recognize responses - rockefeller...

49
Taking Dendritic Cells into Medicine Ralph M. Steinman and Jacques Banchereau The Rockefeller University, New York, NY 10021 and Baylor Institute for Immunology Research, Dallas, TX 75204 Dendritic cells (DCs) orchestrate a repertoire of immune responses that bring about resistance to infection and silencing or tolerance to self. In the settings of infection and cancer, microbes and tumors can exploit DCs to evade immunity, but DCs also can generate resistance, a capacity that is readily enhanced with DC targeted vaccines. During allergy, autoimmunity and transplant rejection, DCs instigate unwanted responses that cause disease, but again DCs can be harnessed to silence these conditions with novel therapies. Here we present some medical implications of DC biology that account for illness and provide opportunities for prevention and therapy. Immunology is a major force in medicine. It is needed to understand how prevalent diseases (Fig. 1) come about and how to develop preventions and treatments. This broad reach of the immune system reflects its two functions: to recognize diverse substances termed antigens, and to generate many qualitatively distinct responses. Dendritic cells (DCs), named for their probing, tree-like or dendritic shapes (Gr. dendron, tree) 1,2 (Fig. 1), are pivotal for both: recognition of a universe of antigens and control of an array of responses. 1

Upload: others

Post on 17-Jun-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

Taking Dendritic Cells into Medicine

Ralph M. Steinman and Jacques Banchereau

The Rockefeller University, New York, NY 10021 and Baylor Institute for Immunology

Research, Dallas, TX 75204

Dendritic cells (DCs) orchestrate a repertoire of immune responses that bring about

resistance to infection and silencing or tolerance to self. In the settings of infection and

cancer, microbes and tumors can exploit DCs to evade immunity, but DCs also can

generate resistance, a capacity that is readily enhanced with DC targeted vaccines. During

allergy, autoimmunity and transplant rejection, DCs instigate unwanted responses that

cause disease, but again DCs can be harnessed to silence these conditions with novel

therapies. Here we present some medical implications of DC biology that account for

illness and provide opportunities for prevention and therapy.

Immunology is a major force in medicine. It is needed to understand how prevalent diseases

(Fig. 1) come about and how to develop preventions and treatments. This broad reach of the

immune system reflects its two functions: to recognize diverse substances termed antigens, and

to generate many qualitatively distinct responses. Dendritic cells (DCs), named for their probing,

tree-like or dendritic shapes (Gr. dendron, tree)1,2 (Fig. 1), are pivotal for both: recognition of a

universe of antigens and control of an array of responses.

1

Page 2: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

Previously, we reviewed some biological features of DCs3,4. Here we illustrate medical

implications of DCs, which control a spectrum of innate and adaptive responses. Innate

immunity encompasses many rapid reactions to infection and other challenges5-9. Adaptive

immunity, in contrast, is learned or acquired more slowly, in days to weeks; it has two hallmarks:

exquisite specificity for antigens, and a durable memory to develop improved function upon re-

exposure to antigen. Adaptive responses are either immunogenic, providing resistance in

infection and cancer, or tolerogenic, leading to silencing as is desirable in transplantation,

autoimmunity and allergy. To date, the successes of immunology in the clinic have largely been

based on antibodies made by B cells, but T cell-mediated immunity, which has enormous yet

untapped therapeutic potential will be stressed here.

DCs are specialized to capture and process antigens in vivo10-14, converting proteins to

peptides that are presented on MHC molecules and recognized by T cells. DCs also migrate to T

cell areas of lymphoid organs where the two cell types interact to bring about clonal selection15-17.

The DC system is thus designed to harness the recognition repertoire of T cells, consisting of

billions of different lymphocytes, each with a distinct but randomly arranged antigen receptor18.

This repertoire in turn represents a virtually infinite “drug library” for specific therapies that

increase or decrease T cell function.

Following clonal selection, DCs control many T cell responses. Antigen-selected T cells

undergo extensive expansion, a thousand fold or more as a result of division at a rate as high as

2-3 cell cycles a day19,20. Clones of lymphocytes are also subject to silencing or tolerance by so-

called “tolerogenic DCs”21-23, which either eliminate (delete)19,20,23-26 or block (suppress) T

cells27-39. If deletion is avoided, the clone undergoes differentiation to bring about an array of

potential helper, killer, and suppressive activities. For example, under the control of DCs, helper

T cells acquire the capacity to produce powerful cytokines like interferon-γ to activate

macrophages to resist infection with facultative and obligate intracellular microbes (Th1 cells)40-

44, or IL-4, 5 and 13 to mobilize white cells that resist helminths (Th2 cells)45,46, or IL-17 to

mobilize phagocytes at body surfaces to resist extracellular bacilli (Th17 cells)47-51.

Alternatively, DCs can guide T cells to become suppressive by making IL-10 (Tr1 cells)28,30,52,53

or by differentiating into foxp3+ cells36,37,39. Finally, DCs induce the T cell clone to acquire

memory, allowing it to persist for prolonged periods and to respond rapidly to a repeated

exposure to antigen54-58.

2

Page 3: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

There are four specialized features (Fig 2) of DCs that contribute to their capacity to

control T cell recognition and responsiveness and in turn, either prevent or generate disease.

Recent advances in DC biology are outlined in Boxes 1-4. Briefly, 1) DCs are positioned to

capture disease-causing antigens and to present these to lymphocytes in lymphoid organs12,59-64,

the sites for the generation of immunity and tolerance. 2) DCs have an endocytic system that is

dedicated to antigen capture and processing, creating ligands for different classes of

lymphocytes13,14. 3) DCs differentiate or mature in response to a spectrum of stimuli14,65,66

(Table 1), allowing them to bring about innate and adaptive responses that are potent and

qualitatively matched to the disease causing agent. 4) DCs are comprised of subsets that differ

from one another in terms of location, antigen presentation, maturation and development67-72.

Location (Box 1) and antigen presentation (Box 2) allow DCs to efficiently select specific clones

from the diverse recognition repertoire. Maturation (Box 3) and subsets (Box 4) allow DCs to

control the diverse response repertoire of T cells and other classes of lymphocytes such as B cells

and NK cells.

3

Page 4: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

Dendritic cells in infectious diseases

DCs induce resistance to infection. When microbial antigens are injected in association with

DCs into mice, the animals acquire adaptive immunity to Borrelia burgdorfei73, chlamydiae74,

Leishmania major75-77, fungi78,79, toxoplasma gondii80-82, malaria83, and HIV84,85. Conversely,

DC depletion reduces defenses to viruses like CMV86,87, HSV-288 and LCMV89. In humans, a

lack of circulating DCs during bacterial sepsis and dengue virus infection is associated with a

poor prognosis90,91.

A key concept is that DCs mature in distinct ways in response to different microbial

components, thereby launching alternative versions of host immunity. The microbial ligands act

on pattern recognition receptors5,7,8 including externally disposed toll like receptors6,44,92-94 and

lectins95,96, and the cytoplasmic NOD/NALP family 97-99, RIG-I and mda5 molecules100. These

pattern recognition receptors can function synergistically44,101-104.

In contrast, several microbes have the capacity to actively block DC maturation, e.g.,

Coxiella burnetti105, Salmonella typhi106, Anthrax lethal factor protein107, Plasmodia108, a M.

ulcerans mycolactone109 and viruses like vaccinia110, herpes simplex111,112, HIV113,114, CMV115-

118, varicella zoster119, HCV120,121, Ebola/Marburg/Lassa fever122,123, and measles124. An

interesting exception is the effective attenuated yellow fever virus vaccine that may work by

infecting and maturing DCs, allowing for antigen presentation to T cells125,126.

Furthermore, pathogens can alter other levels of DC physiology to evade an immune

response. For example, the agents of plague, Yersinia pestis, and typhoid fever, Salmonella typhi,

selectively inject toxins into phagocytes, including DCs, and destroy the cells required for innate

and adaptive protection127,128. Influenza, measles and HSV-2 can induce apoptotic cell death in

DCs112,124,129-131. With some viruses, cell death occurs through the intermediate formation of

giant cells or syncytia132-136. CMV, herpes and M. tuberculosis inhibit the migration step of DC

function by blocking expression of CCR7137-139, a chemokine receptor that guides DCs into

lymphatic vessels and onwards to lymphoid tissues15,17(Box 1, next page).

4

Page 5: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

Box 1, Location of DCs. DCs are a uniquely positioned, prime target for disease relevant stimuli. DCs are abundant at body surfaces like skin, pharynx, upper esophagus, vagina, ectocervix, and anus419, and at so called internal or mucosal surfaces, such as the respiratory and gastro-intestinal systems61,62,364,420-423. DCs actually extend their processes through the tight junctions of epithelia, which likely involves DC expression of the tight junction proteins claudens and occludens, without altering epithelial barrier function424. This increases DC capture of antigens from the environment61,163 even when there is no overt infection or inflammation, probably allowing for the silencing of the immune system to harmless environmental antigens26,363.

DCs at body surfaces can function locally, e.g., to convert vitamins A and D to active retinoic acid and 1,25(OH)2D3. One consequence of the overlooked metabolic capacities of DCs is to increase the homing of immune cells to that mucosal surface425-

427 and, in the case of retinoic acid, to help DCs differentiate suppressor T cells that block autoimmune and inflammatory conditions428-431.

After leaving peripheral tissues, DCs migrate with environmental, self, and microbial antigens to lymphoid organs, a process that is guided by chemokines15,17 and can be enhanced by vaccination432,433.

DCs have now been studied in intact lymphoid tissues without the need for cell isolation. These are the sites where immune resistance and tolerance are initiated. The DCs create a labyrinthine system within T cell areas, while probing the environment through the continuous formation and retraction of processes63 and displaying antigens and other stimuli needed to initiate responses by appropriate clones of specific T cells434-438. New research reveals interactions of DCs in lymphoid tissues with other major classes of lymphocytes, B cells423,439-448 and NK cells257,260,261,449.

Microbes also can alter the function of DCs so that they switch T cell responses from

protective Th1 to nonprotective Th2, as in infections with Aspergillus fumigatus140,141 malaria142,

and hepatitis C143, or to IL-10 production in the case of Bordetella pertussis144.

At this time, there are no therapies that try to interrupt the microbial immune evasion

pathways that are summarized above.

Several microbes additionally can exploit DCs for purposes of replication and spread in

the infected host. The lectin DC-SIGN/CD209 is used by dengue virus145,146 and Ebola virus147

to infect DCs. In the case of HIV-1, CMV, and Ebola virus, the lectin additionally sequesters

virus within DCs, which later transmit infectious virus to other targets like T cells147-153. DCs are

5

Page 6: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

also implicated in the spread of varicella zoster154, measles virus155, poliovirus156, Aspergillus

fumigatus157,158, LCMV159, Toxoplasma gondii 160, prions161, Listeria monocytogenes 162, and

Bacillus anthracis spores163.

Box 2, Antigen presentation by DCs. DCs contain a specialized endocytic system450, having many uptake receptors that deliver antigen to processing compartments14,68,451-

458. DCs then present peptides from processed proteins to CD4+ and CD8+ T cells454,459, self and microbial glycolipids to NKT cells460-462, and native antigens to B cells440,463,464.

Many uptake receptors on DCs are lectins with carbohydrate recognition capacity95,96. Some, like DEC-205/CD205 and mannose receptor/CD206179,312, are type I transmembrane proteins with multiple contiguous lectin domains451. A majority to date, such as DC-SIGN/CD209465, Langerin/CD207466, ASGPR467, LOX-1178 and DCIR68,468, are type II proteins with a single external lectin domain. Among these, CD209 has attracted wide interest because it binds a number of microbes including HIV148, dengue145, cytomegalovirus150, mycobacteria469,470 and candida471, and because it can hinder DC maturation and contribute to immune evasion472-474. Another lectin, Langerin/CD207, is reported to degrade HIV in DCs, thereby reducing transmission to T cells475.

Early literature on DCs took antigen uptake for granted, without realizing that ligation of uptake receptors increases the efficiency with which antigens are delivered to the immune system in vivo by ~100 fold19,56,176,476. Targeting vaccines to these DC receptors should significantly improve the efficacy of T cell mobilizing vaccines.

Following uptake of antigen, DCs are able to “cross present” antigens on MHC I to elicit CD8+ killer T cells235,263,307,476-482. During cross presentation, nonreplicating protein antigens are internalized and somehow gain access to the cytoplasm before being processed by the proteasome for peptide presentation on MHC class I. Critical steps seem to occur from less acidic compartments456,458. Cross presentation allows DCs to induce CD8+ T cell responses to immune complexes, nonreplicating forms of microbes and vaccines, and dying cells205,206,287,299,482,483.

For dying cells60,242,243,390,484, DC receptors for uptake remain to be defined in situ. Since cell death accompanies infection420, cancer246,248, transplantation390,391, the normal turnover of self tissues485, and some viral vaccines486, the uptake of dying cells is a starting point for DCs to capture antigens in many clinical settings. Fc receptors, which recognize antigen-antibody complexes, mediate antigen uptake and both activating and inhibitory signals in DCs204,205,207,487,488. For example, if inhibitory receptors are blocked, the binding of tumor cells coated with antibody leads to improved presentation of tumor antigens and production of the immune stimulating cytokine IL-12207,209.

6

Page 7: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

All these consequences of the microbial-DC interaction have been analyzed on myeloid

DCs, mainly in tissue culture but not in patients. We expect that microbes also evade and exploit

plasmacytoid DCs, which also help to resist pathogens164 and are diminished in the blood during

infections with HIV165-169, HTLV-1170 and RSV171.

Box 3, Maturation of DCs. DCs differentiate or mature in distinct ways in response to a spectrum of environmental and endogenous stimuli (Table 1). The maturation pathway then helps to designate which lymphocyte functions will be induced and which products will be made by both DCs and lymphocytes66,489-492.

In the steady state in the overt absence of maturation stimuli, DCs can induce tolerance when they capture self and environmental antigens19,23,363. Maintenance of tolerance can require PD-L125 as well as fas493 on DCs. Upon infection or other causes of maturation, the DCs redeploy, but it is still not understood why DC differentiation is so rapid and extensive. One factor may be the high levels of required NF-κB family proteins494,495 with each family member (s) able to control different DC responses496.

Maturing DCs can induce different types of CD4+ T cells (see text), such as Th192,93,497, Th245, or Th1748,51,498-502 to increase resistance. Other stimuli can yield “tolerogenic” DCs, which induce Tr1 and foxp3+ T regs28,33,35-37,52,356,361,503-509 to silence immunity. Maturing DCs also express more IL-15 and activate inflammation and NK cells in vivo261,510

When DCs mature in response to microbial products, expression of hundreds of genes is altered511,512, leading to synthesis of cytokines, e.g., IL-12 and type I interferons that enhance innate and adaptive resistance86,513-519. While cytokines in turn can induce some components of DC maturation, the DCs that directly interact with microbial ligands are the immunologically more active ones520,521. The types of cytokines are influenced by the DC subset and the mode of DC activation522-524. Several chemokines are also secreted in groups at a time525, attracting different cells in succession to the site of DC maturation: phagocytes, memory lymphocytes, and naïve T cells526. Numerous mechanisms also dampen the DC response to microbial products185,527.

Maturation regulates antigen processing by lowering the pH of endocytic vacuoles, activating proteolysis, and transporting peptide-MHC complexes to the cell surface453,454,528. Importantly, maturing DCs remodel their surface, typically expressing many membrane associated costimulatory molecules243,520. These include members of the B7, TNF and Notch families529-532.

continued, next page

7

Page 8: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

Box 3, continued. A critical unknown is to define the changes in DCs that link innate to adaptive immunity in vivo533. The standard view that a combination of MHC-peptide (“signal one”) and high B7-2/CD86 is sufficient to drive T cell immunity is oversimplified. To influence T cell differentiation, DCs additionally need to produce cytokines like IL-12 and type I interferons, or membrane associated TNF-family receptors, like CD40243and lymphotoxin receptors534 and TNF family members, like CD70 to induce Th1176,535 and OX40L to induce Th2536.

To counteract these mechanisms for pathogenesis of infectious disease, DCs are now

being considered in the design of vaccines to prevent and treat infection by enhancing

immunogenesis. A new concept is to deliver vaccine antigens to specific receptors on DCs (Box

2), along with stimuli to control DC maturation. For example, microbial proteins are genetically

engineered into anti-receptor monoclonal antibodies, which then quickly and selectively target to

large numbers of DCs within intact lymphoid tissues19,56,172. The CD205 receptor on DCs in

human lymphoid tissues173,174 delivers antigen for processing onto both MHC class I and II,

increasing presentation efficiency >100 fold relative to nontargeted antigen. Th1 responses,

considered valuable for protection against many intracellular pathogens and tumors, also are

induced when antigens from HIV, malaria, Leishmania and tumors are targeted to maturing

CD205+ DCs56,172,175,176. Significantly, these responses are broad, i.e., capable of recognizing

many peptides from a given microbial protein and in several MHC haplotypes56,177, and the

responses take place at mucosal surfaces, both key criteria in vaccine design. Other potential DC

targets are being addressed including LOX-1178, MMR179,180, DCIR68, DC-SIGN181, toxin

receptors182,183 and CD40184.

Beyond the value of antigen targeting, there is a need to mature DCs during vaccination

in a way that is appropriate to the pathogen at hand. Here, one needs to define correlates of

immunogenicity in vivo, i.e., what specific changes in DCs are required to generate protective

responses from the response repertoire?

Dendritic cells in cancer

DCs are found in tumors in mice185 and patients186. Yet tumors suppress immunity, especially

locally and by many pathways187-189. Tumors express cytokines, like IL6190,vascular endothelial

8

Page 9: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

growth factor191 and IL-10192, which suppress DCs through STAT3 signaling193,194. Tumors may

condition local DCs to form suppressive T cells, such as foxp3+195 and IL-13 producing CD4+

T196 and NKT197 cells. DCs even support the clonogenic growth of tumors in multiple

myeloma198. Therefore, like some infections, cancers have ways to evade and exploit DCs199.

Nevertheless, immunology is providing treatments for cancer, mainly in the form of

monoclonal antibodies200,201. Antibodies, by virtue of their antigen binding variable Fab regions,

can block critical functions on cancer cells202; and by virtue of their constant or Fc regions,

antibodies can mobilize an attack by other Fc receptor (FcR) bearing cells such as innate

phagocytes and NK cells203,204. Antibodies mediate the uptake and processing of tumor cells by

DCs and also can trigger DC maturation (Table 1)205-210. This portends the design of antibodies

that harness select FcRs to induce better innate and adaptive anti-tumor immunity.

More emphasis is needed on cell mediated immunity that also has a clear capacity to

resist cancer211,212. In phase I clinical research, adoptive transfer of killer T cells leads to

regressions of melanoma and other cancers213-216. There is also a major survival benefit from

allogeneic bone marrow transplantation, in which lymphocytes from the marrow donor resist

leukemia and other hematologic malignancies (the “graft vs. leukemia” reaction)217-220. In

multiple myeloma, T cells that recognize glycolipids and peptides in the tumor are found in a

premalignant stage of disease but not when the tumor grows out of control221,222. In

paraneoplastic diseases, T cells respond to antigens shared by the tumor and the nervous system,

and likely resist the tumor but at the same time cause severe neurologic sequelae223.

Examination of colorectal cancers indicates that the presence of a Th1 type immune response in

the tumor correlates with a better prognosis224,225. On the other hand, immune suppression

predisposes to higher frequencies of several cancers226. All of the above examples imply a role

for immune surveillance by T cells against human cancer, as is also seen in mice227,228.

DCs can be marshaled for the prevention and treatment of cancer for the following

reasons229. 1) Tumors are replete with potential antigens230-232, and they can become

immunogenic when presented by DCs229,233-238. This means that the immune attack on cancer

can be broad enough to encompass multiple targets including mutant proteins expressed by the

cancer, and not just one target, where the latter favors immune escape. 2) Likewise, DCs can

activate and expand the different arms of cell-mediated resistance such as NK239,240, NKT221,241-

244, γδ T245 and αβ T227,246-248 cells, each of which recognizes different alterations in cancer cells.

9

Page 10: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

3) DCs in systemic lymphoid organs and DCs generated ex vivo from progenitors in blood likely

retain their immunizing capacities in cancer patients43,247,249-254. Therefore one can test whether

the specific features of DCs (Fig. 2) can be harnessed either to generate therapeutic immunity or

to prevent cancer during premalignant or minimal residual disease stages255,256.

Two DC based immune therapy approaches are currently available. In one, DCs are

generated ex vivo, loaded with tumor antigens, and reinjected to take advantage of the ability of

DCs to migrate to the T cell areas of lymphoid organs to induce strong T cell and perhaps NK

immunity257-261(Fig. 3). Less explored is the evidence that DCs themselves can acquire killer

activity for human tumors and express granzyme and perforin killer molecules262. In another

strategy, tumor cells or tumor antigens are targeted directly to DCs in the T cell regions (Fig 3),

e.g., within monoclonal anti-DC antibodies as discussed above.

Already a large number of phase I studies in humans have used the first approach in

which DCs are generated from precursors ex vivo. These early trials have only occasionally

yielded significant tumor regressions263-265, and there are no studies yet showing improved

survival.

10

Page 11: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

The current deficiencies in DC therapy need to be addressed in patients to deal with

significant scientific and other obstacles. First, DC vaccines are being tried only in late stage

cancer patients who are immunosuppressed as a result of extensive radiation, chemotherapy

and/or large tumor burdens. Why limit research to these patients, especially when DC therapy is

nontoxic, including in the few patients who have experienced major tumor regressions263-265?

Second, most injected DCs remain at the injection site. Only a few migrate to the draining

lymphoid tissue, ~ 1 % in mice and humans266,267. Research in patients will be required to

overcome this limitation. One example is to condition the injected site with cytokines like TNF,

so that the injected DCs receive needed cues for migration into lymphatics268. Third, more

emphasis is needed on DC quality, by testing DCs for their capacity to induce helper and killer T

cells with a high avidity for tumor antigens but few T regs269-271. Fourth, many initial vaccine

studies used DCs charged with one or a few tumor antigens, whereas the potential of DCs rests

with their still untapped capacity to elicit a strong and broad immune attack to lessen the chances

of tumor escape. Fifth, vaccine studies need to be accompanied by in depth immune monitoring

to define assays for protective lymphocytes whose presence correlates with tumor regression

and/or improved survival. For example in HIV vaccines, it is thought that protective T cells will

produce several and higher levels of cytokines like IFN-γ, IL-2, and TNF-α272,273; express low

levels of the PD-1 regulatory protein274-276; and have high functional avidity for antigen277. Sixth,

DC based immune therapies require a coordinated research effort to generate DCs for clinical use

and to help investigators systematically evaluate the many variables pertinent to efficacy278.

Even with these limitations, early trials with ex vivo DCs have in patients with advanced

cancer expanded T cells that recognize multiple tumor antigens279 and make the protective

cytokine, IFN-γ43,280,281. These observations provide a starting point to use DCs in a more

concerted way to study their immune capacities in cancer patients, and to combine immunization

with other therapeutic modalities282. Chemotherapy283-287, radiotherapy288, and removal of tumor

regulatory mechanisms, e.g., CTLA4289,290, B7-H4291,292, IL-10293,294, TGF-β195,295, IL-13196,197,

could be more specific and less toxic if combined with DC-based therapies.

The ex vivo technique is being developed so that it is easier to carry out, including in

collaborative studies296-298. This approach allows the basic features of DCs (Fig. 2) to be

controlled, e.g., to load them appropriately with multiple antigens including tumor

11

Page 12: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

cells263,265,299,300 tumor RNA301, or potentially cancer stem cells302; to select from many options

for DC maturation300,303,304(Box 3); and to specify the DC subset for clinical use 279(Box 4).

Box 4, Subsets of DCs. There are several types of DCs, each with distinct markers and functions69. Plasmacytoid DCs, so named because of cytologic similarities to antibody-producing plasma cells537-539, can be involved in tolerance in their immature state540. For maturation, these DCs selectively express activating FcR209,210 as well as TLR7 and TLR9328,457,541-544. When immune complexes containing DNA and RNA are bound and ingested, they signal potentially pathologic levels of type I interferons323,327-

329,331. Other DCs, termed “myeloid,” also produce type I interferons515,545,546.

Myeloid DCs in different tissues can be further subdivided based on expression of certain markers and functions. In the case of skin, DCs in the epidermis (Langerhans cells) express Langerin/CD207547 and DEC-205/CD205, and induce strong killer T cell responses548; some DCs in the dermis express DC-SIGN/CD209 and mannose receptor/CD206549 and can activate antibody forming B cells550. Distinct skin DC populations also migrate to different areas of the draining lymph nodes551,552. Thus the outcome of skin vaccination may depend on the lectin and DC subset that pick up the vaccine.

In mouse spleen, a DC subset expresses DEC-205 and is particularly efficient for the cross presentation of antigens on MHC I60,67,68,478,481,553, including tumor cells248, and also for the induction of IFN-γ producing, Th1 helper T cells41,42,176. A second subset expresses DCIR2 and other lectins and is more efficient at processing antigens for presentation on MHC II68. In lymph nodes draining mucosal tissues, the presence of an integrin, CD103, distinguishes a DC subset that cross presents on MHC class I554 but also synthesizes retinoic acid for the differentiation of foxp3+ T reg428-430.

DC subsets communicate with each other, e.g., plasmacytoid DCs produce interferons and membrane bound costimulators that recruit other DCs to participate in immunity555,556. The production of most DC subsets is controlled in the steady state by the cytokine, flt-3 ligand557,558, whereas during inflammation and infection, another cytokine GM-CSF mobilizes increased numbers of monocyte-derived DCs559-564. In the steady state, DCs in lymphoid tissues emanate from marrow progenitors in the blood565,566 but not monocytes567,568, although monocytes give rise to DCs in some nonlymphoid tissues in the steady state568 and in many sites during inflammation69.

A second DC-based strategy in cancer therapy would be to mobilize DCs directly within

the patient, either within the tumor or within lymphoid organs. In mice, irradiated tumor cells

12

Page 13: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

injected intravenously are taken up by DCs60,248. Alternatively, one can inject irradiated tumor

cells that are transduced to express GM-CSF, which recruits the patient’s DCs to capture and

present tumor antigens305-308. A new approach to the maturation of antigen capturing DCs in

vivo uses innate NKT cells that recognize glycolipids309. The activated NKT cells mimic the

effects of a combination of TLR and CD40 ligation on DCs and elicit long lasting, protective

CD4+ and CD8+ T cell resistance248,310,311. Another example is to selectively target cancer

antigens within monoclonal antibodies to receptors on DCs175,312, thus ensuring the delivery of

antigen to large numbers of DCs in the T cell areas of lymphoid organs, as opposed to the small

numbers that successfully home from an injection site to these areas (Fig. 3).

Unfortunately, DC-mediated immunization against cancer is still an underdeveloped field.

Yet this approach can exploit the patient’s responses to cancer to produce more specific, less

toxic, broader, and longer lasting therapies. There is an important unmet need to expand and

coordinate research on DC based therapies against proliferating and initiating cancer cells.

Dendritic cells in autoimmunity

Inappropriate responses to self constituents, in select genetic backgrounds, can lead to chronic

inflammatory conditions, termed autoimmune diseases. DCs bearing self antigens are able to

induce autoreactive T cells in mouse models of multiple sclerosis313, cardiomyopathy314,315, and

systemic lupus erythematosus (SLE)316.

It is increasingly appreciated that a pivotal step leading to human autoimmunity is an

overproduction of a particular cytokine(s)317-320, and subsets of DCs can be a major source. For

example:

• TNFα is a key cytokine in rheumatoid arthritis and other diseases, like psoriasis,

and TNFα blockade is a powerful therapy for many patients321. In psoriasis, a

major source of TNFα is a DC subset infiltrating the affected skin322.

• Lupus erythematosus, a systemic disease in which antibodies are formed against

self constituents especially nucleoproteins, is accompanied by what is termed an

interferon signature in white blood cells323. Type I interferon is made in large

amounts by plasmacytoid DCs324,325, which infiltrate the skin lesions of SLE326.

Viral nucleic acids, as well as self nucleoproteins internalized in the form of

immune complexes, trigger TLR7 and TLR9 leading to type I interferon

13

Page 14: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

production327-329. Moreover, the DNA binding protein and cytokine, HMGB1,

can deliver self DNA to TLR9 and ligate the RAGE receptor on DCs330.

Interferon in turn drives the mobilization of activated granulocytes323 and

differentiates monocytes into DCs, which may present dying cells in an

immunogenic rather than tolerogenic manner331. Ligation of activating FcR by

the immune complexes that are formed during autoimmunity also drives DC

maturation209,210,332.

• Type I interferon and plasmacytoid DCs are also proposed to be pathogenic in

such other diseases as psoriasis333, dermatomyositis334, and Sjogren’s syndrome335.

• IL-23336 is a cytokine that drives disease, including psoriasis and inflammatory

bowel disease337-341, and DCs are major IL-23 producers342-344.

Despite their role in inducing autoimmunity, DCs are also relevant to the therapy of these

diseases. 1) Various treatments especially glucocorticosteroids can reduce DC numbers and

functions323,345-352. 2) Since DCs seem to be a source of pathogenic cytokines, the stimuli for

cytokine production need to be identified and obviated. 3) One potential antigen-specific

strategy relates to the capacity of DCs to expand and induce T cells that suppress immunity,

usually termed “T regs”31,33,37,353-358. T regs can suppress other DCs that present disease

producing antigens359,360. In the case of a spontaneous model of autoimmunity, diabetes in NOD

mice, T regs that recognize antigens in insulin producing β cells can be generated by DCs and

provide a therapeutic benefit even after the onset of disease36,361.

Dendritic cells in allergy

DCs normally tolerize the immune system upon exposure to harmless environmental

antigens22,34,362,363, including the pollens, dust mites, and foods that cause respiratory and

intestinal allergy. The DCs at mucosal surfaces capture proteins364 and silence the corresponding

T cells, possibly by expressing ICOS-L, which in turn differentiates antigen-specific, IL-10

producing T regs365,366. In allergy, instead of this immune tolerance, CD4+ helper T cells

develop along the atopic Th2 pathway and yield interleukins (IL-4, 5, 13) responsible for many

aspects of disease, such as the formation of IgE type antibodies. DCs both initiate the formation

of proallergic Th2 T cells and boost symptom-producing responses in already allergic mice367-371.

14

Page 15: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

How might the normal tolerizing roles of DCs be altered during allergy? Cytokines may

again be pivotal, e.g., recent studies reveal a role for thymic stromal lymphopoietin (TSLP)372 in

atopic dermatitis46,373 and asthma374. TSLP matures human DCs from blood46 and skin375 to

elicit unusual Th2 cells that secrete not only IL-4, 5, and 13 but also high levels of TNF. TSLP-

treated mature DCs express OX40L, a TNF family member. OX40L can instruct T cells to

develop a Th2 type of memory along with the production of allergic mediators like prostaglandin

D257,376.

Several new directions for allergy research involve DCs. 1) DC maturation by TSLP

needs to be understood and blocked; likewise for TSLP production by epithelial cells, which is

controlled by retinoid receptors377 as well as IL-1 and other inflammatory cytokines378. 2) The

substances that cause allergy may modify DC function directly by blocking IL-12 production by

DCs and in turn favoring formation of allergic Th2 cells379. 3) In humans, new synthetic

oligonucleotides can suppress the presentation of allergens by DCs to Th2 cells380; these

compounds have shown promise in treating ragweed induced allergic rhinitis381. 4) DCs might

be targeted to allow for formation of allergen-specific T regs, which can treat allergy as observed

in mouse models382,383. 5) New drugs can interfere with the pro-allergic functions of DCs and

treat experimental asthma384. The sphingosine-1-phosphate receptor agonist FTY720 blocks the

migration of DCs from lung to lymph nodes where the DCs could present antigens to disease-

causing Th2 cells385, while agonists for the D prostanoid 1 receptor condition DCs to induce

disease-reducing T regs386.

Dendritic cells in transplantation

DCs play a key role in the outcome of organ and hematopoietic transplantation. DCs in grafted

organs mature and migrate into the recipient387-389, where they stimulate alloreactive T cells that

bring about graft rejection. Recipient DCs also can capture portions of the graft390,391 and elicit

organ rejection. Interestingly, suppressive drugs in current clinical use may act on the rejection-

inducing DCs as well as the rejecting T cells392-396. In hematopoietic transplantation, recipient

DCs are key initiators of T cell-induced graft versus host reactions397,398. Mechanisms have yet

to be pinpointed to explain the maturation and migration of DCs that accompany transplantation.

It seems reasonable to propose that strategies to block these DC functions during

transplantation will promote acceptance399. However, recent discoveries show that DCs in

15

Page 16: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

grafted tissues can regenerate locally400, thus providing a long term source of antigen to stimulate

rejection398. In this light, alternative pathways to use DCs to induce transplantation tolerance are

being assessed401-403. One involves the activation of recipient NK cells, which reject donor DCs

in tissue culture models404,405 and in vivo406. Another pathway is to induce DCs to become

tolerogenic389, e.g., to express the tolerogenic ILT3 molecule407 or to induce graft-specific foxp3+

T reg to suppress graft-rejecting T cells35,391. In hematopoietic cell transplantation, strategies

that lead to recipient DC depletion are currently being tested in the clinic.

Extending DC biology into medicine.

Immunology, including T cell mediated immunity, has a central role in understanding how

disease develops and in designing new treatments (Fig. 4). Here we suggest therapies aimed at

the upstream events initiated by DCs. DCs can either intensify or subdue T cell responses,

depending on whether resistance or tolerance needs to be increased.

Considerable evidence from studies in mice, which predominates in this review, shows

how DCs act in a disease specific manner. Increased emphasis on the antigens that elicit disease

should provide medical approaches that are longer lasting and less subject to side effects than

current antigen-nonspecific ones. To repeat, DC-based analyses and therapies emphasize initial

events in complex disease cascades, but more patient-based research is needed to take DCs into

medicine.

16

Page 17: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

The scientific rationale is that a patient’s response to disease involves not only antigens

and lymphocytes but also DCs that control antigen presentation and clonal selection (the immune

recognition repertoire) as well as lymphocyte growth, differentiation and memory (the immune

response repertoire). Each feature of DCs in fig. 2 is an intricate area for research, whose

findings are facilitating a deeper analysis of disease and how to interrupt its progress. However,

major gaps exist to take this science into medicine.

In mice, the current research emphasis has been on model antigens, isolated DCs, and

highly selected populations of T cells with a single transgenic antigen receptor. Instead, research

on DCs needs to be directed to the control of immune responses 1) in situ in intact animals, 2)

within the natural immune repertoire, 3) to clinically relevant antigens, and 4) with new immune

enhancers or adjuvants that act in defined ways on DCs. Even so, the immune systems of mice

and men differ in many aspects408. A demanding subject, but one which will enhance disease

research, is the development of mice with immune systems derived from human sources196,409-416.

More research needs to be done with patients. The patient sets the standards for the

quality of knowledge that is required to understand many aspects of disease and its treatment.

Often the pathogen (tumor, microbe, allergen, stimulus for autoimmune or autoinflammatory

disease) is not easily or completely modeled in mice.

Although scientists and the public both desire “translation” from mice to humans, there is

an underappreciated need to create a sizeable limb of the scientific enterprise that will bring new

methods, concepts, and coordination to the basic study of disease with patients. This need is

illustrated by the paucity of DC-based studies on the immunotherapy of cancer. Basic research

with patients differs substantially from current outcome studies and drug licensing trials, which

test whether existing practices and concepts are clinically effective. In contrast, disease research

yields new ideas and therapies. The biology of DCs is ready to be extended to dissect disease

pathways and to direct its prevention and treatment. Despite the obstacles to research in

patients417,418, physicians and scientists have the knowledge and tools to think systemically about

diseases, plan their therapies, and investigate how human beings respond. DCs are an early

player in disease development and an unavoidable target in the design of treatments.

17

Page 18: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

Acknowledgement. The authors thank Carol Moberg and Michel Nussenzweig for extensive

comments and Judy Adams for assistance with the manuscript. We are grateful to our patients

and colleagues for their many contributions, and to the NIH and several foundations for support.

18

Page 19: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

1 Steinman, R. M. & Cohn, Z. A., Identification of a novel cell type in peripheral lymphoid organs of mice. I. Morphology, quantitation, tissue distribution. J. Exp. Med. 137, 1142 (1973).

2 Steinman, R. M. & Cohn, Z. A., Identification of a novel cell type in peripheral lymphoid organs of mice. II. Functional properties in vitro. J. Exp. Med. 139, 380 (1974).

3 Banchereau, J. & Steinman, R. M., Dendritic cells and the control of immunity. Nature 392, 245 (1998).

4 Banchereau, J. et al., Immunobiology of dendritic cells. Annu. Rev. Immunol. 18, 767 (2000).

5 Janeway, C. A., Jr. & Medzhitov, R., Innate immune recognition. Annu. Rev. Immunol. 20, 197 (2002).

6 Iwasaki, A. & Medzhitov, R., Toll-like receptor control of the adaptive immune responses. Nat. Immunol. 5, 987 (2004).

7 Akira, S., Uematsu, S. &Takeuchi, O., Pathogen recognition and innate immunity. Cell 124, 783 (2006).

8 Beutler, B. et al., Genetic analysis of host resistance: toll-like receptor signaling and immunity at large. Annu. Rev. Immunol. 24, 353 (2006).

9 Lemaitre, B. & Hoffmann, J., The host defense of Drosophila melanogaster. Annu. Rev. Immunol. 25, 697 (2007).

10 Bujdoso, R., Hopkins, J., Dutia, B. M., Young, P. &McConnell, I., Characterization of sheep afferent lymph dendritic cells and their role in antigen carriage. J. Exp. Med. 170, 1285 (1989).

11 Crowley, M., Inaba, K. &Steinman, R. M., Dendritic cells are the principal cells in mouse spleen bearing immunogenic fragments of foreign proteins. J. Exp. Med. 172, 383 (1990).

12 Liu, L. M. & MacPherson, G. G., Antigen acquisition by dendritic cells: intestinal dendritic cells acquire antigen administered orally and can prime naive T cells in vivo. J. Exp. Med. 177, 1299 (1993).

13 Guermonprez, P., Valladeau, J., Zitvogel, L., Thery, C. &Amigorena, S., Antigen presentation and T cell stimulation by dendritic cells. Annu. Rev. Immunol. 20, 621 (2002).

14 Trombetta, E. S. & Mellman, I., Cell biology of antigen processing in vitro and in vivo. Annu. Rev. Immunol. 23, 975 (2005).

15 Cyster, J. G., Chemokines and the homing of dendritic cells to the T cell areas of lymphoid organs. J. Exp. Med. 189, 447 (1999).

16 Itano, A. A. & Jenkins, M. K., Antigen presentation to naive CD4 T cells in the lymph node. Nat. Immunol. 4, 733 (2003).

17 Randolph, G. J., Angeli, V. &Swartz, M. A., Dendritic-cell trafficking to lymph nodes through lymphatic vessels. Nat. Rev. Immunol. 5, 617 (2005).

18 Germain, R. N., An innately interesting decade of research in immunology. Nat. Med. 10, 1307 (2004).

19 Hawiger, D. et al., Dendritic cells induce peripheral T cell unresponsiveness under steady state conditions in vivo. J. Exp. Med. 194, 769 (2001).

20 Bonifaz, L. et al., Efficient targeting of protein antigen to the dendritic cell receptor DEC-205 in the steady state leads to antigen presentation on major histocompatibility

19

Page 20: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

complex class I products and peripheral CD8+ T cell tolerance. J. Exp. Med. 196, 1627 (2002).

21 Steinman, R. M. & Nussenzweig, M. C., Avoiding horror autotoxicus: the importance of dendritic cells in peripheral T cell tolerance. Proc. Natl. Acad. Sci. USA 99, 351 (2002).

22 Moser, M., Dendritic cells in immunity and tolerance-do they display opposite functions? Immunity 19, 5 (2003).

23 Probst, H. C., Lagnel, J., Kollias, G. &van den Broek, M., Inducible transgenic mice reveal resting dendritic cells as potent inducers of CD8+ T cell tolerance. Immunity 18, 713 (2003).

24 Liu, K. et al., Immune tolerance after delivery of dying cells to dendritic cells in situ. J. Exp. Med. 196, 1091 (2002).

25 Probst, H. C., McCoy, K., Okazaki, T., Honjo, T. &van den Broek, M., Resting dendritic cells induce peripheral CD8+ T cell tolerance through PD-1 and CTLA-4. Nat. Immunol. 6, 280 (2005).

26 Worbs, T. et al., Oral tolerance originates in the intestinal immune system and relies on antigen carriage by dendritic cells. J. Exp. Med. 203, 519 (2006).

27 Manavalan, J. S. et al., High expression of ILT3 and ILT4 is a general feature of tolerogenic dendritic cells. Transpl. Immunol. 11, 245 (2003).

28 Levings, M. K. et al., Differentiation of Tr1 cells by immature dendritic cells requires IL-10 but not CD25+CD4+ Tr cells. Blood 105, 1162 (2005).

29 Steinbrink, K. et al., Interleukin-10-treated human dendritic cells induce a melanoma-antigen- specific anergy in CD8(+) T cells resulting in a failure to lyse tumor cells. Blood 93, 1634 (1999).

30 Dhodapkar, M. V. & Steinman, R. M., Antigen-bearing, immature dendritic cells induce peptide-specific, CD8+ regulatory T cells in vivo in humans. Blood 100, 174 (2002).

31 Kretschmer, K. et al., Inducing and expanding regulatory T cell populations by foreign antigen. Nat. Immunol. 6, 1219 (2005).

32 Zinser, E., Lechmann, M., Golka, A., Lutz, M. B. &Steinkasserer, A., Prevention and treatment of experimental autoimmune encephalomyelitis by soluble CD83. J. Exp. Med. 200, 345 (2004).

33 Tarbell, K. V., Yamazaki, S., Olson, K., Toy, P. &Steinman, R. M., CD25+ CD4+ T cells, expanded with dendritic cells presenting a single autoantigenic peptide, suppress autoimmune diabetes. J. Exp. Med. 199, 1467 (2004).

34 Macpherson, A. J. & Smith, K., Mesenteric lymph nodes at the center of immune anatomy. J. Exp. Med. 203, 497 (2006).

35 Yamazaki, S. et al., Effective expansion of alloantigen-specific Foxp3+ CD25+ regulatory T cells by dendritic cells during the mixed leukocyte reaction. Proc. Natl. Acad. Sci. USA 103, 2758 (2006).

36 Luo, X. et al., Dendritic cells with TGF-β1 differentiate naive CD4+CD25- T cells into islet-protective Foxp3+ regulatory T cells. Proc. Natl. Acad. Sci. USA 104, 2821 (2007).

37 Yamazaki, S. et al., Dendritic cells are specialized accessory cells for the induction of Foxp3+ CD4+ regulatory T cells with TGF-β. Blood In Press (2007).

38 Travis, M. A. et al., Loss of integrin αvβ8 on dendritic cells causes autoimmunity and colitis in mice. Nature 449, 361 (2007).

20

Page 21: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

39 Dipaolo, R. J. et al., Autoantigen-specific TGFβ-induced Foxp3+ regulatory T cells prevent autoimmunity by inhibiting dendritic cells from activating autoreactive T cells. J. Immunol. 179, 4685 (2007).

40 Macatonia, S. E. et al., Dendritic cells produce IL-12 and direct the development of Th1 cells from naive CD4+ T cells. J. Immunol. 154, 5071 (1995).

41 Pulendran, B. et al., Distinct dendritic cell subsets differentially regulate the class of immune responses in vivo. Proc. Natl. Acad. Sci. USA 96, 1036 (1999).

42 Maldonado-Lopez, R. et al., CD8α+ and CD8α- subclasses of dendritic cells direct the development of distinct T helper cells in vivo. J. Exp. Med. 189, 587 (1999).

43 Schuler-Thurner, B. et al., Rapid induction of tumor-specific type 1 T helper cells in metastatic melanoma patients by vaccination with mature, cryopreserved, peptide-loaded monocyte-derived dendritic cells. J. Exp. Med. 195, 1279 (2002).

44 Napolitani, G., Rinaldi, A., Bertoni, F., Sallusto, F. &Lanzavecchia, A., Selected Toll-like receptor agonist combinations synergistically trigger a T helper type 1-polarizing program in dendritic cells. Nat. Immunol. 6, 769 (2005).

45 Seder, R. A., Paul, W. E., Davis, M. M. &Fazekas de St.Groth, B., The presence of interleukin 4 during in vitro priming determines the lymphokine-producing potential of CD4+ T cells from T cell receptor transgenic mice. J. Exp. Med. 176, 1091 (1992).

46 Soumelis, V. et al., Human epithelial cells trigger dendritic cell-mediated allergic inflammation by producing TSLP. Nat. Immunol. 3, 673 (2002).

47 Higgins, S. C., Jarnicki, A. G., Lavelle, E. C. &Mills, K. H., TLR4 mediates vaccine-induced protective cellular immunity to Bordetella pertussis: role of IL-17-producing T cells. J. Immunol. 177, 7980 (2006).

48 Bailey, S. L., Schreiner, B., McMahon, E. J. &Miller, S. D., CNS myeloid DCs presenting endogenous myelin peptides 'preferentially' polarize CD4+ TH-17 cells in relapsing EAE. Nat. Immunol. 8, 172 (2007).

49 Kleinschek, M. A. et al., IL-25 regulates Th17 function in autoimmune inflammation. J. Exp. Med. 204, 161 (2007).

50 Leibundgut-Landmann, S. et al., Syk- and CARD9-dependent coupling of innate immunity to the induction of T helper cells that produce interleukin 17. Nat. Immunol. 8, 630 (2007).

51 Denning, T. L., Wang, Y. C., Patel, S. R., Williams, I. R. &Pulendran, B., Lamina propria macrophages and dendritic cells differentially induce regulatory and interleukin 17-producing T cell responses. Nat. Immunol. 8, 1086 (2007).

52 Jonuleit, H., Schmitt, E., Schuler, G., Knop, J. &Enk, A. H., Induction of interleukin 10-producing, nonproliferating CD4+ T cells with regulatory properties by repetitive stimulation with allogeneic immature human dendritic cells. J. Exp. Med. 192, 1213 (2000).

53 Dhodapkar, M. V., Steinman, R. M., Krasovsky, J., Münz, C. &Bhardwaj, N., Antigen specific inhibition of effector T cell function in humans after injection of immature dendritic cells. J. Exp. Med. 193, 233 (2001).

54 Badovinac, V. P., Messingham, K. A., Jabbari, A., Haring, J. S. &Harty, J. T., Accelerated CD8+ T-cell memory and prime-boost response after dendritic-cell vaccination. Nat. Med. 11, 748 (2005).

55 Zammit, D. J., Cauley, L. S., Pham, Q. M. &Lefrancois, L., Dendritic cells maximize the memory CD8 T cell response to infection. Immunity 22, 561 (2005).

21

Page 22: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

56 Trumpfheller, C. et al., Intensified and protective CD4+ T cell immunity at a mucosal surface after a single dose of anti-dendritic cell HIV gag fusion antibody vaccine. J. Exp. Med. 203, 607 (2006).

57 Wang, Y. H. et al., Maintenance and polarization of human TH2 central memory T cells by thymic stromal lymphopoietin-activated dendritic cells. Immunity 24, 827 (2006).

58 Pulendran, B. & Ahmed, R., Translating innate immunity into immunological memory: implications for vaccine development. Cell 124, 849 (2006).

59 Huang, F.-P. et al., A discrete subpopulation of dendritic cells transports apoptotic intestinal epithelial cells to T cell areas of mesenteric lymph nodes. J. Exp. Med. 191, 435 (2000).

60 Iyoda, T. et al., The CD8+ dendritic cell subset selectively endocytoses dying cells in culture and in vivo. J. Exp. Med. 195, 1289 (2002).

61 Niess, J. H. et al., CX3CR1-mediated dendritic cell access to the intestinal lumen and bacterial clearance. Science (New York, N.Y 307, 254 (2005).

62 Chieppa, M., Rescigno, M., Huang, A. Y. C. &Germain, R. N., Dynamic imaging of dendritic cell extension into the small bowel lumen in response to epithelial cell TLR engagement. J. Exp. Med. 203, 2841 (2006).

63 Lindquist, R. L. et al., Visualizing dendritic cell networks in vivo. Nat. Immunol. 5, 1243 (2004).

64 Kabashima, K. et al., Intrinsic lymphotoxin-β receptor requirement for homeostasis of lymphoid tissue dendritic cells. Immunity 22, 439 (2005).

65 Mellman, I. & Steinman, R. M., Dendritic cells: specialized and regulated antigen processing machines. Cell 106, 255 (2001).

66 Reis e Sousa, C., Dendritic cells in a mature age. Nat. Rev. Immunol. 6, 476 (2006). 67 Schnorrer, P. et al., The dominant role of CD8+ dendritic cells in cross-presentation is not

dictated by antigen capture. Proc. Natl. Acad. Sci. USA 103, 10729 (2006). 68 Dudziak, D. et al., Differential antigen processing by dendritic cell subsets in vivo.

Science (New York, N.Y 315, 107 (2007). 69 Shortman, K. & Naik, S. H., Steady-state and inflammatory dendritic-cell development.

Nat. Rev. Immunol. 7, 19 (2007). 70 Wu, L. et al., RelB is essential for the development of myeloid-related CD8α- dendritic

cells but not of lymphoid-related CD8α+ dendritic cells. Immunity 9, 839 (1998). 71 Tamura, T. et al., IFN regulatory factor-4 and -8 govern dendritic cell subset

development and their functional diversity. J. Immunol. 174, 2573 (2005). 72 Caton, M. L., Smith-Raska, M. R. &Reizis, B., Notch-RBP-J signaling controls the

homeostasis of CD8- dendritic cells in the spleen. J. Exp. Med. 204, 1653 (2007). 73 Mbow, M. L., Zeidner, N., Panella, N., Titus, R. G. &Piesman, J., Borrelia burgdorfei-

pulsed dendritic cells induce a protective immune response against tick-transmitted spirochetes. Infect. Immun. 65, 3386 (1997).

74 Su, H. et al., Vaccination against chlamydial genital tract infection following immunization with dendritic cells pulsed ex vivo with non-viable chlamydiae. J. Exp. Med. 188, 809 (1998).

75 Flohe, S. B., Bauer, C., Flohe, S. &Moll, H., Antigen-pulsed epidermal Langerhans cells protect susceptible mice from infection with the intracellular parasite Leishmania major. Eur. J. Immunol. 28, 3800 (1998).

22

Page 23: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

76 Ahuja, S. S. et al., Dendritic cell (DC)-based anti-infective strategies: DCs engineered to secrete IL-12 are a potent vaccine in a murine model of an intracellular infection. J. Immunol. 163, 3890 (1999).

77 Ghosh, M. et al., Dendritic cell-based immunotherapy combined with antimony-based chemotherapy cures established murine visceral leishmaniasis. J. Immunol. 170, 5625 (2003).

78 Bozza, S. et al., A dendritic cell vaccine against invasive aspergillosis in allogeneic hematopoietic transplantation. Blood 102, 3807 (2003).

79 Bozza, S. et al., Dendritic cell-based vaccination against opportunistic fungi. Vaccine 22, 857 (2004).

80 Bourguin, I. et al., Murine dendritic cells pulsed in vitro with Toxoplasma gondii antigens induce protective immunity in vivo. Infect. Immun. 66, 4867 (1998).

81 Dimier-Poisson, I. et al., Protective mucosal Th2 immune response against Toxoplasma gondii by murine mesenteric lymph node dendritic cells. Infect. Immun. 71, 5254 (2003).

82 Liu, C. H. et al., Dendritic cells are essential for in vivo IL-12 production and development of resistance against Toxoplasma gondii infection in mice. J. Immunol. 177, 31 (2006).

83 Pouniotis, D. S. et al., Dendritic cells induce immunity and long-lasting protection against blood-stage malaria despite an in vitro parasite-induced maturation defect. Infect. Immun. 72, 5331 (2004).

84 Gorantla, S. et al., Human dendritic cells transduced with herpes simplex virus amplicons encoding human immunodeficiency virus type 1 (HIV-1) gp120 elicit adaptive immune responses from human cells engrafted into NOD/SCID mice and confer partial protection against HIV-1 challenge. J. Virol. 79, 2124 (2005).

85 Ayash-Rashkovsky, M. et al., Enhanced HIV-1 specific immune response by CpG ODN and HIV-1 immunogen-pulsed dendritic cells confers protection in the Trimera murine model of HIV-1 infection. FASEB J. 19, 1152 (2005).

86 Dalod, M. et al., Interferon α/β and interleukin 12 responses to viral infections: pathways regulating dendritic cell cytokine expression in vivo. J. Exp. Med. 195, 517 (2002).

87 Asselin-Paturel, C., Brizard, G., Pin, J. J., Briere, F. &Trinchieri, G., Mouse strain differences in plasmacytoid dendritic cell frequency and function revealed by a novel monoclonal antibody. J. Immunol. 171, 6466 (2003).

88 Lund, J. M., Linehan, M. M., Iijima, N. &Iwasaki, A., Plasmacytoid dendritic cells provide innate immune protection against mucosal viral infection in situ. J. Immunol. 177, 7510 (2006).

89 Probst, H. C. & van den Broek, M., Priming of CTLs by lymphocytic choriomeningitis virus depends on dendritic cells. J. Immunol. 174, 3920 (2005).

90 Guisset, O. et al., Decrease in circulating dendritic cells predicts fatal outcome in septic shock. Intensive Care Med. 33, 148 (2006).

91 Pichyangkul, S. et al., A blunted blood plasmacytoid dendritic cell response to an acute systemic viral infection is associated with increased disease severity. J. Immunol. 171, 5571 (2003).

92 Pulendran, B., Palucka, K. &Banchereau, J., Sensing pathogens and tuning immune responses. Science (New York, N.Y 293, 253 (2001).

23

Page 24: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

93 Boonstra, A. et al., Flexibility of mouse classical and plasmacytoid-derived dendritic cells in directing T helper type 1 and 2 cell development: dependency on antigen dose and differential Toll-like receptor ligation. J. Exp. Med. 197, 101 (2003).

94 Takeda, K., Kaisho, T. &Akira, S., Toll-like receptors. Annu. Rev. Immunol. 21, 335 (2003).

95 Figdor, C. G., van Kooyk, Y. &Adema, G. J., C-type lectin receptors on dendritic cells and Langerhans cells. Nat. Rev. Immunol. 2, 77 (2002).

96 Robinson, M. J., Sancho, D., Slack, E. C., Leibundgut-Landmann, S. &Sousa, C. R., Myeloid C-type lectins in innate immunity. Nat. Immunol. 7, 1258 (2006).

97 Inohara, N. & Nunez, G., NODs: intracellular proteins involved in inflammation and apoptosis. Nat. Rev. Immunol. 3, 371 (2003).

98 Mariathasan, S. & Monack, D. M., Inflammasome adaptors and sensors: intracellular regulators of infection and inflammation. Nat. Rev. Immunol. 7, 31 (2007).

99 Martinon, F. & Tschopp, J., NLRs join TLRs as innate sensors of pathogens. Trends Immunol 26, 447 (2005).

100 Honda, K., Takaoka, A. &Taniguchi, T., Type I inteferon gene induction by the interferon regulatory factor family of transcription factors. Immunity 25, 349 (2006).

101 Gantner, B. N., Simmons, R. M., Canavera, S. J., Akira, S. &Underhill, D. M., Collaborative induction of inflammatory responses by dectin-1 and toll-like receptor 2. J. Exp. Med. 197, 1107 (2003).

102 Brown, G. D. et al., Dectin-1 mediates the biological effects of β-glucans. J. Exp. Med. 197, 1119 (2003).

103 Gautier, G. et al., A type I interferon autocrine-paracrine loop is involved in Toll-like receptor-induced interleukin-12p70 secretion by dendritic cells. J. Exp. Med. 201, 1435 (2005).

104 Fritz, J. H. et al., Nod1-mediated innate immune recognition of peptidoglycan contributes to the onset of adaptive immunity. Immunity 26, 445 (2007).

105 Shannon, J. G., Howe, D. &Heinzen, R. A., Virulent Coxiella burnetii does not activate human dendritic cells: Role of lipopolysaccharide as a shielding molecule. Proc. Natl. Acad. Sci. USA 102, 8722 (2005).

106 Tobar, J. A. et al., Virulent Salmonella enterica serovar typhimurium evades adaptive immunity by preventing dendritic cells from activating T cells. Infect. Immun. 74, 6438 (2006).

107 Agrawal, A. et al., Impairment of dendritic cells and adaptive immunity by anthrax lethal toxin. Nature 424, 329 (2003).

108 Urban, B. C. et al., Plasmodium falciparum-infected erythrocytes modulate the maturation of dendritic cells. Nature 400, 73 (1999).

109 Coutanceau, E. et al., Selective suppression of dendritic cell functions by Mycobacterium ulcerans toxin mycolactone. J. Exp. Med. 204, 1395 (2007).

110 Engelmayer, J. et al., Vaccinia virus inhibits the maturation of human dendritic cells: a novel mechanism of immune evasion. J. Immunol. 163, 6762 (1999).

111 Salio, M., Cella, M., Suter, M. &Lanzavecchia, A., Inhibition of dendritic cell maturation by herpes simplex virus. Eur. J. Immunol. 29, 3245 (1999).

112 Muller, D. B., Raftery, M. J., Kather, A., Giese, T. &Schonrich, G., Induction of apoptosis and modulation of c-FLIPL and p53 in immature dendritic cells infected with herpes simplex virus. Eur. J. Immunol. 34, 941 (2004).

24

Page 25: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

113 Granelli-Piperno, A., Golebiowska, A., Trumpfheller, C., Siegal, F. P. &Steinman, R. M., HIV-1-infected monocyte-derived dendritic cells do not undergo maturation but can elicit IL-10 production and T cell regulation. Proc. Natl. Acad. Sci. USA 101, 7669 (2004).

114 Majumder, B. et al., Human immunodeficiency virus type 1 Vpr impairs dendritic cell maturation and T-cell activation: implications for viral immune escape. J. Virol. 79, 7990 (2005).

115 Andrews, D. M., Andoniou, C. E., Granucci, F., Ricciardi-Castagnoli, P. &Degli-Esposti, M. A., Infection of dendritic cells by murine cytomegalovirus induces functional paralysis. Nat. Immunol. 2, 1077 (2001).

116 Moutaftsi, M., Mehl, A. M., Borysiewicz, L. K. &Tabi, Z., Human cytomegalovirus inhibits maturation and impairs function of monocyte-derived dendritic cells. Blood 99, 2913 (2002).

117 Senechal, B., Boruchov, A. M., Reagan, J. L., Hart, D. N. &Young, J. W., Infection of mature monocyte-derived dendritic cells with human cytomegalovirus inhibits stimulation of T-cell proliferation via the release of soluble CD83. Blood 103, 4207 (2004).

118 Raftery, M. J. et al., Targeting the function of mature dendritic cells by human cytomegalovirus. A multilayered viral defense strategy. Immunity 15, 997 (2001).

119 Morrow, G., Slobedman, B., Cunningham, A. L. &Abendroth, A., Varicella-zoster virus productively infects mature dendritic cells and alters their immune function. J. Virol. 77, 4950 (2003).

120 Sarobe, P. et al., Hepatitis C virus structural proteins impair dendritic cell maturation and inhibit in vivo induction of cellular immune responses. J. Virol. 77, 10862 (2003).

121 Dolganiuc, A. et al., Hepatitis C virus core and nonstructural protein 3 proteins induce pro- and anti-inflammatory cytokines and inhibit dendritic cell differentiation. J. Immunol. 170, 5615 (2003).

122 Mahanty, S. et al., Impairment of dendritic cells and adaptive immunity by Ebola and Lassa viruses. J. Immunol. 170, 2797 (2003).

123 Bosio, C. M. et al., Ebola and Marburg viruses replicate in monocyte-derived dendritic cells without inducing the production of cytokines and full maturation. J. Infect. Dis. 188, 1630 (2003).

124 Marie, J. C. et al., Mechanism of measles virus-induced suppression of inflammatory immune responses. Immunity 14, 69 (2001).

125 Barba-Spaeth, G., Longman, R. S., Albert, M. L. &Rice, C. M., Live attenuated yellow fever 17D infects human DCs and allows for presentation of endogenous and recombinant T cell epitopes. J. Exp. Med. 202, 1179 (2005).

126 Querec, T. et al., Yellow fever vaccine YF-17D activates multiple dendritic cell subsets via TLR2, 7, 8, and 9 to stimulate polyvalent immunity. J. Exp. Med. 203, 413 (2006).

127 Marketon, M. M., DePaolo, R. W., DeBord, K. L., Jabri, B. &Schneewind, O., Plague bacteria target immune cells during infection. Science (New York, N.Y 309, 1739 (2005).

128 Van Der Velden, A. W., Velasquez, M. &Starnbach, M. N., Salmonella rapidly kill dendritic cells via a caspase-1- dependent mechanism. J. Immunol. 171, 6742 (2003).

129 Servet-Delprat, C. et al., Consequences of Fas-mediated human dendritic cell apoptosis induced by measles virus. J. Virol. 74, 4387 (2000).

130 Jones, C. A. et al., Herpes simplex virus type 2 induces rapid cell death and functional impairment of murine dendritic cells in vitro. J. Virol. 77, 11139 (2003).

25

Page 26: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

131 Siavoshian, S., Abraham, J. D., Thumann, C., Kieny, M. P. &Schuster, C., Hepatitis C virus core, NS3, NS5A, NS5B proteins induce apoptosis in mature dendritic cells. J. Med. Virol. 75, 402 (2005).

132 Grosjean, I. et al., Measles virus infects human dendritic cells and blocks their allostimulatory properties for CD4+ T cells. J. Exp. Med. 186, 801 (1997).

133 Fugier-Vivier, I. et al., Measles virus suppresses cell-mediated immunity by interfering with the survival and functions of dendritic and T cells. J. Exp. Med. 186, 813 (1997).

134 Pope, M. et al., Conjugates of dendritic cells and memory T lymphocytes from skin facilitate productive infection with HIV-1. Cell 78, 389 (1994).

135 Frankel, S. S. et al., Replication of HIV-1 in dendritic cell-derived syncytia at the mucosal surface of the adenoid. Science (New York, N.Y 272, 115 (1996).

136 Frankel, S. S. et al., Active replication of HIV-1 at the lymphoepithelial surface of the tonsil. Am. J. Pathol. 151, 89 (1997).

137 Moutaftsi, M., Brennan, P., Spector, S. A. &Tabi, Z., Impaired lymphoid chemokine-mediated migration due to a block on the chemokine receptor switch in human cytomegalovirus-infected dendritic cells. J. Virol. 78, 3046 (2004).

138 Prechtel, A. T. et al., Infection of mature dendritic cells with herpes simplex virus type 1 dramatically reduces lymphoid chemokine-mediated migration. J. Gen. Virol. 86, 1645 (2005).

139 Khader, S. A. et al., Interleukin 12p40 is required for dendritic cell migration and T cell priming after Mycobacterium tuberculosis infection. J. Exp. Med. 203, 1805 (2006).

140 Benjamim, C. F., Lundy, S. K., Lukacs, N. W., Hogaboam, C. M. &Kunkel, S. L., Reversal of long-term sepsis-induced immunosuppression by dendritic cells. Blood 105, 3588 (2005).

141 d'Ostiani, C. F. et al., Dendritic cells discriminate between yeasts and hyphae of the fungus Candida albicans. Implications for initiation of T helper cell immunity in vitro and in vivo. J. Exp. Med. 191, 1661 (2000).

142 Sponaas, A. M. et al., Malaria infection changes the ability of splenic dendritic cell populations to stimulate antigen-specific T cells. J. Exp. Med. 203, 1427 (2006).

143 Kanto, T. et al., Reduced numbers and impaired ability of myeloid and plasmacytoid dendritic cells to polarize T helper cells in chronic hepatitis C virus infection. J. Infect. Dis. 190, 1919 (2004).

144 McGuirk, P., McCann, C. &Mills, K. H. G., Pathogen-specific T regulatory 1 cells induced in the respiratory tract by a bacterial molecule that stimulates interleukin 10 production by dendritic cells: a novel strategy for evasion of protective T helper type 1 responses by Bordetella pertussis. J. Exp. Med. 195, 221 (2002).

145 Tassaneetrithep, B. et al., DC-SIGN (CD209) mediates dengue virus infection of human dendritic cells. J. Exp. Med. 197, 823 (2003).

146 Navarro-Sanchez, E. et al., Dendritic-cell-specific ICAM3-grabbing non-integrin is essential for the productive infection of human dendritic cells by mosquito-cell-derived dengue viruses. EMBO Rep. 4, 1 (2003).

147 Alvarez, C. P. et al., C-type lectins DC-SIGN and L-SIGN mediate cellular entry by Ebola virus in cis and in trans. J. Virol. 76, 6841 (2002).

148 Geijtenbeek, T. B. H. et al., DC-SIGN, a dendritic cell specific HIV-1 binding protein that enhances TRANS-infection of T cells. Cell 100, 587 (2000).

26

Page 27: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

149 Kwon, D. S., Gregario, G., Bitton, N., Hendrickson, W. A. &Littman, D. R., DC-SIGN mediated internalization of HIV is required for trans-enhancement of T cell infection. Immunity 16, 135 (2002).

150 Halary, F. et al., Human cytomegalovirus binding to DC-SIGN is required for dendritic cell infection and target cell trans-infection. Immunity 17, 653 (2002).

151 Gurney, K. B. et al., Binding and transfer of human immunodeficiency virus by DC-SIGN+ cells in human rectal mucosa. J. Virol. 79, 5762 (2005).

152 Arrighi, J. F. et al., DC-SIGN-mediated infectious synapse formation enhances X4 HIV-1 transmission from dendritic cells to T cells. J. Exp. Med. 200, 1279 (2004).

153 Moris, A. et al., Dendritic cells and HIV-specific CD4+ T cells: HIV antigen presentation, T-cell activation, and viral transfer. Blood 108, 1643 (2006).

154 Abendroth, A., Morrow, G., Cunningham, A. L. &Slobedman, B., Varicella-zoster virus infection of human dendritic cells and transmission to T cells: implications for virus dissemination in the host. J. Virol. 75, 6183 (2001).

155 Shingai, M. et al., Wild-type measles virus infection in human CD46/CD150-transgenic mice: CD11c-positive dendritic cells establish systemic viral infection. J. Immunol. 175, 3252 (2005).

156 Wahid, R., Cannon, M. J. &Chow, M., Dendritic cells and macrophages are productively infected by poliovirus. J. Virol. 79, 401 (2005).

157 Bozza, S. et al., Dendritic cells transport conidia and hyphae of Aspergillus fumigatus from the airways to the draining lymph nodes and initiate disparate Th responses to the fungus. J. Immunol. 168, 1362 (2002).

158 Serrano-Gomez, D. et al., Dendritic cell-specific intercellular adhesion molecule 3-grabbing nonintegrin mediates binding and internalization of Aspergillus fumigatus conidia by dendritic cells and macrophages. J. Immunol. 173, 5635 (2004).

159 Sevilla, N. et al., Immunosuppression and resultant viral persistence by specific viral targeting of dendritic cells. J. Exp. Med. 192, 1249 (2000).

160 Lambert, H., Hitziger, N., Dellacasa, I., Svensson, M. &Barragan, A., Induction of dendritic cell migration upon Toxoplasma gondii infection potentiates parasite dissemination. Cell. Microbiol. 8, 1611 (2006).

161 Huang, F.-P., Farquhar, C. F., Mabbott, N. A., Bruce, M. E. &MacPherson, G. G., Migrating intestinal dendritic cells transport PrPSc from the gut. J. Gen. Virol. 83, 267 (2002).

162 Neuenhahn, M. et al., CD8α+ dendritic cells are required for efficient entry of Listeria monocytogenes into the spleen. Immunity 25, 619 (2006).

163 Cleret, A. et al., Lung dendritic cells rapidly mediate anthrax spore entry through the pulmonary route. J. Immunol. 178, 7994 (2007).

164 Colonna, M., Krug, A. &Cella, M., Interferon-producing cells: on the front line in immune responses against pathogens. Curr. Opin. Immunol. 14, 373 (2002).

165 Pacanowski, J. et al., Reduced blood CD123+ (lymphoid) and CD11c+ (myeloid) dendritic cell numbers in primary HIV-1 infection. Blood 98, 3016 (2001).

166 Soumelis, V. et al., Depletion of circulating natural type 1 interferon-producing cells in HIV-infected AIDS patients. Blood 98, 906 (2001).

167 Donaghy, H. et al., Loss of blood CD11c(+) myeloid and CD11c(-) plasmacytoid dendritic cells in patients with HIV-1 infection correlates with HIV-1 RNA virus load. Blood 98, 2574 (2001).

27

Page 28: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

168 Chehimi, J. et al., Persistent decreases in blood plasmacytoid dendritic cell number and function despite effective highly active antiretroviral therapy and increased blood myeloid dendritic cells in HIV-infected individuals. J. Immunol. 168, 4796 (2002).

169 Finke, J. S., Shodell, M., Shah, K., Siegal, F. P. &Steinman, R. M., Dendritic cell numbers in the blood of HIV-1 infected patients before and after changes in antiretroviral therapy. J. Clin. Immunol. 24, 647 (2004).

170 Hishizawa, M. et al., Depletion and impaired interferon-alpha-producing capacity of blood plasmacytoid dendritic cells in human T-cell leukaemia virus type I-infected individuals. Br. J. Haematol. 125, 568 (2004).

171 Gill, M. A. et al., Mobilization of plasmacytoid and myeloid dendritic cells to mucosal sites in children with respiratory syncytial virus and other viral respiratory infections. J. Infect. Dis. 191, 1105 (2005).

172 Boscardin, S. B. et al., Antigen targeting to dendritic cells elicits long-lived T cell help for antibody responses. J. Exp. Med. 203, 599 (2006).

173 Granelli-Piperno, A. et al., Dendritic cell -specific intercellular adhesion molecule 3-grabbing nonintegrin/CD209 is abundant on macrophages in the normal human lymph node and is not required for dendritic cell stimulation of the mixed leukocyte reaction. J. Immunol. 175, 4265 (2005).

174 Pack, M. et al., Dec-205/CD205+ dendritic cells are abundant in the white pulp of human spleen and are also found in the border region between the red and white pulp. Immunol. In Press (2007).

175 Charalambous, A., Oks, M., Nchinda, G., Yamazaki, S. &Steinman, R. M., Dendritic cell targeting of survivin protein in a xenogeneic form elicits strong CD4+ T cell immunity to mouse survivin. J. Immunol. 177, 8410 (2006).

176 Soares, H. et al., A subset of dendritic cells induces CD4+ T cells to produce IFN-γ by an IL-12-independent but CD70 dependent mechanism in vivo. J. Exp. Med. 204, 1095 (2007).

177 Bozzacco, L. et al., DEC-205 receptor on dendritic cells mediates presentation of HIV gag protein to CD8+ T cells in a spectrum of human MHC I haplotypes. Proc. Natl. Acad. Sci. USA 104, 1289 (2007).

178 Delneste, Y. et al., Involvement of LOX-1 in dendritic cell-mediated antigen cross-presentation. Immunity 17, 353 (2002).

179 Ramakrishna, V. et al., Toll-like receptor activation enhances cell-mediated immunity induced by an antibody vaccine targeting human dendritic cells. Journal of translational medicine 5, 5 (2007).

180 McKenzie, E. J. et al., Mannose receptor expression and function define a new population of murine dendritic cells. J. Immunol. 178, 4975 (2007).

181 Tacken, P. J. et al., Effective induction of naive and recall T-cell responses by targeting antigen to human dendritic cells via a humanized anti-DC-SIGN antibody. Blood 106, 1278 (2005).

182 Vingert, B. et al., The Shiga toxin B-subunit targets antigen in vivo to dendritic cells and elicits anti-tumor immunity. Eur J Immunol 36, 1124 (2006).

183 Jeannin, P. et al., OmpA targets dendritic cells, induces their maturation and delivers antigen into the MHC class 1 presentation pathway. Nat. Immunol. 1, 502 (2000).

184 Schjetne, K. W., Fredriksen, A. B. &Bogen, B., Delivery of antigen to CD40 induces protective immune responses against tumors. J Immunol 178, 4169 (2007).

28

Page 29: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

185 Vicari, A. P. et al., Reversal of tumor-induced dendritic cell paralysis by CpG immunostimulatory oligonucleotide and anti-interleukin 10 receptor antibody. J. Exp. Med. 196, 541 (2002).

186 Bell, D. et al., In breast carcinoma tissue, immature dendritic cells reside within the tumor, whereas mature dendritic cells are located in peritumoral areas. J. Exp. Med. 190, 1417 (1999).

187 Rabinovich, G. A., Gabrilovich, D. &Sotomayor, E. M., Immunosuppressive strategies that are mediated by tumor cells. Annu. Rev. Immunol. 25, 267 (2007).

188 Zou, W., Regulatory T cells, tumour immunity and immunotherapy. Nat. Rev. Immunol. 6, 295 (2006).

189 Gabrilovich, D., Mechanisms and functional significance of tumour-induced dendritic-cell defects. Nat. Rev. Immunol. 4, 941 (2004).

190 Chomarat, P., Banchereau, J., Davoust, J. &Palucka, A. K., IL-6 switches the differentiation of monocytes from dendritic cells to macrophages. Nat. Immunol. 1, 510 (2000).

191 Gabrilovitch, D. I. et al., Production of vascular endothelial growth factor by human tumors inhibits the functional maturation of dendritic cells. Nat. Med. 2, 1096 (1996).

192 Sumimoto, H., Imabayashi, F., Iwata, T. &Kawakami, Y., The BRAF-MAPK signaling pathway is essential for cancer-immune evasion in human melanoma cells. J. Exp. Med. 203, 1651 (2006).

193 Kortylewski, M. et al., Inhibiting Stat3 signaling in the hematopoietic system elicits multicomponent antitumor immunity. Nat. Med. 11, 1314 (2005).

194 Wang, T. et al., Regulation of the innate and adaptive immune responses by Stat-3 signaling in tumor cells. Nat. Med. 10, 48 (2004).

195 Ghiringhelli, F. et al., Tumor cells convert immature myeloid dendritic cells into TGF-β-secreting cells inducing CD4+CD25+ regulatory T cell proliferation. J. Exp. Med. 202, 919 (2005).

196 Aspord, C. et al., Breast cancer instructs dendritic cells to prime interleukin 13-secreting CD4+ T cells that facilitate tumor development. J. Exp. Med. 204, 1037 (2007).

197 Terabe, M. et al., NKT cell-mediated repression of tumor immunosurveillance by IL-13 and the IL-4R-STAT6 pathway. Nat. Immunol. 1, 515 (2000).

198 Kukreja, A. et al., Enhancement of clonogenicity of human multiple myeloma by dendritic cells. J. Exp. Med. 203, 1859 (2006).

199 Vicari, A. P., Caux, C. &Trinchieri, G., Tumour escape from immune surveillance through dendritic cell inactivation. Semin. Cancer Biol. 12, 33 (2002).

200 Adams, G. P. & Weiner, L. M., Monoclonal antibody therapy of cancer. Nat. Biotechnol. 23, 1147 (2005).

201 Waldmann, T. A., Immunotherapy: past, present and future. Nat. Med. 9, 269 (2003). 202 Slamon, D. J. et al., Use of chemotherapy plus a monoclonal antibody against HER2 for

metastatic breast cancer that overexpresses HER2. N. Engl. J. Med. 344, 783 (2001). 203 Clynes, R. A., Towers, T. L., Presta, L. G. &Ravetch, J. V., Inhibitory Fc receptors

modulate in vivo cytotoxicity against tumor targets. Nat. Med. 6, 443 (2000). 204 Ravetch, J. V. & Bolland, S., IgG Fc receptors. Annu. Rev. Immunol. 19, 275 (2001). 205 Regnault, A. et al., Fcγ receptor-mediated induction of dendritic cell maturation and

major histocompatibility complex class I-restricted antigen presentation after immune complex internalization. J. Exp. Med. 189, 371 (1999).

29

Page 30: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

206 Dhodapkar, K. M., Krasovsky, J., Williamson, B. &Dhodapkar, M. V., Anti-tumor monoclonal antibodies enhance cross-presentation of cellular antigens and the generation of myeloma-specific killer T cells by dendritic cells. J. Exp. Med. 195, 125 (2002).

207 Kalergis, A. M. & Ravetch, J. V., Inducing tumor immunity through the selective engagement of activating Fcγ receptors on dendritic cells. J. Exp. Med. 195, 1653 (2002).

208 Rafiq, K., Bergtold, A. &Clynes, R., Immune complex-mediated antigen presentation induces tumor immunity. J. Clin. Invest. 110, 71 (2002).

209 Dhodapkar, K. M. et al., Selective blockade of inhibitory Fcγ receptor enables human dendritic cell maturation with IL-12p70 production and immunity to antibody-coated tumor cells. Proc. Natl. Acad. Sci. USA 102, 2910 (2005).

210 Boruchov, A. M. et al., Activating and inhibitory IgG Fc receptors on human DCs mediate opposing functions. J. Clin. Invest. 115, 2914 (2005).

211 Blattman, J. N. & Greenberg, P. D., Cancer immunotherapy: a treatment for the masses. Science (New York, N.Y 305, 200 (2004).

212 Rosenberg, S. A., Shedding light on immunotherapy for cancer. N. Engl. J. Med. 350, 1461 (2004).

213 Dudley, M. E. et al., Cancer regression and autoimmunity in patients after clonal repopulation with antitumor lymphocytes. Science (New York, N.Y 298, 850 (2002).

214 Morgan, R. A. et al., Cancer regression in patients after transfer of genetically engineered lymphocytes. Science (New York, N.Y 314, 126 (2006).

215 Brodie, S. J. et al., In vivo migration and function of transferred HIV-1-specific cytotoxic T cells. Nature Med. 5, 34 (1999).

216 Yee, C. et al., Adoptive T cell therapy using antigen-specific CD8+ T cell clones for the treatment of patients with metastatic melanoma: in vivo persistence, migration, and antitumor effect of transferred T cells. Proc. Natl. Acad. Sci. USA 99, 16168 (2002).

217 Butturini, A. & Gale, R. P., The role of T-cells in preventing relapse in chronic myelogenous leukemia. Bone Marrow Transplant. 2, 351 (1987).

218 Giralt, S. A. & Kolb, H. J., Donor lymphocyte infusions. Curr. Opin. Oncol. 8, 96 (1996). 219 Velardi, A., Ruggeri, L., Alessandro, Moretta &Moretta, L., NK cells: a lesson from

mismatched hematopoietic transplantation. Trends Immunol 23, 438 (2002). 220 Appelbaum, F. R., Haematopoietic cell transplantation as immunotherapy. Nature 411,

385 (2001). 221 Dhodapkar, M. V. et al., A reversible defect in natural killer T cell function characterizes

the progression of premalignant to malignant multiple myeloma. J. Exp. Med. 197, 1667 (2003).

222 Dhodapkar, M. V., Krasovsky, J., Osman, K. &Geller, M. D., Vigorous premalignancy-specific effector T cell response in the bone marrow of patients with monoclonal gammopathy. J. Exp. Med. 198, 1753 (2003).

223 Darnell, R. B. & Posner, J. B., Observing the invisible: successful tumor immunity in humans. Nat. Immunol. 4, 201 (2003).

224 Pages, F. et al., Effector memory T cells, early metastasis, and survival in colorectal cancer. N. Engl. J. Med. 353, 2654 (2005).

225 Galon, J. et al., Type, density, and location of immune cells within human colorectal tumors predict clinical outcome. Science (New York, N.Y 313, 1960 (2006).

226 Gutierrez-Dalmau, A. & Campistol, J. M., Immunosuppressive therapy and malignancy in organ transplant recipients: a systematic review. Drugs 67, 1167 (2007).

30

Page 31: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

227 Shankaran, V. et al., IFNγ and lymphocytes prevent primary tumour development and shape tumour immunogenicity. Nature 410, 1107 (2001).

228 Dunn, G. P., Old, L. J. &Schreiber, R. D., The immunobiology of cancer immunosurveillance and immunoediting. Immunity 21, 137 (2004).

229 Banchereau, J. & Palucka, A. K., Dendritic cells as therapeutic vaccines against cancer. Nat. Rev. Immunol. 5, 296 (2005).

230 Renkvist, N., Castelli, C., Robbins, P. F. &Parmiani, G., A listing of human tumor antigens recognized by T cells. Cancer Immunol. Immunother. 50, 3 (2001).

231 Boon, T., Coulie, P. G., Van den Eynde, B. J. &van der Bruggen, P., Human T cell responses against melanoma. Annu. Rev. Immunol. 24, 175 (2006).

232 Simpson, A. J., Caballero, O. L., Jungbluth, A., Chen, Y. T. &Old, L. J., Cancer/testis antigens, gametogenesis and cancer. Nat. Rev. Cancer 5, 615 (2005).

233 Timmerman, J. M. & Levy, R., Dendritic cell vaccines for cancer immunotherapy. Annu. Rev. Med. 50, 507 (1999).

234 Fong, L. & Engleman, E. G., Dendritic cells in cancer immunotherapy. Annu. Rev. Immunol. 18, 245 (2000).

235 Berard, F. et al., Cross-priming of naive CD8 T cells against melanoma antigens using dendritic cells loaded with killed allogeneic melanoma cells. J. Exp. Med. 192, 1535 (2000).

236 Schuler, G., Schuler-Thurner, B. &Steinman, R. M., The use of dendritic cells in cancer immunotherapy. Curr. Opin. Immunol. 15, 138 (2003).

237 Figdor, C. G., De Vries, I. J., Lesterhuis, W. J. &Melief, C. J., Dendritic cell immunotherapy: mapping the way. Nat. Med. 10, 475 (2004).

238 Cerundolo, V., Hermans, I. F. &Salio, M., Dendritic cells: a journey from laboratory to clinic. Nat. Immunol. 5, 7 (2004).

239 Zitvogel, L., Dendritic and natural killer cells cooperate in the control/switch of innate immunity. J. Exp. Med. 195, F9 (2002).

240 Walzer, T., Dalod, M., Robbins, S. H., Zitvogel, L. &Vivier, E., Natural-killer cells and dendritic cells: "l'union fait la force". Blood 106, 2252 (2005).

241 Fujii, S., Shimizu, K., Kronenberg, M. &Steinman, R. M., Prolonged interferon-γ producing NKT response induced with α-galactosylceramide-loaded dendritic cells. Nat. Immunol. 3, 867 (2002).

242 Fujii, S., Shimizu, K., Smith, C., Bonifaz, L. &Steinman, R. M., Activation of natural killer T cells by α-galactosylceramide rapidly induces the full maturation of dendritic cells in vivo and thereby acts as an adjuvant for combined CD4 and CD8 T cell immunity to a co-administered protein. J. Exp. Med. 198, 267 (2003).

243 Fujii, S., Liu, K., Smith, C., Bonito, A. J. &Steinman, R. M., The linkage of innate to adaptive immunity via maturing dendritic cells in vivo requires CD40 ligation in addition to antigen presentation and CD80/86 costimulation. J. Exp. Med. 199, 1607 (2004).

244 Hermans, I. F. et al., NKT cells enhance CD4+ and CD8+ T cell responses to soluble antigen in vivo through direct interaction with dendritic cells. J. Immunol. 171, 5140 (2003).

245 Gao, Y. et al., γδ T cells provide an early source of interferon γ in tumor immunity. J. Exp. Med. 198, 433 (2003).

246 van Mierlo, G. J. et al., Activation of dendritic cells that cross-present tumor-derived antigen licenses CD8+ CTL to cause tumor eradication. J. Immunol. 173, 6753 (2004).

31

Page 32: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

247 Paczesny, S. et al., Expansion of melanoma-specific cytolytic CD8+ T cell precursors in patients with metastatic melanoma vaccinated with CD34+ progenitor-derived dendritic cells. J. Exp. Med. 199, 1503 (2004).

248 Liu, K. et al., Innate NKT lymphocytes confer superior adaptive immunity via tumor-capturing dendritic cells. J. Exp. Med. 202, 1507 (2005).

249 Nestle, F. O. et al., Vaccination of melanoma patients with peptide-or tumor lysate-pulsed dendritic cells. Nat. Med. 4, 328 (1998).

250 Hsu, F. J. et al., Vaccination of patients with B-cell lymphoma using autologous antigen-pulsed dendritic cells. Nat. Med. 2, 52 (1996).

251 Thurner, B. et al., Vaccination with Mage-3A1 peptide-pulsed mature, monocyte-derived dendritic cells expands specific cytotoxic T cells and induces regression of some metastases in advanced stage IV melanoma. J. Exp. Med. 190, 1669 (1999).

252 Timmerman, J. M. & Levy, R., Linkage of foreign carrier protein to a self-tumor antigen enhances the immunogenicity of a pulsed dendritic cell vaccine. J. Immunol. 164, 4797 (2000).

253 Schuler-Thurner, B. et al., Mage-3 and influenza-matrix peptide-specific cytotoxic T cells are inducible in terminal stage HLA-A.2.1+ melanoma patients by mature monocyte-derived dendritic cells. J. Immunol. 165, 3492 (2000).

254 Timmerman, J. M. et al., Idiotype-pulsed dendritic cell vaccination for B-cell lymphoma: clinical and immune responses in 35 patients. Blood 99, 1517 (2002).

255 Finn, O. J., Premalignant lesions as targets for cancer vaccines. J. Exp. Med. 198, 1623 (2003).

256 Lollini, P. L., Cavallo, F., Nanni, P. &Forni, G., Vaccines for tumour prevention. Nat. Rev. Cancer 6, 204 (2006).

257 Fernandez, N. C. et al., Dendritic cells directly trigger NK cell functions: cross-talk relevant in innate anti-tumor immune responses in vivo. Nat. Med. 5, 405 (1999).

258 Ferlazzo, G. et al., Human dendritic cells activate resting NK cells and are recognized via the NKp30 receptor by activated NK cells. J. Exp. Med. 195, 343 (2002).

259 Gerosa, F. et al., Reciprocal activating interaction between Natural Killer cells and dendritic cells. J. Exp. Med. 195, 327 (2002).

260 Ferlazzo, G. et al., Distinct roles of IL-12 and IL-15 in human natural killer cell activation by dendritic cells from secondary lymphoid organs. Proc. Natl. Acad. Sci. USA 101, 16606 (2004).

261 Lucas, M., Schachterle, W., Oberle, K., Aichele, P. &Diefenbach, A., Dendritic cells prime natural killer cells by trans-presenting interleukin 15. Immunity 26, 503 (2007).

262 Stary, G. et al., Tumoricidal activity of TLR7/8-activated inflammatory dendritic cells. J. Exp. Med. 204, 1441 (2007).

263 Palucka, A. K. et al., Dendritic cells loaded with killed allogeneic melanoma cells can induce objective clinical responses and MART-1 specific CD8+ T-cell immunity. J. Immunother. 29, 545 (2006).

264 Fong, L. et al., Altered peptide ligand vaccination with Flt3 ligand expanded dendritic cells for tumor immunotherapy. Proc. Natl. Acad. Sci. USA 98, 8809 (2001).

265 O'Rourke, M. G. et al., Durable complete clinical responses in a phase I/II trial using an autologous melanoma cell/dendritic cell vaccine. Cancer Immunol. Immunother. 52, 387 (2003).

32

Page 33: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

266 Josien, R. et al., TRANCE, a tumor necrosis family member, enhances the longevity and adjuvant properties of dendritic cells in vivo. J. Exp. Med. 191, 495 (2000).

267 De Vries, I. J. et al., Effective migration of antigen-pulsed dendritic cells to lymph nodes in melanoma patients is determined by their maturation state. Cancer Res. 63, 12 (2003).

268 Martin-Fontecha, A. et al., Regulation of dendritic cell migration to the draining lymph node: impact on T lymphocyte traffic and priming. J. Exp. Med. 198, 615 (2003).

269 Bullock, T. N., Mullins, D. W. &Engelhard, V. H., Antigen density presented by dendritic cells in vivo differentially affects the number and avidity of primary, memory, and recall CD8+ T cells. J. Immunol. 170, 1822 (2003).

270 Berzofsky, J. A. et al., Progress on new vaccine strategies for the immunotherapy and prevention of cancer. J. Clin. Invest. 113, 1515 (2004).

271 Sutmuller, R. P. et al., Synergism of cytotoxic T lymphocyte-associated antigen 4 blockade and depletion of CD25+ regulatory T cells in antitumor therapy reveals alternative pathways for suppression of autoreactive cytotoxic T lymphocyte responses. J. Exp. Med. 194, 823 (2001).

272 Betts, M. R. et al., HIV nonprogressors preferentially maintain highly functional HIV-specific CD8+ T cells. Blood 107, 4781 (2006).

273 Darrah, P. A. et al., Multifunctional TH1 cells define a correlate of vaccine-mediated protection against Leishmania major. Nat. Med. 13, 843 (2007).

274 Petrovas, C. et al., PD-1 is a regulator of virus-specific CD8+ T cell survival in HIV infection. J. Exp. Med. 203, 2281 (2006).

275 Trautmann, L. et al., Upregulation of PD-1 expression on HIV-specific CD8+ T cells leads to reversible immune dysfunction. Nat. Med. 12, 1198 (2006).

276 Day, C. L. et al., PD-1 expression on HIV-specific T cells is associated with T-cell exhaustion and disease progression. Nature 443, 350 (2006).

277 Belyakov, I. M. et al., Impact of vaccine-induced mucosal high-avidity CD8+ CTLs in delay of AIDS viral dissemination from mucosa. Blood 107, 3258 (2006).

278 Schadendorf, D. et al., Dacarbazine (DTIC) versus vaccination with autologous peptide-pulsed dendritic cells (DC) in first-line treatment of patients with metastatic melanoma: a randomized phase III trial of the DC study group of the DeCOG. Ann. Oncol. 17, 563 (2006).

279 Banchereau, J. et al., Immune and clinical responses in patients with metastatic melanoma to CD34+ progenitor-derived dendritic cell vaccine. Cancer Res. 61, 6451 (2001).

280 Fong, L., Brockstedt, D., Benike, C., Wu, L. &Engleman, E. G., Dendritic cells injected via different routes induce immunity in cancer patients. J. Immunol. 166, 4254 (2001).

281 Su, Z. et al., Telomerase mRNA-transfected dendritic cells stimulate antigen-specific CD8+ and CD4+ T cell responses in patients with metastatic prostate cancer. J. Immunol. 174, 3798 (2005).

282 Pardoll, D. & Allison, J., Cancer immunotherapy: breaking the barriers to harvest the crop. Nat. Med. 10, 887 (2004).

283 Tong, Y., Song, W. &Crystal, R. G., Combined intratumoral injection of bone marrow-derived dendritic cells and systemic chemotherapy to treat pre-existing murine tumors. Cancer Res. 61, 7530 (2001).

284 Casares, N. et al., Caspase-dependent immunogenicity of doxorubicin-induced tumor cell death. J. Exp. Med. 202, 1691 (2005).

33

Page 34: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

285 Obeid, M. et al., Calreticulin exposure dictates the immunogenicity of cancer cell death. Nat. Med. 13, 54 (2007).

286 Asavaroengchai, W., Kotera, Y. &Mulé, J. J., Tumor lysate-pulsed dendritic cells can elicit an effective antitumor immune response during early lymphoid recovery. Proc. Natl. Acad. Sci. USA 99, 931 (2002).

287 Apetoh, L. et al., Toll-like receptor 4-dependent contribution of the immune system to anticancer chemotherapy and radiotherapy. Nat. Med. 13, 1050 (2007).

288 Newcomb, E. W. et al., The combination of ionizing radiation and peripheral vaccination produces long-term survival of mice bearing established invasive GL261 gliomas. Clin. Cancer Res. 12, 4730 (2006).

289 Quezada, S. A., Peggs, K. S., Curran, M. A. &Allison, J. P., CTLA4 blockade and GM-CSF combination immunotherapy alters the intratumor balance of effector and regulatory T cells. J. Clin. Invest. 116, 1935 (2006).

290 Peggs, K. S., Quezada, S. A., Korman, A. J. &Allison, J. P., Principles and use of anti-CTLA4 antibody in human cancer immunotherapy. Curr. Opin. Immunol. (2006).

291 Curiel, T. J. et al., Specific recruitment of regulatory T cells in ovarian carcinoma fosters immune privilege and predicts reduced survival. Nat. Med. 10, 942 (2004).

292 Kryczek, I. et al., Induction of B7-H4 on APCs through IL-10: novel suppressive mode for regulatory T cells. J. Immunol. 177, 40 (2006).

293 Roncarolo, M. G. et al., Interleukin-10-secreting type 1 regulatory T cells in rodents and humans. Immunol. Rev. 212, 28 (2006).

294 Vence, L. et al., Circulating tumor antigen-specific regulatory T cells in patients with metastatic melanoma. Proc. Natl. Acad. Sci. USA (In Press).

295 Ghiringhelli, F. et al., CD4+CD25+ regulatory T cells inhibit natural killer cell functions in a transforming growth factor-{beta}-dependent manner. J. Exp. Med. 202, 1075 (2005).

296 Berger, T. G. et al., Efficient elutriation of monocytes within a closed system (Elutratrade TM) for clinical-scale generation of dendritic cells. J. Immunol. Methods 298, 61 (2005).

297 Schaft, N. et al., Generation of an optimized polyvalent monocyte-derived dendritic cell vaccine by transfecting defined RNAs after rather than before maturation. J. Immunol. 174, 3087 (2005).

298 Erdmann, M. et al., Effective clinical-scale production of dendritic cell vaccines by monocyte elutriation directly in medium, subsequent culture in bags and final antigen loading using peptides or RNA transfection. J. Immunotherapy 30, 663 (2007).

299 Dhodapkar, M. V., Krasovsky, J. &Olson, K., T cells from the tumor microenvironment of patients with progressive myeloma can generate strong, tumor-specific cytolytic responses to autologous, tumor-loaded dendritic cells. Proc. Natl. Acad. Sci. USA 99, 13009 (2002).

300 Dubsky, P. et al., IL-15-induced human DC efficiently prime melanoma-specific naive CD8+ T cells to differentiate into CTL. Eur. J. Immunol., 1 (2007).

301 Gilboa, E. & Vieweg, J., Cancer immunotherapy with mRNA-transfected dendritic cells. Immunol. Rev. 199, 251 (2004).

302 Spisek, R. et al., Frequent and specific immunity to the embryonal stem cell associated antigen SOX2 in patients with monoclonal gammopathy. J. Exp. Med. 204, 831 (2007).

303 Mailliard, R. B. et al., alpha-type-1 polarized dendritic cells: a novel immunization tool with optimized CTL-inducing activity. Cancer Res. 64, 5934 (2004).

34

Page 35: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

304 Mohamadzadeh, M. et al., Interleukin 15 skews monocyte differentiation into dendritic cells with features of Langerhans cells. J. Exp. Med. 194, 1013 (2001).

305 Soiffer, R. et al., Vaccination with irradiated autologous melanoma cells engineered to secrete human granulocyte-macrophage colony-stimulating factor generates potent antitumor immunity in patients with metastatic melanoma. Proc. Natl. Acad. Sci. USA 95, 13141 (1998).

306 Salgia, R. et al., Vaccination with irradiated autologous tumor cells engineered to secrete granulocyte-macrophage colony-stimulating factor augments antitumor immunity in some patients with metastatic non-small-cell lung carcinoma. J Clin Oncol 21, 624 (2003).

307 Thomas, A. M. et al., Mesothelin-specific CD8+ T cell responses provide evidence of in vivo cross-priming by antigen-presenting cells in vaccinated pancreatic cancer patients. J. Exp. Med. 200, 297 (2004).

308 Laheru, D. & Jaffee, E. M., Immunotherapy for pancreatic cancer - science driving clinical progress. Nat. Rev. Cancer 5, 459 (2005).

309 Fujii, S., Shimizu, K., Hemmi, H. &Steinman, R. M., Innate Vα14+ NKT cells mature dendritic cells, leading to strong adaptive immunity. Immunol. Rev. In Press (2007).

310 Ahonen, C. L. et al., Combined TLR and CD40 triggering induces potent CD8+ T cell expansion with variable dependence on type I IFN. J. Exp. Med. 199, 775 (2004).

311 Shimizu, K., Kurosawa, Y., Taniguchi, M., Steinman, R. M. &Fujii, S., Cross presentation of tumor cells loaded with α-galactosylceramide leads to potent and long lived T cell mediated immunity via dendritic cells. J. Exp. Med. In Press (2007).

312 Ramakrishna, V. et al., Mannose receptor targeting of tumor antigen pmel17 to human dendritic cells directs anti-melanoma T cell responses via multiple HLA molecules. J. Immunol. 172, 2845 (2004).

313 Dittel, B. N., Visintin, I., Merchant, R. M. &Janeway, C. A., Jr., Presentation of the self antigen myelin basic protein by dendritic cells leads to experimental autoimmune encephalomyelitis. J. Immunol. 163, 32 (1999).

314 Eriksson, U. et al., Dendritic cell-induced autoimmune heart failure requires cooperation between adaptive and innate immunity. Nat. Med. 9, 1484 (2003).

315 Marty, R. R. & Eriksson, U., Dendritic cells and autoimmune heart failure. Int. J. Cardiol. 112, 34 (2006).

316 Bondanza, A. et al., Dissociation between autoimmune response and clinical disease after vaccination with dendritic cells. J. Immunol. 170, 24 (2003).

317 Banchereau, J., Pascual, V. &Palucka, A. K., Autoimmunity through cytokine-induced dendritic cell activation. Immunity 20, 539 (2004).

318 Banchereau, J. & Pascual, V., Type I interferon in systemic lupus erythematosus and other autoimmune diseases. Immunity 25, 383 (2006).

319 Palucka, A. K., Blanck, J. P., Bennett, L., Pascual, V. &Banchereau, J., Cross-regulation of TNF and IFN-α in autoimmune diseases. Proc. Natl. Acad. Sci. USA 102, 3372 (2005).

320 Vogelsang, P., Jonsson, M. V., Dalvin, S. T. &Appel, S., Role of dendritic cells in Sjogren's syndrome. Scand. J. Immunol. 64, 219 (2006).

321 Feldmann, M. & Maini, R. N., Lasker Clinical Medical Research Award. TNF defined as a therapeutic target for rheumatoid arthritis and other autoimmune diseases. Nat. Med. 9, 1245 (2003).

35

Page 36: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

322 Lowes, M. A. et al., Increase in TNFα and inducible nitric oxide synthase-expressing dendritic cells in psoriasis and reduction with efalizumab (anti-CD11a). Proc. Natl. Acad. Sci. USA 102, 19057 (2005).

323 Bennett, L. et al., Interferon and granulopoiesis signatures in systemic lupus erythematosus blood. J. Exp. Med. 197, 711 (2003).

324 Siegal, F. P. et al., The nature of the principal type 1 interferon-producing cells in human blood. Science (New York, N.Y 284, 1835 (1999).

325 Asselin-Paturel, C. & Trinchieri, G., Production of type I interferons: plasmacytoid dendritic cells and beyond. J. Exp. Med. 202, 461 (2005).

326 Farkas, L., Beiske, K., Lund-Johansen, F., Brandtzaeg, P. &Jahnsen, F. L., Plasmacytoid dendritic cells (natural interferon-α/β-producing cells) accumulate in cutaneous lupus erythematosus lesions. Am. J. Pathol. 159, 237 (2001).

327 Boule, M. W. et al., Toll-like receptor 9-dependent and -independent dendritic cell activation by chromatin-immunoglobulin G complexes. J. Exp. Med. 199, 1631 (2004).

328 Means, T. K. et al., Human lupus autoantibody-DNA complexes activate DCs through cooperation of CD32 and TLR9. J. Clin. Invest. 115, 407 (2005).

329 Marshak-Rothstein, A., Toll-like receptors in systemic autoimmune disease. Nat. Rev. Immunol. 6, 823 (2006).

330 Tian, J. et al., Toll-like receptor 9-dependent activation by DNA-containing immune complexes is mediated by HMGB1 and RAGE. Nat. Immunol. 8, 487 (2007).

331 Blanco, P., Palucka, A. K., Gill, M., Pascual, V. &Banchereau, J., Induction of dendritic cell differentiation by IFN-α in systemic lupus erythematosus. Science (New York, N.Y 294, 1540 (2001).

332 Dhodapkar, K. M. et al., Selective blockade of the inhibitory Fcγ receptor (FcγRIIB) in human dendritic cells and monocytes leads to the induction of type I interferon response program. J. Exp. Med. 204, 1359 (2007).

333 Nestle, F. O. et al., Plasmacytoid predendritic cells initiate psoriasis through interferon-α production. J. Exp. Med. 202, 135 (2005).

334 Greenberg, S. A. et al., Interferon-alpha/beta-mediated innate immune mechanisms in dermatomyositis. Ann Neurol 57, 664 (2005).

335 Gottenberg, J. E. et al., Activation of IFN pathways and plasmacytoid dendritic cell recruitment in target organs of primary Sjogren's syndrome. Proc. Natl. Acad. Sci. USA 103, 2770 (2006).

336 Kastelein, R. A., Hunter, C. A. &Cua, D. J., Discovery and biology of IL-23 and IL-27: related but functionally distinct regulators of inflammation. Annu. Rev. Immunol. 25, 221 (2007).

337 Hue, S. et al., Interleukin-23 drives innate and T cell-mediated intestinal inflammation. J. Exp. Med. 203, 2473 (2006).

338 Kullberg, M. C. et al., IL-23 plays a key role in Helicobacter hepaticus-induced T cell-dependent colitis. J. Exp. Med. 203, 2485 (2006).

339 Langrish, C. L. et al., IL-23 drives a pathogenic T cell population that induces autoimmune inflammation. J. Exp. Med. 201, 233 (2005).

340 Duerr, R. H. et al., A genome-wide association study identifies IL23R as an inflammatory bowel disease gene. Science (New York, N.Y 314, 1461 (2006).

341 Lowes, M. A., Bowcock, A. M. &Krueger, J. G., Pathogenesis and therapy of psoriasis. Nature 445, 866 (2007).

36

Page 37: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

342 Smits, H. H. et al., Commensal Gram-negative bacteria prime human dendritic cells for enhanced IL-23 and IL-27 expression and enhanced Th1 development. Eur J Immunol 34, 1371 (2004).

343 Lee, E. et al., Increased expression of interleukin 23 p19 and p40 in lesional skin of patients with psoriasis vulgaris. J. Exp. Med. 199, 125 (2004).

344 Uhlig, H. H. et al., Differential activity of IL-12 and IL-23 in mucosal and systemic innate immune pathology. Immunity 25, 309 (2006).

345 Piemonti, L. et al., Glucocorticoids affect human dendritic cell differentiation and maturation. J. Immunol. 162, 6473 (1999).

346 Shodell, M., Shah, K. &Siegal, F. P., Circulating human plasmacytoid dendritic cells are highly sensitive to corticosteroid administration. Lupus 12, 222 (2003).

347 Chen, X., Murakami, T., Oppenheim, J. J. &Howard, O. M., Triptolide, a constituent of immunosuppressive Chinese herbal medicine, is a potent suppressor of dendritic-cell maturation and trafficking. Blood 106, 2409 (2005).

348 Whartenby, K. A. et al., Inhibition of FLT3 signaling targets DCs to ameliorate autoimmune disease. Proc. Natl. Acad. Sci. USA 102, 16741 (2005).

349 Tussiwand, R., Onai, N., Mazzucchelli, L. &Manz, M. G., Inhibition of natural type I IFN-producing and dendritic cell development by a small molecule receptor tyrosine kinase inhibitor with Flt3 affinity. J. Immunol. 175, 3674 (2005).

350 Iruretagoyena, M. I. et al., Inhibition of nuclear factor-ΚB enhances the capacity of immature dendritic cells to induce antigen-specific tolerance in experimental autoimmune encephalomyelitis. J. Pharmacol. Exp. Ther. 318, 59 (2006).

351 Vizzardelli, C. et al., Effects of dexamethazone on LPS-induced activationand migration of mouse dendritic cells revealed by a genome-wide transcriptional analysis. Eur. J. Immunol. 36, 1504 (2006).

352 Bros, M. et al., A newly established murine immature dendritic cell line can be differentiated into a mature state, but exerts tolerogenic function upon maturation in the presence of glucocorticoid. Blood 109, 3820 (2007).

353 Menges, M. et al., Repetitive injections of dendritic cells matured with tumor necrosis factor α induce antigen-specific protection of mice from autoimmunity. J. Exp. Med. 195, 15 (2002).

354 Lechmann, M. et al., The extracellular domain of CD83 inhibits dendritic cell-mediated T cell stimulation and binds to a ligand on dendritic cells. J. Exp. Med. 194, 1813 (2001).

355 Verginis, P., Li, H. S. &Carayanniotis, G., Tolerogenic semimature dendritic cells suppress experimental autoimmune thyroiditis by activation of thyroglobulin-specific CD4+CD25+ T cells. J. Immunol. 174, 7433 (2005).

356 Yamazaki, S. et al., Direct expansion of functional CD25+ CD4+ regulatory T cells by antigen processing dendritic cells. J. Exp. Med. 198, 235 (2003).

357 Nishimura, E., Sakihama, T., Setoguchi, R., Tanaka, K. &Sakaguchi, S., Induction of antigen-specific immunologic tolerance by in vivo and in vitro antigen-specific expansion of naturally arising Foxp3+CD25+CD4+ regulatory T cells. Int. Immunol. 16, 1189 (2004).

358 Watanabe, N. et al., Hassall's corpuscles instruct dendritic cells to induce CD4+CD25+ regulatory T cells in human thymus. Nature 436, 1181 (2005).

359 Annacker, O. et al., Essential role for CD103 in the T cell-mediated regulation of experimental colitis. J. Exp. Med. 202, 1051 (2005).

37

Page 38: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

360 Tang, Q. et al., Visualizing regulatory T cell control of autoimmune responses in nonobese diabetic mice. Nat. Immunol. 7, 83 (2006).

361 Tarbell, K. V. et al., Dendritic cell-expanded, islet-specific, CD4+ CD25+ CD62L+ regulatory T cells restore normoglycemia in diabetic NOD mice. J. Exp .Med. 204, 191 (2007).

362 Steinman, R. M., Hawiger, D. &Nussenzweig, M. C., Tolerogenic dendritic cells. Annu. Rev. Immunol. 21, 685 (2003).

363 Brimnes, M. K., Bonifaz, L., Steinman, R. M. &Moran, T. M., Influenza virus-induced dendritic cell maturation is associated with the induction of strong T cell immunity to a coadministered, normally nonimmunogenic protein. J. Exp. Med. 198, 133 (2003).

364 Vermaelen, K. Y., Carro-Muino, I., Lambrecht, B. N. &Pauwels, R. A., Specific migratory dendritic cells rapidly transport antigen from the airways to the thoracic lymph nodes. J. Exp. Med. 193, 51 (2001).

365 Akbari, O., DeKruyff, R. H. &Umetsu, D. T., Pulmonary dendritic cells producing IL-10 mediate tolerance induced by respiratory exposure to antigen. Nat. Immunol. 2, 725 (2001).

366 Akbari, O. et al., Antigen-specific regulatory T cells develop via the ICOS/ICOS- ligand pathway and inhibit allergen-induced airway hyperreactivity. Nat. Med. 8, 1024 (2002).

367 van Rijt, L. S. et al., In vivo depletion of lung CD11c+ dendritic cells during allergen challenge abrogates the characteristic features of asthma. J. Exp. Med. 201, 981 (2005).

368 van Rijt, L. S., van Kessel, C. H., Boogaard, I. &Lambrecht, B. N., Respiratory viral infections and asthma pathogenesis: a critical role for dendritic cells? J. Clin. Virol. 34, 161 (2005).

369 Lambrecht, B. N. & Hammad, H., Taking our breath away: dendritic cells in the pathogenesis of asthma. Nat. Rev. Immunol. 3, 994 (2003).

370 Lambrecht, B. N., Dendritic cells and the regulation of the allergic immune response. Allergy 60, 271 (2005).

371 Holt, P. G. & Thomas, W. R., Sensitization to airborne environmental allergens: unresolved issues. Nat. Immunol. 6, 957 (2005).

372 Ziegler, S. F. & Liu, Y. J., Thymic stromal lymphopoietin in normal and pathogenic T cell development and function. Nat. Immunol. 7, 709 (2006).

373 Yoo, J. et al., Spontaneous atopic dermatitis in mice expressing an inducible thymic stromal lymphopoietin transgene specifically in the skin. J. Exp. Med. 202, 541 (2005).

374 Zhou, B. et al., Thymic stromal lymphopoietin as a key initiator of allergic airway inflammation in mice. Nat. Immunol. 6, 1047 (2005).

375 Ebner, S. et al., Thymic stromal lemphopoietin converts human epidermal Langerhans cells into antigen presenting cells that induce pro-allergic T cells. J. Allergy Clin. Immunol. 119, 982 (2007).

376 Jenkins, S. J., Perona-Wright, G., Worsley, A. G., Ishii, N. &Macdonald, A. S., Dendritic cell expression of OX40 ligand acts as a costimulatory, not polarizing, signal for optimal Th2 priming and memory induction in vivo. J. Immunol. 179, 3515 (2007).

377 Li, M. et al., Retinoid X receptor ablation in adult mouse keratinocytes generates an atopic dermatitis triggered by thymic stromal lymphopoietin. Proc. Natl. Acad. Sci. USA 102, 14795 (2005).

38

Page 39: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

378 Bogiatzi, S. I. et al., Cutting Edge: Proinflammatory and Th2 cytokines synergize to induce thymic stromal lymphopoietin production by human skin keratinocytes. J Immunol 178, 3373 (2007).

379 Traidl-Hoffmann, C. et al., Pollen-associated phytoprostanes inhibit dendritic cell interleukin-12 production and augment T helper type 2 cell polarization. J. Exp. Med. 201, 627 (2005).

380 Hessel, E. M. et al., Immunostimulatory oligonucleotides block allergic airway inflammation by inhibiting Th2 cell activation and IgE-mediated cytokine induction. J. Exp. Med. 202, 1563 (2005).

381 Creticos, P. S. et al., Immunotherapy with a ragweed-toll-like receptor 9 agonist vaccine for allergic rhinitis. N. Engl. J. Med. 355, 1445 (2006).

382 Lewkowich, I. P. et al., CD4+CD25+ T cells protect against experimentally induced asthma and alter pulmonary dendritic cell phenotype and function. J. Exp. Med. 202, 1549 (2005).

383 Strickland, D. H. et al., Reversal of airway hyperresponsiveness by induction of airway mucosal CD4+CD25+ regulatory T cells. J. Exp. Med. 203, 2649 (2006).

384 Idzko, M. et al., Inhaled iloprost suppresses the cardinal features of asthma via inhibition of airway dendritic cell function. J. Clin. Invest. 117, 464 (2007).

385 Idzko, M. et al., Local application of FTY720 to the lung abrogates experimental asthma by altering dendritic cell function. J. Clin. Invest. 116, 2935 (2006).

386 Hammad, H. et al., Activation of the D prostanoid 1 receptor suppresses asthma by modulation of lung dendritic cell function and induction of regulatory T cells. J. Exp. Med. 204, 357 (2007).

387 Larsen, C. P., Morris, P. J. &Austyn, J. M., Migration of dendritic leukocytes from cardiac allografts into host spleens: a novel pathway for initiation of rejection. J. Exp. Med. 171, 307 (1990).

388 Larsen, C. P. et al., Migration and maturation of Langerhans cells in skin transplants and explants. J. Exp. Med. 172, 1483 (1990).

389 Larosa, D. F., Rahman, A. H. &Turka, L. A., The innate immune system in allograft rejection and tolerance. J. Immunol. 178, 7503 (2007).

390 Inaba, K. et al., Efficient presentation of phagocytosed cellular fragments on the MHC class II products of dendritic cells. J. Exp. Med. 188, 2163 (1998).

391 Ochando, J. C. et al., Alloantigen-presenting plasmacytoid dendritic cells mediate tolerance to vascularized grafts. Nat. Immunol. 7, 652 (2006).

392 Matsue, H. et al., Contrasting impacts of immunosuppressive agents (rapamycin, FK506, cyclosporin A, and dexamethasone) on bidirectional dendritic cell-T cell interaction during antigen presentation. J. Immunol. 169, 3555 (2002).

393 Chen, T. et al., Cyclosporin A impairs dendritic cell migration by regulating chemokine receptor expression and inhibiting cyclooxygenase-2 expression. Blood 103, 413 (2004).

394 Hackstein, H. et al., Rapamycin inhibits IL-4--induced dendritic cell maturation in vitro and dendritic cell mobilization and function in vivo. Blood 101, 4457 (2003).

395 Hackstein, H. & Thomson, A. W., Dendritic cells: emerging pharmacological targets of immunosuppressive drugs. Nat. Rev. Immunol. 4, 24 (2004).

396 Turnquist, H. R. et al., Rapamycin-conditioned dendritic cells are poor stimulators of allogeneic CD4+ T cells, but enrich for antigen-specific Foxp3+ T regulatory cells and promote organ transplant tolerance. J. Immunol. 178, 7018 (2007).

39

Page 40: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

397 Shlomchik, W. D. et al., Prevention of graft versus host disease by inactivation of host antigen- presenting cells. Science (New York, N.Y 285, 412 (1999).

398 Merad, M. et al., Depletion of host Langerhans cells before transplantation of donor alloreactive T cells prevents skin graft-versus-host disease. Nat. Med. 10, 510 (2004).

399 Land, W. G., Innate immunity-mediated allograft rejection and strategies to prevent it. Transplant. Proc. 39, 667 (2007).

400 Merad, M. et al., Langerhans cells renew in the skin throughout life under steady-state conditions. Nat. Immunol. 3, 1135 (2002).

401 Lechler, R. I., Sykes, M., Thomson, A. W. &Turka, L. A., Organ transplantation-how much of the promise has been realized? Nat. Med. 11, 605 (2005).

402 Lan, Y. Y. et al., "Alternatively activated" dendritic cells preferentially secrete IL-10, expand Foxp3+CD4+ T cells, and induce long-term organ allograft survival in combination with CTLA4-Ig. J. Immunol. 177, 5868 (2006).

403 Morelli, A. E. & Thomson, A. W., Tolerogenic dendritic cells and the quest for transplant tolerance. Nat. Rev. Immunol. 7, 610 (2007).

404 Cooper, M. A., Fehniger, T. A., Fuchs, A., Colonna, M. &Caligiuri, M. A., NK cell and DC interactions. Trends Immunol. 25, 47 (2004).

405 Ferlazzo, G. & Münz, C., NK cell compartments and their activation by dendritic cells. J. Immunol. 172, 1333 (2004).

406 Yu, G., Xu, X., Vu, M. D., Kilpatrick, E. D. &Li, X. C., NK cells promote transplant tolerance by killing donor antigen-presenting cells. J. Exp. Med. 203, 1851 (2006).

407 Vlad, G., Cortesini, R. &Suciu-Foca, N., License to heal: bidirectional interaction of antigen-specific regulatory T cells and tolerogenic APC. J. Immunol. 174, 5907 (2005).

408 Mestas, J. & Hughes, C. C. W., Of mice and men: differences between mouse and human immunology. J. Immunol. 172, 2731 (2004).

409 van Rijn, R. S. et al., A new xenograft model for graft-versus-host disease by intravenous transfer of human peripheral blood mononuclear cells in RAG2-/- γc-/- double-mutant mice. Blood 102, 2522 (2003).

410 Traggiai, E. et al., Development of a human adaptive immune system in cord blood cell-transplanted mice. Science (New York, N.Y 304, 104 (2004).

411 Melkus, M. W. et al., Humanized mice mount specific adaptive and innate immune responses to EBV and TSST-1. Nat. Med. 12, 1316 (2006).

412 Macchiarini, F., Manz, M. G., Palucka, A. K. &Shultz, L. D., Humanized mice: are we there yet? J. Exp. Med. 202, 1307 (2005).

413 Baenziger, S. et al., Disseminated and sustained HIV infection in CD34+ cord blood cell-transplanted Rag2-/-γc-/- mice. Proc. Natl. Acad. Sci. USA 103, 15951 (2006).

414 Legrand, N., Weijer, K. &Spits, H., Experimental models to study development and function of the human immune system in vivo. J. Immunol. 176, 2053 (2006).

415 Palucka, A. K. et al., Human dendritic cell subsets in NOD/SCID mice engrafted with CD34+ hematopoietic progenitors. Blood 102, 3302 (2003).

416 Shultz, L. D., Ishikawa, F. &Greiner, D. L., Humanized mice in translational biomedical research. Nat. Rev. Immunol. 7, 118 (2007).

417 Steinman, R. M. & Mellman, I., Immunotherapy: bewitched, bothered, and bewildered no more. Science (New York, N.Y 305, 197 (2004).

418 Nathan, D. G., The several Cs of translational clinical research. J. Clin. Invest. 115, 795 (2005).

40

Page 41: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

419 Nishibu, A. et al., Behavioral responses of epidermal Langerhans cells in situ to local pathological stimuli. J. Invest. Dermatol. 126, 787 (2006).

420 Fleeton, M. N. et al., Peyer's patch dendritic cells process viral antigen from apoptotic epithelial cells in the intestine of reovirus-infected mice. J. Exp. Med. 200, 235 (2004).

421 Holt, P. G., Schon-Hegrad, M. A. &Oliver, J., MHC class II antigen-bearing dendritic cells in pulmonary tissues of the rat. Regulation of antigen presentation activity by endogenous macrophage populations. J. Exp. Med. 167, 262 (1987).

422 Uematsu, S. et al., Detection of pathogenic intestinal bacteria by Toll-like receptor 5 on intestinal CD11c+ lamina propria cells. Nat. Immunol. 7, 868 (2006).

423 Tezuka, H. et al., Regulation of IgA production by naturally occurring TNF/iNOS-producing dendritic cells. Nature 448, 929 (2007).

424 Rescigno, M. et al., Dendritic cells express tight junction proteins and penetrate gut epithelial monolayers to sample bacteria. Nat. Immunol. 2, 361 (2001).

425 Iwata, M. et al., Retinoic acid imprints gut-homing specificity on T cells. Immunity 21, 527 (2004).

426 Sigmundsdottir, H. et al., DCs metabolize sunlight-induced vitamin D3 to 'program' T cell attraction to the epidermal chemokine CCL27. Nat. Immunol. 8, 285 (2007).

427 Mebius, R. E., Vitamins in control of lymphocyte migration. Nat. Immunol. 8, 229 (2007). 428 Coombes, J. L. et al., A functionally specialized population of mucosal CD103+ DCs

induces Foxp3+ regulatory T cells via a TGF-β- and retinoic acid-dependent mechanism. J. Exp. Med. 204, 1757 (2007).

429 Sun, C. M. et al., Small intestine lamina propria dendritic cells promote de novo generation of Foxp3 T reg cells via retinoic acid. J. Exp. Med. 204, 1775 (2007).

430 Benson, M. J., Pino-Lagos, K., Rosemblatt, M. &Noelle, R. J., All-trans retinoic acid mediates enhanced T reg cell growth, differentiation, and gut homing in the face of high levels of co-stimulation. J. Exp. Med. 204, 1765 (2007).

431 Saurer, L., McCullough, K. C. &Summerfield, A., In vitro induction of mucosa-type dendritic cells by all-trans retinoic Acid. J. Immunol. 179, 3504 (2007).

432 Angeli, V. et al., B cell-driven lymphangiogenesis in inflamed lymph nodes enhances dendritic cell mobilization. Immunity 24, 203 (2006).

433 Webster, B. et al., Regulation of lymph node vascular growth by dendritic cells. J. Exp. Med. 203, 1903 (2006).

434 Bousso, P. & Robey, E., Dynamics of CD8+ T cell priming by dendritic cells in intact lymph nodes. Nat. Immunol. 4, 579 (2003).

435 Miller, M. J., Safrina, O., Parker, I. &Cahalan, M. D., Imaging the single cell dynamics of CD4+ T cell activation by dendritic cells in lymph nodes. J. Exp. Med. 200, 847 (2004).

436 Germain, R. N. & Jenkins, M. K., In vivo antigen presentation. Curr. Opin. Immunol. 16, 120 (2004).

437 Castellino, F. et al., Chemokines enhance immunity by guiding naive CD8 T cells to sites of CD4+ T cell-dendritic cell interaction. Nature 440, 890 (2006).

438 Shakhar, G. et al., Stable T cell–dendritic cell interactions precede the development of both tolerance and immunity in vivo. Nat. Immunol. 6, 707 (2005).

439 Mora, J. R. et al., Generation of gut-homing IgA-secreting B cells by intestinal dendritic cells. Science (New York, N.Y 314, 1157 (2006).

440 Qi, H., Egen, J. G., Huang, A. Y. &Germain, R. N., Extrafollicular activation of lymph node B cells by antigen-bearing dendritic cells. Science (New York, N.Y 312, 1672 (2006).

41

Page 42: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

441 Balazs, M., Martin, F., Zhou, T. &Kearney, J. F., Blood dendritic cells interact with splenic marginal zone B cells to initiate T-independent immune responses. Immunity 17, 341 (2002).

442 Bergtold, A., Desai, D. D., Gavhane, A. &Clynes, R., Cell surface recycling of internalized antigen permits dendritic cell priming of B cells. Immunity 23, 503 (2005).

443 Dubois, B. et al., Critical role of IL-12 in dendritic cell-induced differentiation of naive B lymphocytes. J. Immunol. 161, 2223 (1998).

444 Dubois, B. et al., Dendritic cells enhance growth and differentiation of CD40-activated B lymphocytes. J. Exp. Med. 185, 941 (1997).

445 Fayette, J. et al., Human dendritic cells skew isotype switching of CD40-activated naive B cells towards IgA1 and IgA2. J. Exp. Med. 185, 1909 (1997).

446 Jego, G. et al., Plasmacytoid dendritic cells induce plasma cell differentiation through type I interferon and interleukin 6. Immunity 19, 225 (2003).

447 Litinskiy, M. B. et al., DCs induce CD40-independent immunoglobulin class switching through BLyS and APRIL. Nat. Immunol. 3, 822 (2002).

448 Sornasse, T. et al., Antigen-pulsed dendritic cells can efficiently induce an antibody response in vivo. J. Exp. Med. 175, 15 (1992).

449 Bajenoff, M. et al., Natural killer cell behavior in lymph nodes revealed by static and real-time imaging. J. Exp. Med. 203, 619 (2006).

450 Delamarre, L., Pack, M., Chang, H., Mellman, I. &Trombetta, E. S., Differential lysosomal proteolysis in antigen-presenting cells determines antigen fate. Science (New York, N.Y 307, 1630 (2005).

451 Jiang, W. et al., The receptor DEC-205 expressed by dendritic cells and thymic epithelial cells is involved in antigen processing. Nature 375, 151 (1995).

452 Mahnke, K. et al., The dendritic cell receptor for endocytosis, DEC-205, can recycle and enhance antigen presentation via major histocompatibility complex class II-positive lysosomal compartments. J. Cell Biol. 151, 673 (2000).

453 Turley, S. J. et al., Transport of peptide-MHC class II complexes in developing dendritic cells. Science (New York, N.Y 288, 522 (2000).

454 Inaba, K. et al., The formation of immunogenic MHC class II- peptide ligands in lysosomal compartments of dendritic cells is regulated by inflammatory stimuli. J. Exp. Med. 191, 927 (2000).

455 Lizee, G. et al., Control of dendritic cell cross-presentation by the major histocompatibility complex class I cytoplasmic domain. Nat. Immunol. 4, 1065 (2003).

456 Savina, A. et al., NOX2 controls phagosomal pH to regulate antigen processing during crosspresentation by dendritic cells. Cell 126, 205 (2006).

457 Guiducci, C. et al., Properties regulating the nature of the plasmacytoid dendritic cell response to Toll-like receptor 9 activation. J. Exp. Med. 203, 1999 (2006).

458 Burgdorf, S., Kautz, A., Bohnert, V., Knolle, P. A. &Kurts, C., Distinct pathways of antigen uptake and intracellular routing in CD4 and CD8 T cell activation. Science (New York, N.Y 316, 612 (2007).

459 Münz, C. et al., Mature myeloid dendritic cell subsets have distinct roles for activation and viability of circulating human natural killer cells. Blood 105, 266 (2005).

460 Zhou, D. et al., Editing of CD1d-bound lipid antigens by endosomal lipid transfer proteins. Science (New York, N.Y 303, 523 (2004).

42

Page 43: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

461 Szatmari, I. et al., Activation of PPARψ specifies a dendritic cell subtype capable of enhances induction of iNKT cell expansion. Immun. 21, 95 (2004).

462 Szatmari, I. et al., PPARγ controls CD1d expression by turning on retinoic acid synthesis in developing human dendritic cells. J. Exp. Med. 203, 2351 (2006).

463 Wykes, M., Pombo, A., Jenkins, C. &MacPherson, G. G., Dendritic cells interact directly with naive B lymphocytes to transfer antigen and initiate class switching in a primary T-dependent response. J. Immunol. 161, 1313 (1998).

464 Colino, J., Shen, Y. &Snapper, C. M., Dendritic cells pulsed with intact Streptococcus pneumoniae elicit both protein- and polysaccharide-specific immunoglobulin isotype responses in vivo through distinct mechanisms. J. Exp. Med. 195, 1 (2001).

465 Geijtenbeek, T. B. H. et al., Identification of DC-SIGN, a novel dendritic cell-specific ICAM-3 receptor that supports primary immune responses. Cell 100, 575 (2000).

466 Valladeau, J. et al., Identification of mouse langerin/CD207 in Langerhans cells and some dendritic cells of lymphoid tissues. J. Immunol. 168, 782 (2002).

467 Valladeau, J. et al., Immature human dendritic cells express asialoglycoprotein receptor isoforms for efficient receptor-mediated endocytosis. J. Immunol. 167, 5767 (2001).

468 Bates, E. E. et al., APCs express DCIR, a novel C-type lectin surface receptor containing an immunoreceptor tyrosine-based inhibitory motif. J. Immunol. 163, 1973 (1999).

469 Geijtenbeek, T. B. et al., Mycobacteria target DC-SIGN to suppress dendritic cell function. J. Exp. Med. 197, 7 (2003).

470 Tailleux, L. et al., DC-SIGN is the major Mycobacterium tuberculosis receptor on human dendritic cells. J. Exp. Med. 197, 121 (2003).

471 Cambi, A. et al., The C-type lectin DC-SIGN (CD209) is an antigen-uptake receptor for Candida albicans on dendritic cells. Eur. J. Immunol. 33, 532 (2003).

472 van Kooyk, Y. & Geijtenbeek, T. B., DC-SIGN: escape mechanism for pathogens. Nat. Rev. Immunol. 3, 697 (2003).

473 Hodges, A. et al., Activation of the lectin DC-SIGN induces an immature dendritic cell phenotype triggering Rho-GTPase activity required for HIV-1 replication. Nat. Immunol. 8, 569 (2007).

474 Gringhuis, S. I. et al., C-type lectin DC-SIGN modulates Toll-like receptor signaling via Raf-1 kinase-dependent acetylation of transcription factor NF-κB. Immunity 26, 605 (2007).

475 de Witte, L. et al., Langerin is a natural barrier to HIV-1 transmission by Langerhans cells. Nat. Med. 13, 367 (2007).

476 Bonifaz, L. C. et al., In vivo targeting of antigens to maturing dendritic cells via the DEC-205 receptor improves T cell vaccination. J. Exp. Med. 199, 815 (2004).

477 Albert, M. L., Sauter, B. &Bhardwaj, N., Dendritic cells acquire antigen from apoptotic cells and induce class I-restricted CTLs. Nature 392, 86 (1998).

478 den Haan, J., Lehar, S. &Bevan, M., CD8+ but not CD8- dendritic cells cross-prime cytotoxic T cells in vivo. J. Exp. Med. 192, 1685 (2000).

479 Jung, S. et al., In vivo depletion of CD11c+ dendritic cells abrogation priming of CD8+ T cells by exogenous cell-associated antigens. Immunity 17, 211 (2002).

480 Heath, W. R. et al., Cross-presentation, dendritic cell subsets, and the generation of immunity to cellular antigens. Immunol. Rev. 199, 9 (2004).

43

Page 44: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

481 Belz, G. T., Shortman, K., Bevan, M. J. &Heath, W. R., CD8α+ dendritic cells selectively present MHC class I-restricted noncytolytic viral and intracellular bacterial antigens in vivo. J. Immunol. 175, 196 (2005).

482 Le Borgne, M. et al., Dendritic cellsr rapidly recruited into epithelial tissues via CCR6/CCL20 are responsible for CD8(+) T cell crosspriming in vivo. Immunity 24, 191 (2006).

483 Rock, K. L. & Goldberg, A. L., Degradation of cell proteins and the generation of MHC class I- presented peptides. Annu. Rev. Immunol. 17, 739 (1999).

484 Schulz, O. & Reis E Sousa, C., Cross-presentation of cell-associated antigens by CD8α+ dendritic cells is attributable to their ability to internalize dead cells. Immunol. 107, 183 (2002).

485 Turley, S., Poirot, L., Hattori, M., Benoist, C. &Mathis, D., Physiological β cell death triggers priming of self-reactive T cells by dendritic cells in a type-1 diabetes model. J. Exp. Med. 198, 1527 (2003).

486 Fonteneau, J.-F., Larsson, M. &Bhardwaj, N., Interactions between dead cells and dendritic cells in the induction of antiviral CTL responses. Curr. Opin. Immunol. 14, 471 (2002).

487 Nimmerjahn, F. & Ravetch, J. V., Fcγ receptors: old friends and new family members. Immunity 24, 19 (2006).

488 Nimmerjahn, F. & Ravetch, J. V., The antiinflammatory activity of IgG: the intravenous IgG paradox. J. Exp. Med. 204, 11 (2007).

489 Sher, A., Pearce, E. &Kaye, P., Shaping the immune response to parasites: role of dendritic cells. Curr. Opin. Immunol. 15, 421 (2003).

490 Pulendran, B., Variegation of the immune response with dendritic cells and pathogen recognition receptors. J. Immunol. 174, 2457 (2005).

491 Thurnher, M., Lipids in dendritic cell biology: messengers, effectors, and antigens. J. Leukoc. Biol. 81, 154 (2007).

492 Kapsenberg, M. L., Dendritic-cell control of pathogen-driven T-cell polarization. Nat. Rev. Immunol. 3, 984 (2003).

493 Stranges, P. B. et al., Elimination of antigen-presenting cells and autoreactive T cells by fas contributes to prevention of autoimmunity. Immunity 26, 629 (2007).

494 Granelli-Piperno, A., Pope, M., Inaba, K. &Steinman, R. M., Coexpression of NF-κB/Regl and Sp1 transcription factors in human immunodeficiency virus 1-induced, dendritic cell-T-cell syncytia. Proc. Natl. Acad. Sci. USA 92, 10944 (1995).

495 Hara, H. et al., The adaptor protein CARD9 is essential for the activation of myeloid cells through ITAM-associated and Toll-like receptors. Nat. Immunol. 8, 619 (2007).

496 Wang, J. et al., Distinct roles of different NF-{kappa}B subunits in regulating inflammatory and T cell stimulatory gene expression in dendritic cells. J. Immunol. 178, 6777 (2007).

497 Cella, M., Facchetti, F., Lanzavecchia, A. &Colonna, M., Plasmacytoid dendritic cells activated by influenza virus and CD40L drive a potent Th1 polarization. Nat. Immunol. 1, 305 (2000).

498 Fossiez, F. et al., T cell interleukin-17 induces stromal cells to produce proinflammatory and hematopoietic cytokines. J. Exp. Med. 183, 2593 (1996).

499 Bettelli, E. et al., Reciprocal developmental pathways for the generation of pathogenic effector TH17 and regulatory T cells. Nature 441, 235 (2006).

44

Page 45: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

500 Weaver, C. T., Hatton, R. D., Mangan, P. R. &Harrington, L. E., IL-17 family cytokines and the expanding diversity of effector T cell lineages. Annu. Rev. Immunol. 25, 821 (2007).

501 Park, H. et al., A distinct lineage of CD4 T cells regulates tissue inflammation by producing interleukin 17. Nat. Immunol. 6, 1133 (2005).

502 Acosta-Rodriguez, E. V. et al., Surface phenotype and antigenic specificity of human interleukin 17-producing T helper memory cells. Nat. Immunol. 8, 639 (2007).

503 Martin, E., O'Sullivan, B., Low, P. &Thomas, R., Antigen-specific suppression of a primed immune response by dendritic cells mediated by regulatory T cells secreting interleukin-10. Immunity 18, 155 (2003).

504 Dillon, S. et al., Yeast zymosan, a stimulus for TLR2 and dectin-1, induces regulatory antigen-presenting cells and immunological tolerance. J. Clin. Invest. 116, 916 (2006).

505 Gonzalez-Rey, E., Chorny, A., Fernandez-Martin, A., Ganea, D. &Delgado, M., Vasoactive intestinal peptide generates human tolerogenic dendritic cells that induce CD4 and CD8 regulatory T cells. Blood 107, 3632 (2006).

506 Brinster, C. & Shevach, E. M., Bone marrow-derived dendritic cells reverse the anergic state of CD4+CD25+ T cells without reversing their suppressive function. J. Immunol. 175, 7332 (2005).

507 Ito, T. et al., Plasmacytoid dendritic cells prime IL-10-producing T regulatory cells by inducible costimulator ligand. J. Exp. Med. 204, 105 (2007).

508 Sato, K., Yamashita, N., Baba, M. &Matsuyama, T., Regulatory dendritic cells protect mice from murine acute graft-versus-host disease and leukemia relapse. Immunity 18, 367 (2003).

509 Watanabe, N. et al., Human thymic stromal lymphopoietin promotes dendritic cell-mediated CD4+ T cell homeostatic expansion. Nat. Immunol. 5, 426 (2004).

510 Ohteki, T. et al., Essential roles of DC-derived IL-15 as a mediator of inflammatory responses in vivo. J. Exp. Med 203, 2329 (2006).

511 Granucci, F. et al., Inducible IL-2 production by dendritic cells revealed by global gene expression analysis. Nat. Immunol. 2, 882 (2001).

512 Edwards, A. D. et al., Relationships among murine CD11chigh dendritic cell subsets as revealed by baseline gene expression patterns. J. Immunol. 171, 47 (2003).

513 Reis e Sousa, C. et al., In vivo microbial stimulation induces rapid CD40L- independent production of IL-12 by dendritic cells and their re-distribution to T cell areas. J. Exp. Med. 186, 1819 (1997).

514 Hochrein, H. et al., Differential production of IL-12, IFN-α, and IFN-γ by mouse dendritic cell subsets. J. Immunol. 166, 5448 (2001).

515 Diebold, S. S. et al., Viral infection switches non-plasmacytoid dendritic cells into high interferon producers. Nature 424, 324 (2003).

516 Lopez, C. B. et al., TLR-independent induction of dendritic cell maturation and adaptive immunity by negative-strand RNA viruses. J. Immunol. 173, 6882 (2004).

517 Johansson, C. et al., Type I interferons produced by hematopoietic cells protect mice against lethal infection by mammalian reovirus. J. Exp. Med. 204, 1349 (2007).

518 Le Bon, A. et al., Direct stimulation of T cells by type I IFN enhances the CD8+ T cell response during cross-priming. J. Immunol. 176, 4682 (2006).

519 Stetson, D. B. & Medzhitov, R., Type I interferons in host defense. Immunity 25, 373 (2006).

45

Page 46: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

520 Sporri, R. & Reis e Sousa, C., Inflammatory mediators are insufficient for full dendritic cell activation and promote expansion of CD4+ T cell populations lacking helper function. Nat. Immunol. 6, 163 (2005).

521 Nolte, M. A., Leibundgut-Landmann, S., Joffre, O. &Sousa, C. R., Dendritic cell quiescence during systemic inflammation driven by LPS stimulation of radioresistant cells in vivo. J. Exp. Med. 204, 1487 (2007).

522 de Saint-Vis, B. et al., The cytokine profile expressed by human dendritic cells is dependent on cell subtype and mode of activation. J. Immunol. 160, 1666 (1998).

523 Flacher, V. et al., Human Langerhans cells express a specific TLR profile and differentially respond to viruses and gram-positive bacteria. J. Immunol. 177, 7959 (2006).

524 Pulendran, B. et al., Lipopolysaccharides from distinct pathogens induce different classes of immune responses in vivo. J Immunol 167, 5067 (2001).

525 Sallusto, F. et al., Rapid and coordinated switch in chemokine receptor expression during dendritic cell maturation. Eur. J. Immunol. 28, 2760 (1998).

526 Piqueras, B., Connolly, J., Freitas, H., Palucka, A. K. &Banchereau, J., Upon viral exposure, myeloid and plasmacytoid dendritic cells produce 3 waves of distinct chemokines to recruit immune effectors. Blood 107, 2613 (2006).

527 Aliberti, J., Hieny, S., Reis, E. S. C., Serhan, C. N. &Sher, A., Lipoxin-mediated inhibition of IL-12 production by DCs: a mechanism for regulation of microbial immunity. Nat. Immunol. 3, 76 (2002).

528 Trombetta, E. S., Ebersold, M., Garrett, W., Pypaert, M. &Mellman, I., Activation of lysosomal function during dendritic cell maturation. Science (New York, N.Y 299, 1400 (2003).

529 Sharpe, A. H. & Freeman, G. J., The B7-CD28 superfamily. Nat. Rev. Immunol. 2, 116 (2002).

530 Watts, T. H., TNF/TNFR family members in costimulation of T cell responses. Annu. Rev. Immunol. 23, 23 (2005).

531 Amsen, D. et al., Instruction of distinct CD4 T helper cell fates by different notch ligands on antigen-presenting cells. Cell 117, 515 (2004).

532 Skokos, D. & Nussenzweig, M. C., CD8- DCs induce IL-12-independent Th1 differentiation through Delta 4 Notch-like ligand in response to bacterial LPS. J. Exp. Med. 204, 1525 (2007).

533 Steinman, R. M. & Hemmi, H., Dendritic cells: translating innate to adaptive immunity. Curr. Top. Microbiol. Immunol. 311, 17 (2006).

534 Summers-Deluca, L. E. et al., Expression of lymphotoxin-αβ on antigen-specific T cells is required for DC function. J. Exp. Med. 204, 1071 (2007).

535 Taraban, V. Y., Rowley, T. F. &Al-Shamkhani, A., A critical role for CD70 in CD8 T cell priming by CD40-licensed APCs. J. Immunol. 173, 6542 (2004).

536 Ito, T. et al., TSLP-activated dendritic cells induce an inflammatory T helper type 2 cell response through OX40 ligand. J. Exp. Med. 202, 1213 (2005).

537 Grouard, G. et al., The enigmatic plasmacytoid T cells develop into dendritic cells with IL-3 and CD40-ligand. J. Exp. Med. 185, 1101 (1997).

538 Colonna, M., Trinchieri, G. &Liu, Y. J., Plasmacytoid dendritic cells in immunity. Nat. Immunol. 5, 1219 (2004).

46

Page 47: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

539 Liu, Y. J., IPC: professional type 1 interferon-producing cells and plasmacytoid dendritic cell precursors. Annu. Rev. Immunol. 23, 275 (2005).

540 De Heer, H. J. et al., Essential role of lung plasmacytoid dendritic cells in preventing asthmatic reactions to harmless inhaled antigen. J. Exp. Med. 200, 89 (2004).

541 Kadowaki, N. et al., Subsets of human dendritic cell precursors express different toll-like receptors and respond to different microbial antigens. J. Exp. Med. 194, 863 (2001).

542 Jarrossay, D., Napolitani, G., Colonna, M., Sallusto, F. &Lanzavecchia, A., Specialization and complementarity in microbial molecule recognition by human myeloid and plasmacytoid dendritic cells. Eur. J. Immunol. 31, 3388 (2001).

543 Vollmer, J. et al., Immune stimulation mediated by autoantigen binding sites within small nuclear RNAs involves Toll-like receptors 7 and 8. J. Exp. Med. 202, 1575 (2005).

544 Barrat, F. J. et al., Nucleic acids of mammalian origin can act as endogenous ligands for Toll-like receptors and may promote systemic lupus erythematosus. J. Exp. Med. 202, 1131 (2005).

545 Lopez, C. B., Garcia-Sastre, A., Williams, B. R. G. &Moran, T. M., Type 1 interferon induction pathway, but not released interferon, participates in the maturation of dendritic cells induced by negative-strand RNA viruses. J. Infect. Dis. 187, 1126 (2003).

546 Kato, H. et al., Cell type-specific involvement of RIG-I in antiviral response. Immunity 23, 19 (2005).

547 Valladeau, J. et al., Langerin, a novel C-type lectin specific to Langerhans cells, is an endocytic receptor that induces the formation of Birbeck granules. Immunity 12, 71 (2000).

548 Caux, C. et al., CD34+ hematopoietic progenitors from human cord blood differentiate along two independent dendritic cell pathways in response to granulocyte-macrophage colony-stimulating factor plus tumor necrosis factor α. Blood 90, 1458 (1997).

549 Ebner, S. et al., Expression of C-type lectin receptors by subsets of dendritic cells in human skin. Int. Immunol. 16, 877 (2004).

550 Caux, C. et al., CD34+ hematopoietic progenitors from human cord blood differentiate along two independent dendritic cell pathways in response to GM-CSF+ TNF α. J. Exp. Med. 184, 695 (1996).

551 Itano, A. A. et al., Distinct dendritic cell populations sequentially present antigen to CD4 T cells and stimulate different aspects of cell-mediated immunity. Immunity 19, 47 (2003).

552 Kissenpfennig, A. et al., Dynamics and function of Langerhans cells in vivo: dermal dendritic cells colonize lymph node areas distinct from slower migrating Langerhans cells. Immunity 22, 643 (2005).

553 Shortman, K. & Liu, Y.-J., Mouse and human dendritic cell subtypes. Nat. Rev. Immunol. 2, 151 (2002).

554 Del Rio, M. L., Rodriguez-Barbosa, J. I., Kremmer, E. &Forster, R., CD103- and CD103+ bronchial lymph node dendritic cells are specialized in presenting and cross-presenting innocuous antigen to CD4+ and CD8+ T Cells. J. Immunol. 178, 6861 (2007).

555 Yoneyama, H. et al., Plasmacytoid DCs help lymph node DCs to induce anti-HSV CTLs. J. Exp. Med. 202, 425 (2005).

556 Kuwajima, S. et al., Interleukin 15-dependent crosstalk between conventional and plasmacytoid dendritic cells is essential for CpG-induced immune activation. Nat. Immunol. 7, 740 (2006).

47

Page 48: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

557 D'Amico, A. & Wu, L., The early progenitors of mouse dendritic cells and plasmacytoid predendritic cells are within the bone marrow hemopoietic precursors expressing Flt3. J. Exp. Med. 198, 293 (2003).

558 Karsunky, H., Merad, M., Cozzio, A., Weissman, I. L. &Manz, M. G., Flt3 ligand regulates dendritic cell development from flt3+ lymphoid and myeloid-committed progenitors to flt3+ dendritic cells in vivo. J. Exp. Med. 198, 305 (2003).

559 Chiodoni, C. et al., Dendritic cells infiltrating tumors cotransduced with granulocyte/macrophage colony-stimulating factor (GM-CSF) and CD40 ligand genes take up and present endogenous tumor-associated antigens, and prime naive mice for a cytotoxic T lymphocyte response. J. Exp. Med. 190, 125 (1999).

560 Nasi, M. L. et al., Intradermal injection of granulocyte-macrophage colony-stimulating factor (GM-CSF) in patients with metastic melanoma recruits dendritic cells. Cytokines, Cellular & Molecular Therapy 5, 139 (1999).

561 Basak, S. K. et al., Increased dendritic cell number and function following continuous in vivo infusion of granulocyte macrophage-colony-stimulating factor and interleukin-4. Blood 99, 2869 (2002).

562 Steptoe, R. J., Ritchie, J. M. &Harrison, L. C., Increased generation of dendritic cells from myeloid progenitors in autoimmune-prone nonobese diabetic mice. J. Immunol. 168, 5032 (2002).

563 Schneeberger, A. et al., Granulocyte-macrophage colony-stimulating factor-based melanoma cell vaccines immunize syngeneic and allogeneic recipients via host dendritic cells. J. Immunol. 171, 5180 (2003).

564 Leon, B., Lopez-Bravo, M. &Ardavin, C., Monocyte-derived dendritic cells formed at the infection site control the induction of protective T helper 1 responses against Leishmania. Immunity 26, 519 (2007).

565 Liu, K. et al., Origin of dendritic cells in peripheral lymphoid organs of mice. Nat. Immunol. 8, 578 (2007).

566 Fogg, D. K. et al., A clonogenic bone marrow progenitor specific for macrophages and dendritic cells. Science (New York, N.Y 311, 83 (2006).

567 Naik, S. H. et al., Intrasplenic steady-state dendritic cell precursors that are distinct from monocytes. Nat. Immunol. 7, 663 (2006).

568 Varol, C. et al., Monocytes give rise to mucosal, but not splenic, conventional dendritic cells. J. Exp. Med. 204, 171 (2007).

569 Reis e Sousa, C., Activation of dendritic cells: translating innate into adaptive immunity. Curr. Opin. Immunol. 16, 21 (2004).

570 Andoniou, C. E. et al., Interaction between conventional dendritic cells and natural killer cells is integral to the activation of effective antiviral immunity. Nat. Immunol. 6, 1011 (2005).

571 Leslie, D. S. et al., CD1-mediated γ/δ T cell maturation of dendritic cells. J. Exp. Med. 196, 1575 (2002).

572 Münz, C., Steinman, R. M. &Fujii, S., Dendritic cell maturation by innate lymphocytes: coordinated stimulation of innate and adaptive immunity. J. Exp. Med. 202, 203 (2005).

573 Conti, L. et al., Reciprocal activating interaction between dendritic cells and pamidronate-stimulated γδ T cells: Role of CD86 and inflammatory cytokines. J. Immunol. 174, 252 (2005).

48

Page 49: recognize responses - Rockefeller Universitylab.rockefeller.edu/steinman/pdf/banchereau_nature_07.pdf · Many uptake receptors on DCs are lectins with carbohydrate recognition capacity

49

574 Ludwig, I. S., Geijtenbeek, T. B. &van Kooyk, Y., Two way communication between neutrophils and dendritic cells. Curr. Opin. Pharmacol. 6, 408 (2006).

575 van Gisbergen, K. P., Sanchez-Hernandez, M., Geijtenbeek, T. B. &van Kooyk, Y., Neutrophils mediate immune modulation of dendritic cells through glycosylation-dependent interactions between Mac-1 and DC-SIGN. J. Exp. Med. 201, 1281 (2005).

576 Chomarat, P., Dantin, C., Bennett, L., Banchereau, J. &Palucka, A. K., TNF skews monocyte differentiation from macrophages to dendritic cells. J. Immunol. 171, 2262 (2003).

577 Luft, T. et al., Type I IFNs enhance the terminal differentiation of dendritic cells. J. Immunol. 161, 1947 (1998).

578 Santini, S. M. et al., Type I interferon as a powerful adjuvant for monocyte-derived dendritic cell development and activity in vitro and in Hu-PBL-SCID mice. J. Exp. Med. 191, 1777 (2000).

579 Liu, Y. J. et al., TSLP: An epithelial cell cytokine that regulates T cell differentiation by conditioning dendritic cell maturation. Annu. Rev. Immunol. 25, 193 (2007).

580 Gallucci, S., Lolkema, M. &Matzinger, P., Natural adjuvants: endogenous activators of dendritic cells. Nat. Med. 5, 1249 (1999).

581 Sauter, B. et al., Consequences of cell death. Exposure to necrotic tumor cells, but not primary tissue cells or apoptotic cells, induces the maturation of immunostimulatory dendritic cells. J. Exp. Med. 191, 423 (2000).

582 Shi, Y., Evans, J. E. &Rock, K. L., Molecular identification of a danger signal that alerts the immune system to dying cells. Nature 425, 516 (2003).

583 Martinon, F., Petrilli, V., Mayor, A., Tardivel, A. &Tschopp, J., Gout-associated uric acid crystals activate the NALP3 inflammasome. Nature 440, 237 (2006).

584 Scaffidi, P., Misteli, T. &Bianchi, M. E., Release of chromatin protein HMGB1 by necrotic cells triggers inflammation. Nature 418, 191 (2002).

585 Mazzoni, A., Siraganian, R. P., Leifer, C. A. &Segal, D. M., Dendritic cell modulation by mast cells controls the TH1/TH2 balance in responding T cells. J. Immunol. 177, 3577 (2006).

586 Singh-Jasuja, H. et al., The heat shock protein gp96 induces maturation of dendritic cells and down-regulation of its receptor. Eur. J. Immunol. 30, 2211 (2000).

587 Messmer, D. et al., High mobility group box protein 1: an endogenous signal for dendritic cell maturation and Th1 polarization. J. Immunol. 173, 307 (2004).

588 Lotze, M. T. & Tracey, K. J., High-mobility group box 1 protein (HMGB1): nuclear weapon in the immune arsenal. Nat. Rev. Immunol. 5, 331 (2005).

589 Biragyn, A. et al., Toll-like receptor 4-dependent activation of dendritic cells by β-defensin 2. Science (New York, N.Y 298, 1025 (2002).

590 Idzko, M. et al., Extracellular ATP triggers and maintains asthmatic airway inflammation by activating dendritic cells. Nat. Med. 13, 913 (2007).

591 Fritz, J. H. et al., Synergistic stimulation of human monocytes and dendritic cells by Toll-like receptor 4 and NOD1- and NOD2-activating agonists. Eur. J. Immunol. 35, 2459 (2005).