Skip to main content
Oncoimmunology logoLink to Oncoimmunology
. 2012 Jul 1;1(4):493–506. doi: 10.4161/onci.20459

Trial Watch

Immunostimulatory cytokines

Erika Vacchelli 1,2,, Lorenzo Galluzzi 3,4,, Alexander Eggermont 3, Jerome Galon 4,5,6,7,8, Eric Tartour 5,7,9, Laurence Zitvogel 3,10, Guido Kroemer 1,4,6,7,11,*
PMCID: PMC3382908  PMID: 22754768

Abstract

During the last two decades, a number of approaches for the activation of the immune system against cancer has been developed. These include highly specific interventions, such as monoclonal antibodies, vaccines and cell-based therapies, as well as relatively unselective strategies, such as the systemic administration of adjuvants and immunomodulatory cytokines. Cytokines constitute a huge group of proteins that, taken together, regulate not only virtually all the aspects of innate and cognate immunity, but also several other cellular and organismal functions. Cytokines operate via specific transmembrane receptors that are expressed on the plasma membrane of target cells and, depending on multiple variables, can engage autocrine, paracrine or endocrine signaling pathways. The most appropriate term for defining the cytokine network is “pleiotropic”: cytokines are produced by - and operate on - multiple, often overlapping, cell types, triggering context-depend biological outcomes as diverse as cell proliferation, chemotaxis, differentiation, inflammation, elimination of pathogens and cell death. Moreover, cytokines often induce the release of additional cytokines, thereby engaging self-amplificatory or self-inhibitory signaling cascades. In this Trial Watch, we will summarize the biological properties of cytokines and discuss the progress of ongoing clinical studies evaluating their safety and efficacy as immunomodulatory agents against cancer.

Keywords: GM-CSF, IFN, IL-2, TGFβ, TNFα, chemokines

Introduction

The word “cytokine” (derived from a Greek term meaning “to set cells in motion”) is employed to describe a highly heterogeneous group of small signaling (glyco)proteins that nowadays includes more than 130 members.1,2 Cytokines are produced throughout the body by an incredible variety of distinct cells, encompassing virtually all cellular components of the immune system as well as epithelial, endothelial and stromal cells. Cytokine-delivered signals, which - depending on the context - can be autocrine, paracrine or endocrine, are transduced into biological outcomes thanks to specific receptors that are expressed on the surface of target cells.1,2 Perhaps the most appropriate word to characterize the cytokine system is “pleiotropic”: taken together, cytokines exert highly diversified functions, yet in a partially overlapping and redundant fashion. Thus, cytokines can trigger biological outcomes as diverse as proliferation, chemotaxis, differentiation, inflammation, elimination of pathogens and cell death. Moreover, cytokines very often stimulate the production and release of additional cytokines, de facto functioning as part of finely regulated and highly intertwined signaling cascades. Such a pleiotropism reflects not only the heterogeneous identity of cytokines as a group, but also (i) the existence of multiple receptors that can bind the same cytokine with different affinity (which are frequently expressed on different target cells), and (ii) the fact that the biological activity of one cytokine on a specific target cell is highly influenced by the concomitant presence of additional cytokines.1,2

A wide array of adverse conditions, encompassing inflammation, infection by pathogens and tumorigenesis, provokes the secretion of cytokines. In this context, cytokines underlie a host response that aims at minimizing the harmful effects of stress, favoring repair mechanisms and, eventually, restoring homeostasis. Indeed, cytokines are often released in subsequent waves, and the terminal molecules of the cascade normally function to extinguish the stress response, along with the reestablishment of homeostasis. One prominent example of this biological behavior is provided by the systemic response to the administration of lipopolysaccharide (LPS, mimicking widespread bacterial infection). In this model, a rapid secretion of tumor necrosis factor α (TNFα) precedes a wave of interleukin-1β (IL-1β), IL-6, IL-8, IL-17A, IL-18 and interferon γ (IFNγ) (all of which exert potent pro-inflammatory effects, at both local and system levels), followed by a relatively delayed secretion of anti-inflammatory IL-10.3-5 In some instances, however, repair mechanisms are inefficient and fail to resolve the cytokine-inducing stimulus, leading to persistent cytokine production and exacerbated tissue damage. This is particularly relevant for inflammation-driven carcinogenesis, as it implies that the sites of chronic inflammation are a source of potentially mutagenic chemicals (e.g., high levels of reactive oxygen species) as well as of cytokine cocktails that may promote survival, proliferation and angiogenesis.6,7 Taken together, these observations suggest that the administration of immunomodulatory cytokines for eliciting an antitumor immune response should always be carefully weighted not only against their acute toxicity (in some cases resembling a state of severe infection) but also against the possibility to exacerbate inflammation-associated oncogenesis.6 In addition, some cytokines are endowed with potent mitogenic functions, de facto precluding their use as anticancer agents (see below).

During the three decades, there have been multiple attempts to classify cytokines based on structural and/or functional parameters. Thus, at some stage, terms including “lymphokines,” “interleukins” and “chemokines” have been introduced to indicate cytokines that are produced by lymphocytes, cytokines that mediate the communication between leucocytes, and cytokines that stimulate chemotaxis, respectively.1,2 Today, according to the Kyoto Encyclopedia of Genes and Genomes (www.genome.jp/kegg/), cytokines can be cataloged into 9 main groups: (1) chemokines, small cytokines with chemotactic activities that can further be subdivided into “C,” “CC,” “CXC” and “CX3C” chemokines, depending on the number and arrangement of conserved cysteine residues; (2) hematopoietic growth factors (or hematopoietins), i.e., cytokines with a prominent role in hematopoiesis, which can be further grouped - based on their respective receptors - into “gp130 (IL6ST) shared,” “IL13RA1 shared,” “IL12RB1 shared,” “IL3RB (CSF2RB) shared,” “ILRG shared” and “others”; (3) interleukin-1 family members; (4) interleukin-10 family members; (5) interleukin-17 family members; (6) interferons (IFNs); (7) platelet-derived growth factor (PDGF) family members; (8) transforming growth factor β (TGFβ) family members; and (9) tumor necrosis factors (Table 1).

Table 1. Operational classification of cytokines.

Family Subfamily Members Ref.
Chemokines
C subfamily
XCL1, XCL2
2 , 8
CC subfamily
CCL1, CCL2, CCL3, CCL4, CCL5, CCL7, CCL8, CCL11, CCL13, CCL14, CCL15, CCL16, CCL17, CCL19, CCL20, CCL21, CCL22, CCL23, CCL24, CCL25, CCL26, CCL27, CCL28
CXC subfamily
CXCL1, CXCL2, CXCL3, CXCL5, CXCL6, CXCL8, CXCL9, CXCL10, CXCL11, CXCL12, CXCL13, CXCL16, PPBP
CX3C subfamily
CX3CL1
Hematopoietins
IL6ST - shared
CLCF1, CNTF, CTF1, IL-11, IL-6, LIF, OSM
17 , 18
IL13RA1 - shared
IL-4, IL-13
IL12RB1 - shared
IL12A, IL23A
IL3RB1 - shared
GM-CSF, IL-3, IL-5
ILRG - shared
IL-2, IL-4, IL-7, IL-9, IL-15, IL-21
Others
G-CSF, EPO, GH1, GH2, LEP, PRL, TSLP, TPO
IL-1s
 
IL-1A, IL-1B, IL-18
43
IL-10s
 
IL-10, IL-19, IL-20, IL-22, IL-24, IL-28A, IL-28B, IL-29
61
IL-17s
 
IL-17A, IL-17B, IL-17C, IL-17D, IL-17E, IL-17F
73
IFNs
 
IFNα1, IFNα2, IFNβ1, IFNε1, IFNγ, IFNκ, IFNω
81 - 83
PDGFs
 
M-CSF, EGF, FLT3L, HGF, KITL,
PLEKHQ1, PGDFA-D, VEGFA-D
89 , 90
TGFβs
 
AMH, BMP2, BMP7, GDF5, INHBA, INHBB,
INHBC, INHBE, TGFβ1, TGFβ2, TGFβ3
103 - 106 , 108
TNFs   CD40L, EDA, FASL, LTA, LTB, RANKL, OX40L,
TNF, TNFSF7, TNFSF8, TNFSF9, TNFSF12,
TNFSF13, TNFSF13B, TNFSF14, TNFSF18, TRAIL
111 - 113

Abbreviations: AMH, anti-Mullerian hormone; BMP, bone morphogenetic protein; CD40L, CD40 ligand; CLC, cardiotrophin-like cytokine factor; CNTF, ciliary neurotrophic factor; CTF, cardiotrophin; EDA, ectodysplasin A; EGF, epidermal growth factor, EPO, erythropoietin; FASL, FAS ligand; FLT3L, FLT3 ligand; GDF, growth differentiation factor preproprotein; GH, growth hormone; G-CSF, granulocyte colony stimulating factor; GM-CSF, granulocyte macrophage colony stimulating factor; HGF, hepatocyte growth factor isoform; IFN, interferon; IL, interleukin; INHB, inhibin β; INHBE, activin β E; KITL, KIT ligand; LEP, leptin; LIF, leukemia inhibitory factor; LT, lymphotoxin; M-CSF, macrophage colony stimulating factor; OSM, oncostatin M, OX40L, OX40 ligand; PDGF, platelet-derived growth factor; PLEKHQ, PH domain-containing protein; PPBP, pro-platelet basic protein; PRL, prolactin; RANKL, RANK ligand; TGFβ, transforming growth factor β, TNF, tumor necrosis factor; TPO, thrombopoietin; TSLP, thymic stromal lymphopoietin, VEGF, vascular endothelial growth factor.

Experts in the field of cytokine research would probably criticize many aspects of this relatively inaccurate but simple classification. Nevertheless, since our aim is not to dwell on the cytokine system in detail, we will use this scheme to provide a conceptual framework to our Trial Watch, in which we will briefly review the biological properties of cytokines and discuss the progress of ongoing (started after January 2008) clinical trials investigating their safety and efficacy as immunomodulatory agents against cancer. In line with this objective, our Trial Watch will deliberately disregard the use of cytokines for hematopoietic reconstitution upon chemotherapy or transplantation. Of note, only 2 cytokines out of more than 130 are nowadays approved by FDA for anticancer therapy: IFNα2b and IL-2 (Table 2).

Table 2. Currently approved cytokines for oncological indications*.

Agent Commercial names   Oncological indications
IFNα2b
Intron A, Roferon-A
 
Cervical intraepithelial neoplasia, CML, FL, HCL, melanoma, MM
IL-2 Aldesleukin, Proleukin   Metastatic melanoma, metastatic renal cell carcinoma

Abbreviations: CML, chronic myeloid leukemia; FL, follicular lymphoma, HCL, hairy cell leukemia, IFN, interferon; IL, interleukin; MM, multiple myeloma. *by FDA or European Medicines Agency (EMA) at the day of submission.

Chemokines

Chemokines are small (8–10 KDa) secreted proteins that share a highly conserved secondary structure and a distinguishing tetracysteine motif, constituting the basis of their systematic classification (see above). Chemokines have first been characterized for their ability to stimulate the migration of cell types as diverse as neutrophils, monocytes, lymphocytes, eosinophils, fibroblasts and keratinocytes. However, it is now clear that chemokines, as a family, contribute to a wide array of physiological and pathological processes, encompassing embryogenesis, immune system function, inflammation, oncogenesis and tumor progression. To date, 44 different human chemokines and 21 G protein-coupled chemokine receptors have been described.2,8

During the last two decades, the inhibition of chemokines or their receptors has been extensively explored - in preclinical models - as a strategy against autoimmune and chronic inflammatory diseases, including rheumatoid arthritis, systemic lupus erythematosus and type 2 diabetes.9 Nowadays, some of these approaches have been translated into clinical trials (www.clinicaltrials.gov). For instance, IL-8 receptor antagonists such as GSK1325756 and SB-656933,10 as well as anti-IL-8 monoclonal antibodies (ABX-IL8) are being tested in patients affected by chronic obstructive pulmonary disease or cystic fibrosis. MDX-1100 (a fully human anti-CXCL10 monoclonal antibody)11 is currently under evaluation for the therapy of rheumatoid arthritis and ulcerative colitis. Moreover, strategies for the inhibition of the chemokine receptor CCR2, including the monoclonal antibody MLN120212 and the PEGylated antisense oligonucleotide NOX-E3613 are being explored in phase I/II clinical trials to treat multiple sclerosis and type 2 diabetes, respectively. Of note, maraviroc (an allosteric inhibitor of CCR5 also known as Selzentry) is currently approved by FDA for HIV treatment, as CCR5 is an essential co-receptor for the cellular entry of most HIV strains.14

Opposed to the large number of preclinical and clinical studies evaluating chemokine inhibition as a strategy for autoimmune disorders and chronic inflammation, only a few trials are evaluating the therapeutic potential of the modulation of the chemokine system in cancer patients (www.clinicaltrials.gov). Vaccines based on autologous dendritic cells (DCs) genetically modified to express CCL21 (a small member of the CC subfamily)15 are being investigated, either as single agents or in combination with CD40L-expressing DCs, against melanoma and lung carcinoma. MLN1202 is currently under investigation as a single agent in unspecified metastatic neoplasms. Moreover, the safety and efficacy of a monoclonal antibody that specifically targets the CCR2 ligand (CNTO888) is being tested in metastatic prostate cancer patients (as monotherapy) as well as in patients affected by unspecified solid tumors (in combination with doxorubicin, gemcitabine, paclitaxel + carboplatin, or docetaxel). Reflecting the important role played by its ligand (CCL12) in normal and malignant hematopoiesis,16 an anti-CXCR4 monoclonal antibody (BMS-936564) is being administered to multiple myeloma patients, alone or in combination with lenalidomide, dexamethasone, or bortezomib, as well as to patients with various B-cell neoplasms. Along similar lines, the anti-CCL12 PEGylated antisense oligonucleotide NOX-A12 is under investigation as a single agent to treat multiple myeloma or lymphocytic leukemia patients.

Hematopoietic growth factors

The group of hematopoietic growth factors includes more than 25 cytokines that are required for the maintenance of hematopoietic stem cells, their proliferation, their differentiation into distinct hematopoietic lineages, as well as for the maintenance of a stable equilibrium in the mature hematopoietic compartment. In addition to this homeostatic role, hematopoietins coordinate the organismal response to stimuli that require the expansion of specific hematopoietic cell lineages, (e.g., pathogen infection).17 The spectrum of cells targeted by different hematopoietins can be large, such in the case of the granulocyte monocyte colony stimulating factor (GM-CSF), or relatively restricted, such in the case of erythropoietin (EPO). In both cases, hematopoietins do not simply convey proliferative signals but also stimulate cell survival, differentiation, lineage commitment and the functional activation of mature cells.18

At present, several hematopoietic growth factors including EPO (Epogen, Procrit), IL-2 (Aldesleukin, Proleukin), granulocyte colony stimulating factor (G-CSF, Filgastrim) and GM-CSF (Sargramostim, Molgramostim) are approved by FDA for the treatment of hematological deficiencies including anemia, neutropenia and thrombocytopenia.19,20 Although these conditions are highly relevant in oncological settings, as they often develop as dose-limiting side effects of chemotherapy, we will not discuss here clinical trials in which hematopoietins are employed (i) to prevent/ameliorate post-chemotherapy cytopenia; and (ii) to facilitate immune reconstitution upon bone marrow transplantation. Rather, we will focus our discussion on the modulation of hematopoietins to obtain specific immunostimulatory or direct anticancer effects.

Besides being employed to boost the proliferation of hematopoietic cells, especially in protocols of adaptive cell transfer,21 IL-2 is currently approved by FDA to treat immunosensitive tumors such as melanoma and renal cell carcinoma (RCC) (Table 2). In line with this notion, IL-2 is also included in approximately 50 early (phase I/II) clinical trials that evaluate various combination regimens for the therapy of melanoma and RCC. Furthermore, IL-2 is being extensively studied (19 phase I/II clinical trials) as an off-label (co-)medication against multiple hematological and solid malignancies, including breast cancer (4 trials) and neuroblastoma (2 trials) (Table 3). In 11 of these trials, unmodified recombinant human IL-2 (rhIL-2) is used, while 7 others employ IL-2 fused to a tumor-targeting peptide (such as the recombinant antibody fragments F16 or L19).22,23 In a rather peculiar approach, the safety and efficacy of an attenuated strain of Salmonella typhimurium genetically engineered to express human IL-2 are being investigated in patients with unresectable hepatic metastases from a solid tumor. In a few cases, IL-2-based chimeras are tested as single agents. More often, IL-2 is co-administered with conventional chemotherapeutics or anticancer vaccines (www.clinicaltrials.gov).

Table 3. Clinical trials* on hematopoietins in cancer therapy (main trends).

Hematopoietin Agent Tumor type Trials* Phase Notes Ref.
IL-2
ALT801
Metastatic urothelial cancer
1
I-II
Combined with cisplatin
and gemcitabine
NCT01326871
CD40L/IL-2-expressing tumor cell vaccine
B-CLL
1
I
As single agent
NCT00609076
DI-Leu16-IL2 immunocytokine
Hematological neoplasms
1
I
Combined with rituximab
NCT00720135
F16IL2 immunocytokine
Breast cancer
1
I-II
Combined with paclitaxel
NCT01134250
Breast cancer
Solid tumors
1
I-II
Combined with doxorubicin
NCT01131364
hu14.18-IL2 immunocytokine
NB
1
II
Combined with GM-CSF
and isotretinoin
NCT01334515
rhIL-2
AML
1
II
Combined with famotine
NCT01289678
Breast cancer
1
II
Combined anastrozol, aromasin, femar, pulsed DCs and thymosin 1α
NCT00935558
MDS
1
I-II
Combined with azacitidine
and ceplene
NCT01324960
Melanoma
1
II
Combined with anticancer vaccine
NCT00784524
Metastatic breast cancer
1
II
NCT00784524
NB
1
I
Combined with zoledronic acid
NCT01404702
NHL
1
II
Combined with rituximab
NCT00994643
Plasma cell neoplasms
1
II
Combined with anticancer vaccine
NCT00616720
Various metastatic neoplasms
1
I
Combined with cyclophosphamide and anticancer vaccine
NCT00676949
IL-2-expressing bacteria
Liver cancer
1
I
As single agent
NCT01099631
L19IL2 immunocytokine
Advanced solid tumors
1
I-II
As single agent
NCT01058538
Pancreatic cancer
1
I-II
Combined with gemcitabine
NCT01198522
Microsphere delivery
HNC
1
n.a.
Combined with GM-CSF
and IL-12
NCT00899821
IL-4
IL-4PE
Glioblastoma
1
II
As single agent
NCT00797940
IL-7
rhIL-7
NB
Sarcoma
1
I-II
Combined with anticancer vaccine
NCT00923351
Metastatic breast cancer
1
II
As single agent
NCT01368107
MGN1601
RCC
1
I-II
Genetically modified anticancer vaccine
NCT01265368
IL-12
Ad-IL-12
Metastatic breast cancer
1
I
As single agent
NCT00849459
Ad-RTS-hIL-12
Melanoma
1
I
Combined with activator ligand
NCT01397708
EGEN-001
CRC
1
I-II
Alone or combined with FOLFIRI or FOLFOX
NCT01300858
Reproductive tract cancer
1
I
Combined with doxorubicin
NCT01489371
1
II
As single agent
NCT01118052
rhIL-12
Breast cancer
1
I-II
Combined with DC/tumor cell fusion vaccine
NCT00622401
Melanoma
1
II
Combined with daclizumab and anticancer vaccine
NCT01307618
SCC
1
I-II
Combined with cetuximab
NCT01468896
IL-12-expressing TILs
Metastatic melanoma
1
I-II
Combined with cyclophosphamide, fludarabine and G-CSF
NCT01236573
Microsphere delivery
HNC
1
n.a.
Combined with GM-CSF and IL-2
NCT00899821
NHS-IL12
Epithelial and mesenchimal malignancies
1
I
As single agent
NCT01417546
pIL-12
Melanoma
1
II
As single agent
NCT01502293
Merkel cell cancer
1
II
NCT01440816
IL-13
IL13-PE38QQR
Brain tumors
1
I
As single agent
NCT00880061
IL-15
IL-15-expressing DCs
Melanoma
1
I-II
Autologous DC-based vaccine
NCT01189383
rhIL-15
Metastatic cancer
1
I
As single agent
NCT01572493
Metastatic melanoma
Metastatic RCC
1
I
NCT01021059
IL-15-activated NK cells
Pediatric refractory solid tumors
1
I-II
Combined with haploidentical stem cell transplantation
NCT01337544
IL-21 rhIL-21 Melanoma
2
I
Combined with ipilimumab
NCT01489059
II
As single agent NCT01152788
Metastatic melanoma
1
II
NCT00601861
RCC 1 I-II NCT00617253

Abbreviations: AML, acute myeloid leukemia; B-CLL, B-cell chronic lymphocytic leukemia; CD40L, CD40 ligand; DC, dendritic cell; FOLFIRI, folinic acid, 5-fluorouracil, irinotecan; FOLFOX, folinic acid, 5-fluorouracil, oxaliplatin; G-CSF, granulocyte colony-stimulating factor; GM-CSF, granulocyte macrophage colony-stimulating factor; HNC, head and neck cancer; IL-, interleukin; MDS, myelodysplastic syndrome; RCC, renal cell carcinoma; NB, neuroblastoma; NHL, non-Hodgkin's lymphoma; p, plasmid-encoded; NK, natural killer; rh, human recombinant; TIL, tumor infiltrating lymphocyte. *started after January, 1st 2008 and not terminated at the day of submission.

G-CSF and GM-CSF stimulate the differentiation of bone marrow stem cells toward the granulocytic (both G-CSF and GM-CSF) and monocytic (GM-CSF only) lineages.20 In addition, G-CSF promotes the mobilization of hematopoietic precursor cells from the bone marrow into the bloodstream.24 Since their approval by FDA (in the early 1990s), G-CSF and GM-CSF have been used in millions of cancer patients to prevent/counterbalance chemotherapy-associated neutropenia.20 Moreover, G-CSF has been tested for its ability to “prime” leukemic cells and hence render them more sensitive to conventional chemotherapy. However, results from a large randomized clinical trial indicate that G-CSF priming might extend disease-free (but not overall) survival, and only in patients with standard-risk acute myeloid leukemia.25 Today, G-CSF and GM-CSF are employed in no less than 100 clinical trials, often - if not always - in combination with therapeutic strategies that are known (or expected) to provoke leucopenia. Multiple studies are testing G-CSF in combination with plerixafor (a derivative of bicyclam with immunostimulatory functions)26 for mobilizing hematopoietic cell stem cells prior to autologous transplantation. Moreover, GM-CSF is extensively being used in the context of anticancer vaccine-based clinical trials, either as a recombinant product (Sargramostim, Molgramostim) or upon genetic engineering of (re-)infused cells. Nevertheless, no clinical studies are currently testing whether G-CSF and GM-CSF might exert anticancer effects independent of their capacity to stimulate early hematopoiesis (www.clinicaltrials.gov).

IL-4 stimulates the differentiation of naive helper T cells into Th2 cells, in turn producing additional IL-4 as well as IL-10 (see below) and IL-13.27 As it drives the local secretion of mitogenic cytokines, the Th2 response has been suggested to promote tumor growth.28 In line with this notion, no trials are currently investigating the efficacy of IL-4 and IL-13 as anticancer agents. Nevertheless, as tumor cells often overexpress the IL-4 and/or the IL-13 receptor,29 IL-4 and IL-13 are being tested as tumor-targeting partners for the enzymatically active portion of Pseudomonas aeruginosa exotoxin A (Table 3).

Besides participating in the acute phase response at the organismal level,3,4 IL-6 can function as a paracrine regulator of inflammation and immunity.30 Moreover, some neoplasms (e.g., most variants of multiple myeloma) produce high levels of IL-6, and these are strictly required for tumor survival.31 Driven by promising preclinical observations,32 several monoclonal antibodies that specifically block IL-6 have been tested in cancer patients during the last decade.31 Nevertheless, the actual efficacy of these drugs for oncological indications remain unclear.33,34 One notable exception is represented by the use of siltuximab (CNTO 328) in ovarian cancer patients experiencing paraneoplastic thrombocytosis.35 The efficacy of siltuximab for oncological indications is currently being tested in no less than 15 phase I/II clinical trials (www.clinicaltrials.gov).

IL-7 is a potent hematopoietic growth factor, stimulating the differentiation of hematopoietic stem cells into lymphoid progenitors as well as the proliferation of mature cells of the lymphoid lineage.36 In spite of early concerns on the possibility that IL-7 might per se sustain hematopoietic tumorigenesis,37 IL-7 is nowadays being investigated in the clinic, in both non-oncological (e.g., in HIV patients in combination with conventional antiretroviral drug) and oncological settings (Table 3). In particular, IL-7 is being tested either as a single agent in metastatic breast cancer patients (mainly to contain lymphopenia and divpenia, i.e., a severe restriction in the TCR repertoire, but also to stimulate an anticancer immune response)38 or in combination with vaccines against RCC, various types of sarcoma or neuroblastoma (www.clinicaltrials.gov).

Besides being involved in the differentiation of naive T cells, IL-12 potently stimulates the functions of T and natural killer (NK) cells. In particular, it promotes the secretion of TNFα and IFNγ and inhibits the generation of IL-4, de facto stimulating a cytotoxic Th1 response.39 In line with this notion, multiple distinct approaches for employing IL-12 in anticancer therapy are being evaluated (Table 3). Thus, in 4 phase I/II clinical trials, rhIL-12 is used in combination with cetuximab (a monoclonal antibody targeting the epidermal growth factor receptor, EGFR) or other cytokines (IL-2 + GM-CSF, delivered with microspheres) in head and neck cancer patients, as well as to boost vaccine-driven anticancer responses in melanoma or breast cancer patients. As an alternative, the IL-12-encoding gene is administered either via plasmid electroporation (to melanoma and Merkel cell carcinoma patients), either with a replication-incompetent adenoviral strain (to melanoma and metastatic breast cancer patients), either within a non viral vector, alone or combined with chemotherapeutics (to colorectal, ovarian and primary peritoneal carcinoma patients), or in the context of adoptive cell transfer immunotherapy with genetically-engineered tumor-infiltrating lymphocytes (www.clinicaltrials.gov).

Similar to IL-12, both IL-15 and IL-21 regulate the proliferation and activation of NK and T cells.40,41 Moreover, IL-15 has been shown to promote the survival of a specific subset of memory CD8+ T cells.42 Besides a few studies in which rhIL-15 is being tested for its ability to stimulate the immune system in HIV+ patients, there are 8 phase I/II clinical trials in which the safety and efficacy of IL-15 and IL-21 is tested in oncological indications (Table 3). In 6 studies, rhIL-15 and rhIL-21 are given, either as standalone medications or combined with targeted anticancer agents, to patients with immunosensitive neoplasms like melanoma and RCC. Alternatively, IL-15 is administered to boost the immune function of freshly infused NK cells, in an adoptive cell transfer approach, or used to engineer DCs for the generation of an efficient vaccine against melanoma (www.clinicaltrials.gov).

Interleukin-1 family members

The IL-1 family of cytokines includes three members: IL-1α, IL-1β and IL-18. IL-1-like proteins are mainly produced by activated macrophages, and play a prominent role in local and systemic inflammation.43 IL-1 family members are synthesized as inactive precursors and are cleaved by intracellular proteases (pro-IL-1α mainly by calpains, pro-IL-1β and pro-IL-18 mainly by the caspase-1-containing complex called inflammasome) before secretion.44-46 So far, two distinct types of monomeric IL-1 receptors and one heteromeric IL-18 receptor have been identified, all belonging to the immunoglobulin superfamily.47,48 Of note, while the type I IL-1 receptor (IL1R1) is biologically active and transduces IL-1 signals, the type II receptor (IL1R2) often acts as a soluble decoy receptor.49,50 Moreover, IL-1α and IL-1β are functionally antagonized by the IL-1 receptor antagonist (IL-1RA), a soluble factor that binds non-productively to IL-1 receptors.51

Blockade of IL-1β signaling with specific monoclonal antibodies or with a recombinant derivative of IL-1RA (anakinra) compromises the therapeutic effects of anthracyclines and oxaliplatin in immunocompetent, but not immunodeficient, mouse models.52,53 This suggests that IL-1β plays a key role in chemotherapeutic responses by virtue of its immunostimulatory functions. Probably due to such a potent pro-inflammatory profile, correlating with a high risk for severe side effects, the administration of recombinant IL-1 to cancer patients has not yet been tested. Conversely, the inhibition of IL-1 signaling with anakinra is being extensively tested, with encouraging results, in patients affected by type 2 diabetes and many other autoimmune disorders (www.clinicaltrials.gov).54-56

Results from a recent phase II trial evaluating the safety and efficacy of rhIL-18 (SB-485232) in metastatic melanoma patients suggest that SB-485232 is safe but has limited anticancer activity as a single agent.57 Nowadays, SB-485232 is being tested, alone or in combination with doxorubicin, in patients affected by epithelial ovarian cancers (phase I, NCT00659178) (www.clinicaltrials.gov), but the results of these trials have not yet been published. Intriguingly, in spite of its potent pro-inflammatory profile, IL-18 has recently been shown to suppress metastasis surveillance by NK cells.58 In line with this notion, elevated levels of circulating IL-18 have been associated with poor disease outcome in pancreatic and lung cancer patients.59,60 Taken together, these observations suggest that the administration of SB-485232 may be beneficial to patients affected by specific neoplasms (such as melanoma) but detrimental in others.

Interleukin-10 family members

According to current classifications, 9 cytokines belong to the IL-10 family: IL-10, IL-19, IL-20, IL-22, IL-24, IL-26, IL-28A, IL-28B and IL-29.61 With the notable exception of IL-24,62 these proteins operate through heteromeric receptors that share the IL-10 receptor β subunit.63,64 Nevertheless, the biological outcomes of IL-10 family members are extremely heterogeneous, encompassing consistent immunosuppression (such as that triggered by IL-10),65 the induction of an antiviral state (by IL-28A, IL-28B and IL-29, which are also known as type III IFNs),66 as well as (often STAT3-dependent) mitogenic and pro-survival effects (by IL-19, IL-20, IL-22 and IL-26).67 Of note, Il10−/− mice develop a chronic and fatal enterocolitis, which can be stopped by administration of IL-10.68 In line with the potent immunosuppressive functions of IL-10, early clinical trials were launched to study the safety and efficacy of rhIL-10 in patients affected by psoriasis and psoriatic arthritis.69 Following the encouraging results of these studies, rhIL-10 (nowadays produced under the name of prevascar or ilodecakin) is being tested in other clinical settings featuring excessive/chronic inflammation, including wound healing and ulcerative colitis (in the latter scenario, IL-10 is produced in situ by genetically engineered Lactococcus lactis). In addition, a humanized anti-IL-10 antibody (SCH708980) is under investigation as an anti-immunosuppressive agent during visceral Leishmaniasis. Of note, there are no ongoing clinical trials that evaluate the therapeutic potential of cytokines from the IL-10 family (or their inhibitors) in cancer patients (www.clinicaltrials.gov). This said, there is a vast literature that correlate elevated serum levels of IL-10 with poor prognosis in multiple hematological and solid malignancies,70-72 suggesting that - at least in some patients - the inhibition of IL-10 may turn out to be therapeutically beneficial.

Interleukin-17 family members

So far, 6 different IL-17 variants have been described (IL-17A-F).73As a family, IL-17 (which is most often secreted by helper T cells in response to IL-23) exerts potent pro-inflammatory functions, stimulating the release of many other cytokines (such as IL-1β, IL-6, GM-CSF, TGFβ and TNFα), chemokines (e.g., MCP-1), and prostaglandins (e.g., PGE2).74 IL-17 family members operate by binding to a heterodimeric receptor, which can be assembled by two 2 of 5 different subunits (IL17RA-E).75 The exact role of IL-17 and IL-17-producing helper T (Th17) cells in tumor progression and response to therapy is controversial.76 On one hand, IL-17-secreting γδ T cells that arise in response to IL-1β + IL-23 appear to be required for optimal therapeutic responses, presumably due to the fact that they are prone to secrete GM-CSF and IFNγ.52,53 On the other hand, Th17 cells that differentiate in the presence of TGFβ and IL-6 produce IL-17 and IL-10, de facto exerting immunosuppressive effects.77 Hence, are no ongoing clinical trials that evaluate the therapeutic potential of this cytokine in oncological settings (www.clinicaltrials.gov). Conversely, monoclonal antibodies targeting IL-17A (ixekizumab and secukinumab) or IL17RA (brodalumab) have given encouraging results for the treatment of diseases with an auto-immune component such as uveitis, rheumatoid arthritis and psoriasis.78-80 Ixekizumab (also known as LY2439821) is currently being tested in a few clinical trials to optimize dosage and administration schedules (www.clinicaltrials.gov).

Interferons

Interferons (IFNs) have first been characterized for their ability to “interfere” with viral replication, and indeed one of their major functions is the establishment of a robust antiviral state in response to infection.81 In addition, IFNs activate immune cells and facilitate the recognition of tumor cells by the immune system as they stimulate antigen presentation to T lymphocytes.82 IFNs are usually subdivided into 3 classes: type I (including IFNα, IFNβ and IFNω), type II (in humans: IFNγ), and type III IFNs (including IL-28A, IL-28B and IL-29, see above). Type I IFNs signal through a receptor complex known as the IFNα receptor (IFNAR), consisting of IFNAR1 and IFNAR2 chains. Along similar lines, IFNγ functions by binding to a heterodimeric IFNγ receptor (IFNGR), which consists of IFNGR1 and IFNGR2 chains.83 Recent data indicate that both IFNAR and IFNGR are required for optimal responses to immunogenic chemotherapy, at least in preclinical settings.84-86

During the last two decades, great efforts have been spent at optimizing strategies to modulate the IFN system. On one hand, humanized monoclonal antibodies specific for IFNγ (fontolizumab) have been developed and tested in clinical settings. Thus, fontolizumab has been shown to be safe and clinically efficient in patients with chronic inflammatory pathologies such as rheumatoid arthritis or moderate to severe Crohn’s disease.87,88 On the other hand, recombinant IFNα2b has been approved by FDA for a wide range of indications, including chronic hepatitis B and C, hairy cell leukemia, chronic myeloid leukemia, multiple myeloma, follicular lymphoma, and malignant melanoma (Table 2). Nowadays, the efficacy of IFNα2 against cancer is being tested in 58 clinical trials, of which 22 evaluate IFNα2b in FDA-approved clinical settings (most often in melanoma patients). Of the remaining 36 studies, 12 investigate the therapeutic potential of IFNα2 in RCC patients, while the others span multiple solid and hematopoietic neoplasms, including bladder cancer, hepatocellular carcinoma, colorectal carcinoma, mesothelioma, pancreatic cancer, and Hodgkin’s lymphoma (Table 4). Frequently (11/58 trials), IFNα2 is tested as monotherapy. As an alternative, combination regimens with bevacizumab (7/58 trials), dacarbazine (3/58 trials), cisplatin (3/58 trials), cyclophosphamide (2/58 trials), sorafenib (2/58 trials) or decitabine (2/58 trials) are being investigated. Of note, a relatively high proportion of IFNα2-based clinical trials (around 20%) consists of advanced studies (phase III/IV), presumably reflecting the limited safety issued associated with the use of a FDA-approved molecule. At present, rhIFNγ (which is approved by FDA for the therapy of granulomatous disease and severe osteopetrosis) is being tested against a wide range of cancer-unrelated pathologies, including infections and inflammatory diseases. Moreover, there are 5 ongoing clinical studies that evaluate the safety and efficacy of IFNγ in oncological indications. In one case, IFNγ is used alone or in combination with bevacizumab, 5-fluorouracil and folinic acid against colorectal cancer. In the others, IFNγ is employed to boost anticancer immune responses driven by the administration of autologous T cells or peptidic vaccines (in soft tissue sarcoma, melanoma or plasma cell neoplasm patients) (www.clinicaltrials.gov).

Table 4. Clinical trials* on IFNs as anticancer agents (main trends).

Chemokine Tumor type Trials* Phase Notes Ref.
Early clinical trials (phase I-II)
IFNα
Adult T-cell leukemia/lymphoma
1
n.a.
Combined with valproic acid
and zidovudine
NCT00854581
Advanced solid tumors
1
I
Combined with decitabine
NCT00701298
1
I
Combined with cyclophosphamide
and vinorelbine
NCT00908869
1
I
Combined with sodium stibogluconate
NCT00629200
1
II
NCT01479309
Bladder cancer
1
II
As single agent
NCT01162785
Cervical cancer
1
II
Combined with radiotherapy and RA
NCT01276730
CML
3
I
Combined with nilotinib
NCT01294618
II
As single agent
NCT01392170
Combined with nilotinib
NCT01220648
CRC
1
I-II
Combined with celecoxib and poly-ICLC
NCT01545141
Cutaneous T-cell lymphoma
1
n.a.
Combined with UV light
NCT00724061
HCC
1
I
As single agent
NCT00838968
Hodgkin lymphoma
1
II
Combined with ABVD
NCT01404936
Kidney cancer
1
I-II
Combined with radioablation
NCT00891475
1
I-II
Combined with pazaponib
NCT01513187
1
II
Combined with sorafenib
NCT00589550
4
II
Combined with bevacizumab
NCT00619268 NCT00719264 NCT00796757 NCT00873236
1
II
Combined with celecoxib
NCT01158534
1
II
Combined with GM-CSF and IL-2
NCT01176552
1
II
Combined with bevacizumab and IL-2
NCT01274273
Malignant pleural mesothelioma
2
I
Combined with cisplatin and pemetrexed
NCT01119664
As single agent
NCT01212367
NHL
1
II
In alkylating agent- or anthracycline-based
regimens plus rituximab
NCT00842114
Neurofibroma
1
II
As single agent
NCT00678951
Pancreatic cancer
2
I
Combined with 5-FU, cisplatin and gemcitabine
NCT00660270
Combined with 5-FU, docetaxel, gemcitabine
and oxaliplatin
NCT00761241
Various solid tumors
1
II
Combined with DC- and TIL-based immunotherapy plus cyclophosphamide and TNFα
NCT00610389
IFNγ
CRC
1
II
Combined with bevacizumab, 5-FU and folinic acid
NCT00786643
Melanoma
1
n.a.
Combined with peptide-based vaccine
NCT00977145
1
I-II
Combined with TIL and IL-2
NCT01082887
Plasma cell neoplasms
1
II
Combined with antitumor vaccine
NCT00616720
Soft tissue sarcoma
1
II
Combined with autologous NY-ESO-1-specific CD8+ T cells, cyclophosphamide and IL-2
NCT01477021
Advanced clinical trials (phase III-IV)
IFNα Bladder cancer
1
III
Combined with BCG, epirubicin and mitomycin C
NCT01094964
CRC
1
III
Combined with 5-FU and folinic acid
NCT01060501
HCV-associated HCC
1
IV
Combined with ribavirin
NCT00834860
Kidney cancer
1
III
Combined with bevacizumab NCT00738530
Urogenital cancer 1 III NCT00631371

Abbreviations: 5-FU, 5-fluorouracil; ABVD, bleomycin, dacarbazine, doxorubicin, vinblastine; BCG, bacillus Calmette-Guérin; CML, chronic myeloid leukemia; CRC, colorectal carcinoma; DC, dendritic cell; GM-CSF, granulocyte macrophage colony stimulating factor; HCC, hepatocellular carcinoma; HCV, hepatitis C virus; IFN, interferon; IL, interleukin; n.a., not available; NHL, Non-Hodgkin's lymphoma; RA, retinoic acid; RCC, renal cell carcinoma; TIL, tumor infiltrating lymphocyte; TNF, tumor necrosis factor; UV, UV. *started after January, 1st 2008 and not terminated at the day of submission.

PDGF family members

The PDGF family of cytokines includes 4 PDGF isoforms (PDGFA-D), 4 isoforms of the vascular endothelial growth factor (VEGFA-D) and other 6 rather heterogeneous proteins. PDGFs, which activate cellular responses via 2 distinct tyrosine kinase receptors (PDGFRα and β), can be released by activated platelets, as well as smooth muscle cells, activated macrophages, and endothelial cells.89 PDGFs exert potent mitogenic functions over a plethora of distinct cell types and play an important role in development and angiogenesis.90 In line with this notion, PDGFRs are often hyperactivated in cancer, often due to point mutations in PDGFRA or PDGFRB, such as in the case of gastrointestinal stromal tumors (GISTs).91 The same holds true for other PDGF-like protein receptors including KIT (which delivers potent mitogenic signals and is mutated in more than 70% GIST patients),92 EGFR (which is overexpressed or mutated in colorectal and lung cancer)93 and FLT3 (which is frequently hyperactivated in acute myeloid leukemia patients).94 Along similar lines, tumor cells often overexpress VEGF, stimulating the so-called “angiogenic switch,” i.e., the transition of a tumor mass from non-vascularized to vascularized.95 The FLT3 ligand (FLT3L) has been shown to expand DCs in vivo (in mice), resulting in the elicitation of NK-mediated antitumor effects.96,97 Nevertheless, given their potent mitogenic (and hence potentially tumorigenic) properties, it is not surprising that PDGF family members are not employed as immunostimulatory agents against cancer, neither alone nor in combination with other drugs. Rather, chemicals that selectively inhibit the tyrosine kinase activity of PDGFR, KIT and FLT3 (including imatinib, lapatinib and sorafenib), as well as monoclonal antibodies that specifically block EGFR (i.e., cetuximab, panitumumab) or VEGFA (i.e., bevacizumab), are currently approved by FDA for an ever-increasing number of oncological indications and are associated with consistent rates of remissions.98 These agents are the subject of an intense wave of clinical trials, aimed at finding new indications as well as at optimizing dosage and administration schedules (www.clinicaltrials.gov). The detailed discussion of such studies largely exceeds the scope of the present Trial Watch and can be found elsewhere.99 This said, it is worth noting that several compounds that were developed to specifically target the receptors for PDGF family members, including imatinib and many others, are less specific than expected, and de facto exert potent off-target immunostimulatory effects.84,100,101 This property has raised great expectations and — as many of these agents have already been approved by FDA for anticancer therapy — will presumably be tested in the short future in appropriate clinical trials.

TGFβ family members

TGFβ1 has first been characterized for its ability to trigger the growth of normal rat kidney (NRK) fibroblast colonies in soft-agar assays (though only if small amounts of EGF were present).102 Since then, 3 distinct subtypes of human TGFβ have been identified (TGFβ1–3), together with at least additional 8 proteins that are now considered part of the TGFβ family, including polypeptides of the activin-inhibin hormonal system and at least 2 bone morphogenic proteins (BMP2 and BMP7).103 TGFβs signals are transduced intracellularly by homo- or heterodimeric serine/threonine kinase receptors, which can be assembled starting from 3 distinct subunits (TGFBR1–3).104 In normal cells, TGFβs, and in particular TGFβ1, exert antiproliferative effects by upregulating the expression of cell cycle inhibitors such as p21CIP1 and p15INK4B or trigger cell death by activating SMAD- or DAXX-dependent apoptosis.105 Conversely, cancer cells often become refractory to the inhibitory effects of TGFβ and - at least in some cases - also overexpress it.106 In this setting, TGFβ stimulates angiogenesis and facilitates the conversion of effector T lymphocytes into FOXP3+ immunosuppressive T cells (Tregs) or Th17 cells, de facto exerting pro-tumor functions.106,107 Of note, BMP2 and BMP7 also activate the SMAD signaling pathway, yet mainly function as osteogenic mediators.108 Over the two last decades, two distinct monoclonal antibodies against TGFβ have been developed: metelimumab (CAT-192, which specifically targets TGFβ1) and fresolimumab (GC-1008, which can block TGFβ1–3).109 In people affected by systemic sclerosis, metelimumab was found to be safe but ineffective,110 leading the proprietary pharmaceutical company to focus on fresolimumab. Today, fresolimumab is being tested, alone or combined with other interventions, in pulmonary fibrosis, myelofibrosis and focal segmental glomerulosclerosis patients, as well as in patients affected by glioma, mesothelioma, melanoma and metastastic breast cancer (www.clinicaltrials.gov). To our knowledge, any other clinical trial is currently investigating the modulation of TGFβ signaling in oncological indications.

Tumor necrosis factors

The TNF family of cytokines includes more than 15 distinct proteins, all of which can - at least in some settings - trigger cell death upon binding to specific transmembrane receptors.111-113 The best known member of the family, TNFα, has first been identified in 1975, thanks to the work of Lloyd Old at the Memorial Sloan-Kettering Cancer Center (New York).114 Another TNF, lymphotoxin (LT, also known as TNFβ), had been discovered a few years earlier, in 1968,115,116 but it was not until the cloning of TNFα and LT cDNAs, in 1985, that the homology between these 2 proteins became evident.117 TNFα can bind 2 distinct transmembrane receptors that are differentially expressed: while type I TNFR (TNFR1) is found at the surface of virtually all cells, type II TNFR (TNFR2) is expressed only by cells of the immune system.113 TNFR1 ligation can induce biological outcomes as diverse as the activation of the NF-κB system (which normally delivers pro-survival signals), the initiation of mitogen-activated protein kinase (MAPK)-transduced signaling cascades (which can regulate proliferation, differentiation and cell death), and the induction of cell death (via apoptosis or regulated necrosis).118,119 The TNF family also includes FAS and CD40 ligands (FASL and CD40L, both exerting a major role in the development, homeostasis and function of the immune system),120,121 RANK ligand (RANKL, which has been involved in the differentiation of osteoclasts, in the maturation of DCs and in hormone-driven breast carcinogenesis),122-124 as well as the TNF-related apoptosis-inducing ligand (TRAIL).125

TNFα has been shown to play a crucial role in the pathogenesis of multiple inflammatory diseases, including (though presumably not limited to) rheumatoid arthritis, Crohn's disease, psoriasis, psoriatic arthritis, ankylosing spondylitis and juvenile idiopathic arthritis.126 During the past 20 y, a large battery of TNF inhibitors has been developed and approved by FDA for the treatment of these conditions. Such agents include anti-TNFα monoclonal antibodies like infliximab (Remicade), adalimumab (Humira), certolizumab pegol (Cimzia), and golimumab (Simponi), as well as the artificial fusion protein etanercept (Enbrel), consisting in two soluble human TNFR moieties linked to the Fc portion of an IgG1.99,126 Denosumab (a fully human anti-RANKL antibody), is currently approved by FDA for the treatment of post-menopausal osteoporosis and for the prevention of skeletal-related events in patients with bone metastases from solid tumors (Table 2).127,128 Furthermore, a monoclonal antibody that neutralizes the OX40 ligand (OX40L, an important co-stimulatory molecule during immune responses) is currently being tested for the prevention of allergen-induced airway obstruction in adults with mild allergic asthma (www.clinicaltrials.gov).

In addition to these applications and studies (which do not directly relate to anticancer therapy), TNFα has been extensively employed in combination with melphalan (an alkylating agent) for isolated limb perfusion in metastatic melanoma and sarcoma patients,129 an approach associated with high rates of limb salvage but limited impact on overall survival.130,131 Moreover, there are 22 ongoing clinical trials that evaluate whether the modulation of TNF family members is beneficial or not to cancer patients (www.clinicaltrials.gov). Three of these 22 studies (all of which are in phase I/II) investigate whether the administration of rhTNFα (as such or fused to a tumor-targeting antibody fragment)132 is efficient (alone or in combination with doxorubicin or melphalan) in patients affected by lymphoma, melanoma or other advanced solid tumors (Table 5). Four clinical trials are testing whether monoclonal antibodies that can activate the TRAIL receptor 1 (TRAILR1), de facto mimicking TRAIL cytotoxicity, exert therapeutic effects, if combined with other chemotherapeutics, against hematological and solid tumors including cervical and colorectal carcinoma (Table 5). The safety and efficacy of strategies based on CD40L are being investigated in 7 clinical trials (Table 5). In this setting, CD40L is either introduced directly into the tumor by local injection of adenoviral vectors, or transfected into autologous DCs or cancer cells that are used to generate vaccines. In addition, CP-870,893 (a monoclonal antibody that activated CD40, de facto mimicking CD40L) is tested, in combination with carboplatin and paclitaxel, in patients bearing advanced solid tumors. Finally, denosumab is currently under investigation (in 9 phase II/III clinical trials) for its capacity to prevent/treat bone metastasis in patients affected by multiple myeloma and solid tumors including breast and prostate cancer (www.clinicaltrials.gov).

Table 5. Clinical trials* on TNFs and TNF-mimicking agents in anticancer therapy (main trends).

Chemokine Agent Tumor type Trials* Phase Notes Ref.
CD40L
CP-870,893
Metastatic solid tumors
1
I
Combined with carboplatin and paclitaxel
NCT00607048
rhCD40L
B-CLL
1
I
CD40L-expressing autologous tumor cell-based vaccine
NCT00609076
Bladder cancer
1
I-II
Adenoviral gene therapy
NCT00891748
Lung cancer
2
II
CD40L-expressing allogeneic tumor cell-based vaccine
NCT00601796
CD40L-expressing
cell-based vaccines
NCT01433172
MDS
1
I
CD40L
NCT00840931
Melanoma
1
I
Adenoviral gene therapy
NCT01455259
TNFα
L19TNFα
Advanced solid tumors
1
I-II
As single agent
NCT01253837
Melanoma
1
I
Combined with melphalan
NCT01213732
rhTNFα
Advanced solid tumors
Lymphoma
1
I
Combined with doxorubicin
NCT01490047
TRAIL Conatumumab
Advanced hematological and solid tumors
2
II
Alone or combined with bevacizumab, FOLFOX and ganitumab
NCT01327612
Advanced solid tumors
I-II
Combined with AMG479
NCT00819169
Dulanermin
CRC
1
I
Combined with bevacizumab
NCT00873756
Mapatumumab Advanced cervical cancer 1 I-II Combined with cisplatin and
radiotherapy
NCT01088347

Abbreviations: Ad, adenovirus; B-CLL, B- chronic lymphocytic leukemia; CRC, colorectal cancer; CD40L, CD40 ligand; DC, dendritic cell; FOLFOX, folinic acid, fluorouracil, oxaliplatin; MDS, myelodysplastic syndrome; MM, multiple myeloma, rh, recombinant human; TNF, tumor necrosis factor; TRAIL, TNF-related apoptosis-inducing ligand. *started after January, 1st 2008 and not terminated at the day of submission.

Concluding remarks

Cytokines constitute a highly heterogeneous group of proteins that - taken together - regulate virtually all aspects of the cell biology. Some of them, such as GM-CSF, potently stimulate the replication and survival of hematopoietic cell precursors, and are mainly employed in tumor patients to help them reconstituting the hematopoietic system upon chemotherapy or transplantation. Others, including multiple members of the PDGF family, exert potent mitogenic functions and hence cannot be employed in cancer therapy, as they would stimulate, rather than prevent, oncogenesis and tumor progression. Finally, there are cytokines that have been approved by FDA (or are being clinically tested) for oncological indications, owing to their ability to stimulate anti-cancer immune responses. For instance, IL-2 has been associated with consistent rates of tumor regression in melanoma and RCC patients,133,134 probably because these malignancies are able to elicit per se elevated levels of antitumor lymphocytes. Although the elevated pleiotropism of the cytokine system constitutes an obstacle for the development of highly targeted therapeutics (implying that these proteins may be intrinsically prone to elicit side effects), we surmise and hope that ongoing and/or future clinical trials will lead to the approval of additional cytokines for use in humans against cancer.

Acknowledgments

Authors are supported by the Ligue contre le Cancer (équipes labelisées), AXA Chair for Longevity Research, Cancéropôle Ile-de-France, Institut National du Cancer (INCa), Fondation Bettencourt-Schueller, Fondation de France, Fondation pour la Recherche Médicale, Agence National de la Recherche, the European Commission (Apo-Sys, ArtForce, ChemoRes. Death-Train) and the LabEx Immuno-Oncology.

Glossary

Abbreviations:

BMP

bone morphogenic protein

CD40L

CD40 ligand

DC

dendritic cell

EGF

epidermal growth factor

EGFR

EGF receptor

EPO

erythropoietin

FASL

FAS ligand

FLT3L, FLT3 ligand

GIST, gastrointestinal stromal tumors

G-CSF

granulocyte colony stimulating factor

GM-CSF

granulocyte macrophage colony stimulating factor

IFN

interferon

IFNAR

IFNα receptor

IFNGR

IFNγ receptor

IL

interleukin

IL-1RA

interleukin-1 receptor antagonist

IL1R1

type I IL-1 receptor

IL1R2

type II interleukin-2 receptor

INHβ

inhibin β

LPS

lipopolysaccharide

LT

lymphotoxin

MAPK

mitogen-activated protein kinase

NK

natural killer

NRK

normal rat kidney

OX40L

OX40 ligand

PDGF

platelet-derived growth factor

PDGFR

PDGF receptor

RANKL

RANK ligand

rh

recombinant human

RCC

renal cell carcinoma

TGFβ

transforming growth factor β

Th17

IL-17-producing helper T

TNF

tumor necrosis factor

TNFR

TNF receptor

TPO

thrombopoietin

Tregs

FOXP3+ immunosuppressive T cells

TRAIL

TNF-related apoptosis-inducing ligand

TRAILR1

TRAIL receptor 1

VEGF

vascular endothelial growth factor

Footnotes

Contributed equally to this article.

References

  • 1.Borish LC, Steinke JW. 2. Cytokines and chemokines. J Allergy Clin Immunol. 2003;111(Suppl):S460–75. doi: 10.1067/mai.2003.108. [DOI] [PubMed] [Google Scholar]
  • 2.Steinke JW, Borish L. 3. Cytokines and chemokines. J Allergy Clin Immunol. 2006;117(Suppl Mini-Primer):S441–5. doi: 10.1016/j.jaci.2005.07.001. [DOI] [PubMed] [Google Scholar]
  • 3.Ohlsson K, Björk P, Bergenfeldt M, Hageman R, Thompson RC. Interleukin-1 receptor antagonist reduces mortality from endotoxin shock. Nature. 1990;348:550–2. doi: 10.1038/348550a0. [DOI] [PubMed] [Google Scholar]
  • 4.Tracey KJ, Fong Y, Hesse DG, Manogue KR, Lee AT, Kuo GC, et al. Anti-cachectin/TNF monoclonal antibodies prevent septic shock during lethal bacteraemia. Nature. 1987;330:662–4. doi: 10.1038/330662a0. [DOI] [PubMed] [Google Scholar]
  • 5.Andrews DM, Chow MT, Ma Y, Cotterell CL, Watt SV, Anthony DA, et al. Homeostatic defects in interleukin 18-deficient mice contribute to protection against the lethal effects of endotoxin. Immunol Cell Biol. 2011;89:739–46. doi: 10.1038/icb.2010.168. [DOI] [PubMed] [Google Scholar]
  • 6.Dranoff G. Cytokines in cancer pathogenesis and cancer therapy. Nat Rev Cancer. 2004;4:11–22. doi: 10.1038/nrc1252. [DOI] [PubMed] [Google Scholar]
  • 7.Coussens LM, Werb Z. Inflammation and cancer. Nature. 2002;420:860–7. doi: 10.1038/nature01322. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Comerford I, McColl SR. Mini-review series: focus on chemokines. Immunol Cell Biol. 2011;89:183–4. doi: 10.1038/icb.2010.164. [DOI] [PubMed] [Google Scholar]
  • 9.Szekanecz Z, Vegvari A, Szabo Z, Koch AE. Chemokines and chemokine receptors in arthritis. Front Biosci (Schol Ed) 2010;2:153–67. doi: 10.2741/s53. [Schol Ed] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Lazaar AL, Sweeney LE, MacDonald AJ, Alexis NE, Chen C, Tal-Singer R. SB-656933, a novel CXCR2 selective antagonist, inhibits ex vivo neutrophil activation and ozone-induced airway inflammation in humans. Br J Clin Pharmacol. 2011;72:282–93. doi: 10.1111/j.1365-2125.2011.03968.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Yellin M, Paliienko I, Balanescu A, Ter-Vartanian S, Tseluyko V, Xu LA, et al. A phase II, randomized, double-blind, placebo-controlled study evaluating the efficacy and safety of MDX-1100, a fully human anti-CXCL10 monoclonal antibody, in combination with methotrexate in patients with rheumatoid arthritis. Arthritis Rheum. 2011 doi: 10.1002/art.34330. [DOI] [PubMed] [Google Scholar]
  • 12.Vergunst CE, Gerlag DM, Lopatinskaya L, Klareskog L, Smith MD, van den Bosch F, et al. Modulation of CCR2 in rheumatoid arthritis: a double-blind, randomized, placebo-controlled clinical trial. Arthritis Rheum. 2008;58:1931–9. doi: 10.1002/art.23591. [DOI] [PubMed] [Google Scholar]
  • 13.Hoffmann S, Hoos J, Klussmann S, Vonhoff S. RNA aptamers and spiegelmers: synthesis, purification, and post-synthetic PEG conjugation. Curr Protoc Nucleic Acid Chem 2011; Chapter 4:Unit 4 46 1-30. [DOI] [PubMed] [Google Scholar]
  • 14.Stephenson J. Researchers buoyed by novel HIV drugs: will expand drug arsenal against resistant virus. JAMA. 2007;297:1535–6. doi: 10.1001/jama.297.14.1535. [DOI] [PubMed] [Google Scholar]
  • 15.Hedrick JA, Zlotnik A. Identification and characterization of a novel beta chemokine containing six conserved cysteines. J Immunol. 1997;159:1589–93. [PubMed] [Google Scholar]
  • 16.Juarez J, Bendall L. SDF-1 and CXCR4 in normal and malignant hematopoiesis. Histol Histopathol. 2004;19:299–309. doi: 10.14670/HH-19.299. [DOI] [PubMed] [Google Scholar]
  • 17.Lyman GH, Dale DC. Introduction to the hematopoietic growth factors. Cancer Treat Res. 2011;157:3–10. doi: 10.1007/978-1-4419-7073-2_1. [DOI] [PubMed] [Google Scholar]
  • 18.Metcalf D. Hematopoietic cytokines. Blood. 2008;111:485–91. doi: 10.1182/blood-2007-03-079681. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Mughal TI. Current and future use of hematopoietic growth factors in cancer medicine. Hematol Oncol. 2004;22:121–34. doi: 10.1002/hon.736. [DOI] [PubMed] [Google Scholar]
  • 20.Metcalf D. The colony-stimulating factors and cancer. Nat Rev Cancer. 2010;10:425–34. doi: 10.1038/nrc2843. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Galluzzi L, Vacchelli E, Eggermont A, Fridman WH, Galon J, Sautès-Fridman C, et al. Trial Watch - Adoptive cell transfer immunotherapy. Oncoimmunol. 2012;1 doi: 10.4161/onci.19549. In press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Johannsen M, Spitaleri G, Curigliano G, Roigas J, Weikert S, Kempkensteffen C, et al. The tumour-targeting human L19-IL2 immunocytokine: preclinical safety studies, phase I clinical trial in patients with solid tumours and expansion into patients with advanced renal cell carcinoma. Eur J Cancer. 2010;46:2926–35. doi: 10.1016/j.ejca.2010.07.033. [DOI] [PubMed] [Google Scholar]
  • 23.Pedretti M, Verpelli C, Mårlind J, Bertani G, Sala C, Neri D, et al. Combination of temozolomide with immunocytokine F16-IL2 for the treatment of glioblastoma. Br J Cancer. 2010;103:827–36. doi: 10.1038/sj.bjc.6605832. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Thomas J, Liu F, Link DC. Mechanisms of mobilization of hematopoietic progenitors with granulocyte colony-stimulating factor. Curr Opin Hematol. 2002;9:183–9. doi: 10.1097/00062752-200205000-00002. [DOI] [PubMed] [Google Scholar]
  • 25.Schiffer CA. Hematopoietic growth factors and the future of therapeutic research on acute myeloid leukemia. N Engl J Med. 2003;349:727–9. doi: 10.1056/NEJMp030076. [DOI] [PubMed] [Google Scholar]
  • 26.DiPersio JF, Uy GL, Yasothan U, Kirkpatrick P. Plerixafor. Nat Rev Drug Discov. 2009;8:105–6. doi: 10.1038/nrd2819. [DOI] [PubMed] [Google Scholar]
  • 27.Kelly-Welch AE, Hanson EM, Boothby MR, Keegan AD. Interleukin-4 and interleukin-13 signaling connections maps. Science. 2003;300:1527–8. doi: 10.1126/science.1085458. [DOI] [PubMed] [Google Scholar]
  • 28.De Monte L, Reni M, Tassi E, Clavenna D, Papa I, Recalde H, et al. Intratumor T helper type 2 cell infiltrate correlates with cancer-associated fibroblast thymic stromal lymphopoietin production and reduced survival in pancreatic cancer. J Exp Med. 2011;208:469–78. doi: 10.1084/jem.20101876. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Kawakami M, Kawakami K, Stepensky VA, Maki RA, Robin H, Muller W, et al. Interleukin 4 receptor on human lung cancer: a molecular target for cytotoxin therapy. Clin Cancer Res. 2002;8:3503–11. [PubMed] [Google Scholar]
  • 30.Zhong Z, Wen Z, Darnell JE., Jr. Stat3: a STAT family member activated by tyrosine phosphorylation in response to epidermal growth factor and interleukin-6. Science. 1994;264:95–8. doi: 10.1126/science.8140422. [DOI] [PubMed] [Google Scholar]
  • 31.Barton BE. Interleukin-6 and new strategies for the treatment of cancer, hyperproliferative diseases and paraneoplastic syndromes. Expert Opin Ther Targets. 2005;9:737–52. doi: 10.1517/14728222.9.4.737. [DOI] [PubMed] [Google Scholar]
  • 32.Voorhees PM, Chen Q, Kuhn DJ, Small GW, Hunsucker SA, Strader JS, et al. Inhibition of interleukin-6 signaling with CNTO 328 enhances the activity of bortezomib in preclinical models of multiple myeloma. Clin Cancer Res. 2007;13:6469–78. doi: 10.1158/1078-0432.CCR-07-1293. [DOI] [PubMed] [Google Scholar]
  • 33.Fizazi K, De Bono JS, Flechon A, Heidenreich A, Voog E, Davis NB, et al. Randomised phase II study of siltuximab (CNTO 328), an anti-IL-6 monoclonal antibody, in combination with mitoxantrone/prednisone versus mitoxantrone/prednisone alone in metastatic castration-resistant prostate cancer. Eur J Cancer. 2012;48:85–93. doi: 10.1016/j.ejca.2011.10.014. [DOI] [PubMed] [Google Scholar]
  • 34.Hunsucker SA, Magarotto V, Kuhn DJ, Kornblau SM, Wang M, Weber DM, et al. Blockade of interleukin-6 signalling with siltuximab enhances melphalan cytotoxicity in preclinical models of multiple myeloma. Br J Haematol. 2011;152:579–92. doi: 10.1111/j.1365-2141.2010.08533.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Stone RL, Nick AM, McNeish IA, Balkwill F, Han HD, Bottsford-Miller J, et al. Paraneoplastic thrombocytosis in ovarian cancer. N Engl J Med. 2012;366:610–8. doi: 10.1056/NEJMoa1110352. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Mackall CL, Fry TJ, Gress RE. Harnessing the biology of IL-7 for therapeutic application. Nat Rev Immunol. 2011;11:330–42. doi: 10.1038/nri2970. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Or R, Abdul-Hai A, Ben-Yehuda A. Reviewing the potential utility of interleukin-7 as a promoter of thymopoiesis and immune recovery. Cytokines Cell Mol Ther. 1998;4:287–94. [PubMed] [Google Scholar]
  • 38.Manuel M, Tredan O, Bachelot T, Clapisson G, Courtier A, Parmentier G, et al. Lymphopenia combined with low TCR diversity (divpenia) predicts poor overall survival in metastatic breast cancer patients. Oncoimmunol. 2012;1 doi: 10.4161/onci.19545. In press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Trinchieri G. Interleukin-12 and the regulation of innate resistance and adaptive immunity. Nat Rev Immunol. 2003;3:133–46. doi: 10.1038/nri1001. [DOI] [PubMed] [Google Scholar]
  • 40.Waldmann TA. The biology of interleukin-2 and interleukin-15: implications for cancer therapy and vaccine design. Nat Rev Immunol. 2006;6:595–601. doi: 10.1038/nri1901. [DOI] [PubMed] [Google Scholar]
  • 41.Parrish-Novak J, Dillon SR, Nelson A, Hammond A, Sprecher C, Gross JA, et al. Interleukin 21 and its receptor are involved in NK cell expansion and regulation of lymphocyte function. Nature. 2000;408:57–63. doi: 10.1038/35040504. [DOI] [PubMed] [Google Scholar]
  • 42.Judge AD, Zhang X, Fujii H, Surh CD, Sprent J. Interleukin 15 controls both proliferation and survival of a subset of memory-phenotype CD8(+) T cells. J Exp Med. 2002;196:935–46. doi: 10.1084/jem.20020772. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Dunne A, O'Neill LA. The interleukin-1 receptor/Toll-like receptor superfamily: signal transduction during inflammation and host defense. Sci STKE 2003; 2003:re3. [DOI] [PubMed]
  • 44.Watanabe N, Kobayashi Y. Selective release of a processed form of interleukin 1 alpha. Cytokine. 1994;6:597–601. doi: 10.1016/1043-4666(94)90046-9. [DOI] [PubMed] [Google Scholar]
  • 45.Zitvogel L, Kepp O, Galluzzi L, Kroemer G. Inflammasomes in carcinogenesis and anticancer immune responses. Nat Immunol. 2012;13:343–51. doi: 10.1038/ni.2224. [DOI] [PubMed] [Google Scholar]
  • 46.Groß O, Yazdi AS, Thomas CJ, Masin M, Heinz LX, Guarda G, et al. Inflammasome activators induce interleukin-1alpha secretion via distinct pathways with differential requirement for the protease function of caspase-1. Immunity. 2012;36:388–400. doi: 10.1016/j.immuni.2012.01.018. [DOI] [PubMed] [Google Scholar]
  • 47.Liu C, Hart RP, Liu XJ, Clevenger W, Maki RA, De Souza EB. Cloning and characterization of an alternatively processed human type II interleukin-1 receptor mRNA. J Biol Chem. 1996;271:20965–72. doi: 10.1074/jbc.271.34.20965. [DOI] [PubMed] [Google Scholar]
  • 48.Sims JE, March CJ, Cosman D, Widmer MB, MacDonald HR, McMahan CJ, et al. cDNA expression cloning of the IL-1 receptor, a member of the immunoglobulin superfamily. Science. 1988;241:585–9. doi: 10.1126/science.2969618. [DOI] [PubMed] [Google Scholar]
  • 49.Dower SK, Kronheim SR, Hopp TP, Cantrell M, Deeley M, Gillis S, et al. The cell surface receptors for interleukin-1 alpha and interleukin-1 beta are identical. Nature. 1986;324:266–8. doi: 10.1038/324266a0. [DOI] [PubMed] [Google Scholar]
  • 50.Symons JA, Young PR, Duff GW. Soluble type II interleukin 1 (IL-1) receptor binds and blocks processing of IL-1 beta precursor and loses affinity for IL-1 receptor antagonist. Proc Natl Acad Sci U S A. 1995;92:1714–8. doi: 10.1073/pnas.92.5.1714. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Arend WP. Interleukin 1 receptor antagonist. A new member of the interleukin 1 family. J Clin Invest. 1991;88:1445–51. doi: 10.1172/JCI115453. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Ma Y, Aymeric L, Locher C, Mattarollo SR, Delahaye NF, Pereira P, et al. Contribution of IL-17-producing gamma delta T cells to the efficacy of anticancer chemotherapy. J Exp Med. 2011;208:491–503. doi: 10.1084/jem.20100269. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Mattarollo SR, Loi S, Duret H, Ma Y, Zitvogel L, Smyth MJ. Pivotal role of innate and adaptive immunity in anthracycline chemotherapy of established tumors. Cancer Res. 2011;71:4809–20. doi: 10.1158/0008-5472.CAN-11-0753. [DOI] [PubMed] [Google Scholar]
  • 54.Aksentijevich I, Masters SL, Ferguson PJ, Dancey P, Frenkel J, van Royen-Kerkhoff A, et al. An autoinflammatory disease with deficiency of the interleukin-1-receptor antagonist. N Engl J Med. 2009;360:2426–37. doi: 10.1056/NEJMoa0807865. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Goldbach-Mansky R, Dailey NJ, Canna SW, Gelabert A, Jones J, Rubin BI, et al. Neonatal-onset multisystem inflammatory disease responsive to interleukin-1beta inhibition. N Engl J Med. 2006;355:581–92. doi: 10.1056/NEJMoa055137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Larsen CM, Faulenbach M, Vaag A, Vølund A, Ehses JA, Seifert B, et al. Interleukin-1-receptor antagonist in type 2 diabetes mellitus. N Engl J Med. 2007;356:1517–26. doi: 10.1056/NEJMoa065213. [DOI] [PubMed] [Google Scholar]
  • 57.Tarhini AA, Millward M, Mainwaring P, Kefford R, Logan T, Pavlick A, et al. A phase 2, randomized study of SB-485232, rhIL-18, in patients with previously untreated metastatic melanoma. Cancer. 2009;115:859–68. doi: 10.1002/cncr.24100. [DOI] [PubMed] [Google Scholar]
  • 58.Terme M, Ullrich E, Aymeric L, Meinhardt K, Desbois M, Delahaye N, et al. IL-18 induces PD-1-dependent immunosuppression in cancer. Cancer Res. 2011;71:5393–9. doi: 10.1158/0008-5472.CAN-11-0993. [DOI] [PubMed] [Google Scholar]
  • 59.Carbone A, Vizio B, Novarino A, Mauri FA, Geuna M, Robino C, et al. IL-18 paradox in pancreatic carcinoma: elevated serum levels of free IL-18 are correlated with poor survival. J Immunother. 2009;32:920–31. doi: 10.1097/CJI.0b013e3181b29168. [DOI] [PubMed] [Google Scholar]
  • 60.Okamoto M, Azuma K, Hoshino T, Imaoka H, Ikeda J, Kinoshita T, et al. Correlation of decreased survival and IL-18 in bone metastasis. Intern Med. 2009;48:763–73. doi: 10.2169/internalmedicine.48.1851. [DOI] [PubMed] [Google Scholar]
  • 61.Fickenscher H, Hör S, Küpers H, Knappe A, Wittmann S, Sticht H. The interleukin-10 family of cytokines. Trends Immunol. 2002;23:89–96. doi: 10.1016/S1471-4906(01)02149-4. [DOI] [PubMed] [Google Scholar]
  • 62.Wang M, Tan Z, Zhang R, Kotenko SV, Liang P. Interleukin 24 (MDA-7/MOB-5) signals through two heterodimeric receptors, IL-22R1/IL-20R2 and IL-20R1/IL-20R2. J Biol Chem. 2002;277:7341–7. doi: 10.1074/jbc.M106043200. [DOI] [PubMed] [Google Scholar]
  • 63.Blumberg H, Conklin D, Xu WF, Grossmann A, Brender T, Carollo S, et al. Interleukin 20: discovery, receptor identification, and role in epidermal function. Cell. 2001;104:9–19. doi: 10.1016/S0092-8674(01)00187-8. [DOI] [PubMed] [Google Scholar]
  • 64.Kotenko SV, Krause CD, Izotova LS, Pollack BP, Wu W, Pestka S. Identification and functional characterization of a second chain of the interleukin-10 receptor complex. EMBO J. 1997;16:5894–903. doi: 10.1093/emboj/16.19.5894. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Sabat R, Grütz G, Warszawska K, Kirsch S, Witte E, Wolk K, et al. Biology of interleukin-10. Cytokine Growth Factor Rev. 2010;21:331–44. doi: 10.1016/j.cytogfr.2010.09.002. [DOI] [PubMed] [Google Scholar]
  • 66.Sheppard P, Kindsvogel W, Xu W, Henderson K, Schlutsmeyer S, Whitmore TE, et al. IL-28, IL-29 and their class II cytokine receptor IL-28R. Nat Immunol. 2003;4:63–8. doi: 10.1038/ni873. [DOI] [PubMed] [Google Scholar]
  • 67.Kunz S, Wolk K, Witte E, Witte K, Doecke WD, Volk HD, et al. Interleukin (IL)-19, IL-20 and IL-24 are produced by and act on keratinocytes and are distinct from classical ILs. Exp Dermatol. 2006;15:991–1004. doi: 10.1111/j.1600-0625.2006.00516.x. [DOI] [PubMed] [Google Scholar]
  • 68.Kühn R, Löhler J, Rennick D, Rajewsky K, Müller W. Interleukin-10-deficient mice develop chronic enterocolitis. Cell. 1993;75:263–74. doi: 10.1016/0092-8674(93)80068-P. [DOI] [PubMed] [Google Scholar]
  • 69.Asadullah K, Sterry W, Volk HD. Interleukin-10 therapy--review of a new approach. Pharmacol Rev. 2003;55:241–69. doi: 10.1124/pr.55.2.4. [DOI] [PubMed] [Google Scholar]
  • 70.De Vita F, Orditura M, Galizia G, Romano C, Lieto E, Iodice P, et al. Serum interleukin-10 is an independent prognostic factor in advanced solid tumors. Oncol Rep. 2000;7:357–61. doi: 10.3892/or.7.2.357. [DOI] [PubMed] [Google Scholar]
  • 71.Lech-Maranda E, Baseggio L, Bienvenu J, Charlot C, Berger F, Rigal D, et al. Interleukin-10 gene promoter polymorphisms influence the clinical outcome of diffuse large B-cell lymphoma. Blood. 2004;103:3529–34. doi: 10.1182/blood-2003-06-1850. [DOI] [PubMed] [Google Scholar]
  • 72.Sarris AH, Kliche KO, Pethambaram P, Preti A, Tucker S, Jackow C, et al. Interleukin-10 levels are often elevated in serum of adults with Hodgkin’s disease and are associated with inferior failure-free survival. Ann Oncol. 1999;10:433–40. doi: 10.1023/A:1008301602785. [DOI] [PubMed] [Google Scholar]
  • 73.Iwakura Y, Ishigame H, Saijo S, Nakae S. Functional specialization of interleukin-17 family members. Immunity. 2011;34:149–62. doi: 10.1016/j.immuni.2011.02.012. [DOI] [PubMed] [Google Scholar]
  • 74.Chang SH, Dong C. Signaling of interleukin-17 family cytokines in immunity and inflammation. Cell Signal. 2011;23:1069–75. doi: 10.1016/j.cellsig.2010.11.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Moseley TA, Haudenschild DR, Rose L, Reddi AH. Interleukin-17 family and IL-17 receptors. Cytokine Growth Factor Rev. 2003;14:155–74. doi: 10.1016/S1359-6101(03)00002-9. [DOI] [PubMed] [Google Scholar]
  • 76.Peters A, Lee Y, Kuchroo VK. The many faces of Th17 cells. Curr Opin Immunol. 2011;23:702–6. doi: 10.1016/j.coi.2011.08.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Ngiow SF, Smyth MJ, Teng MW. Does IL-17 suppress tumor growth? Blood. 2010;115:2554–5, author reply 2556-7. doi: 10.1182/blood-2009-11-254607. [DOI] [PubMed] [Google Scholar]
  • 78.Hueber W, Patel DD, Dryja T, Wright AM, Koroleva I, Bruin G, et al. Psoriasis Study Group. Rheumatoid Arthritis Study Group. Uveitis Study Group Effects of AIN457, a fully human antibody to interleukin-17A, on psoriasis, rheumatoid arthritis, and uveitis. Sci Transl Med. 2010;2:52ra72. doi: 10.1126/scitranslmed.3001107. [DOI] [PubMed] [Google Scholar]
  • 79.Leonardi C, Matheson R, Zachariae C, Cameron G, Li L, Edson-Heredia E, et al. Anti-interleukin-17 monoclonal antibody ixekizumab in chronic plaque psoriasis. N Engl J Med. 2012;366:1190–9. doi: 10.1056/NEJMoa1109997. [DOI] [PubMed] [Google Scholar]
  • 80.Papp KA, Leonardi C, Menter A, Ortonne JP, Krueger JG, Kricorian G, et al. Brodalumab, an anti-interleukin-17-receptor antibody for psoriasis. N Engl J Med. 2012;366:1181–9. doi: 10.1056/NEJMoa1109017. [DOI] [PubMed] [Google Scholar]
  • 81.Katze MG, He Y, Gale M., Jr. Viruses and interferon: a fight for supremacy. Nat Rev Immunol. 2002;2:675–87. doi: 10.1038/nri888. [DOI] [PubMed] [Google Scholar]
  • 82.González-Navajas JM, Lee J, David M, Raz E. Immunomodulatory functions of type I interferons. Nat Rev Immunol. 2012;12:125–35. doi: 10.1038/nri3133. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Platanias LC. Mechanisms of type-I- and type-II-interferon-mediated signalling. Nat Rev Immunol. 2005;5:375–86. doi: 10.1038/nri1604. [DOI] [PubMed] [Google Scholar]
  • 84.Galluzzi L, Senovilla L, Zitvogel L, Kroemer G. The secret ally: immunostimulation by anticancer drugs. Nat Rev Drug Discov. 2012;11:215–33. doi: 10.1038/nrd3626. [DOI] [PubMed] [Google Scholar]
  • 85.Hannani D, Sistigu A, Kepp O, Galluzzi L, Kroemer G, Zitvogel L. Prerequisites for the antitumor vaccine-like effect of chemotherapy and radiotherapy. Cancer J. 2011;17:351–8. doi: 10.1097/PPO.0b013e3182325d4d. [DOI] [PubMed] [Google Scholar]
  • 86.Michaud M, Martins I, Sukkurwala AQ, Adjemian S, Ma Y, Pellegatti P, et al. Autophagy-dependent anticancer immune responses induced by chemotherapeutic agents in mice. Science. 2011;334:1573–7. doi: 10.1126/science.1208347. [DOI] [PubMed] [Google Scholar]
  • 87.Hommes DW, Mikhajlova TL, Stoinov S, Stimac D, Vucelic B, Lonovics J, et al. Fontolizumab, a humanised anti-interferon gamma antibody, demonstrates safety and clinical activity in patients with moderate to severe Crohn’s disease. Gut. 2006;55:1131–7. doi: 10.1136/gut.2005.079392. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Reinisch W, de Villiers W, Bene L, Simon L, Rácz I, Katz S, et al. Fontolizumab in moderate to severe Crohn’s disease: a phase 2, randomized, double-blind, placebo-controlled, multiple-dose study. Inflamm Bowel Dis. 2010;16:233–42. doi: 10.1002/ibd.21038. [DOI] [PubMed] [Google Scholar]
  • 89.Pietras K, Sjöblom T, Rubin K, Heldin CH, Ostman A. PDGF receptors as cancer drug targets. Cancer Cell. 2003;3:439–43. doi: 10.1016/S1535-6108(03)00089-8. [DOI] [PubMed] [Google Scholar]
  • 90.Li H, Fredriksson L, Li X, Eriksson U. PDGF-D is a potent transforming and angiogenic growth factor. Oncogene. 2003;22:1501–10. doi: 10.1038/sj.onc.1206223. [DOI] [PubMed] [Google Scholar]
  • 91.Corless CL, Schroeder A, Griffith D, Town A, McGreevey L, Harrell P, et al. PDGFRA mutations in gastrointestinal stromal tumors: frequency, spectrum and in vitro sensitivity to imatinib. J Clin Oncol. 2005;23:5357–64. doi: 10.1200/JCO.2005.14.068. [DOI] [PubMed] [Google Scholar]
  • 92.Corless CL, Barnett CM, Heinrich MC. Gastrointestinal stromal tumours: origin and molecular oncology. Nat Rev Cancer. 2011;11:865–78. doi: 10.1038/nrc3143. [DOI] [PubMed] [Google Scholar]
  • 93.Linardou H, Dahabreh IJ, Bafaloukos D, Kosmidis P, Murray S. Somatic EGFR mutations and efficacy of tyrosine kinase inhibitors in NSCLC. Nat Rev Clin Oncol. 2009;6:352–66. doi: 10.1038/nrclinonc.2009.62. [DOI] [PubMed] [Google Scholar]
  • 94.Stirewalt DL, Radich JP. The role of FLT3 in haematopoietic malignancies. Nat Rev Cancer. 2003;3:650–65. doi: 10.1038/nrc1169. [DOI] [PubMed] [Google Scholar]
  • 95.Bergers G, Benjamin LE. Tumorigenesis and the angiogenic switch. Nat Rev Cancer. 2003;3:401–10. doi: 10.1038/nrc1093. [DOI] [PubMed] [Google Scholar]
  • 96.Fernandez NC, Lozier A, Flament C, Ricciardi-Castagnoli P, Bellet D, Suter M, et al. Dendritic cells directly trigger NK cell functions: cross-talk relevant in innate anti-tumor immune responses in vivo. Nat Med. 1999;5:405–11. doi: 10.1038/7403. [DOI] [PubMed] [Google Scholar]
  • 97.Lynch DH, Andreasen A, Maraskovsky E, Whitmore J, Miller RE, Schuh JC. Flt3 ligand induces tumor regression and antitumor immune responses in vivo. Nat Med. 1997;3:625–31. doi: 10.1038/nm0697-625. [DOI] [PubMed] [Google Scholar]
  • 98.Baselga J. Targeting tyrosine kinases in cancer: the second wave. Science. 2006;312:1175–8. doi: 10.1126/science.1125951. [DOI] [PubMed] [Google Scholar]
  • 99.Galluzzi L, Vacchelli E, Fridman WH, Galon J, Sautès-Fridman C, Tartour E, et al. Trial Watch - Monoclonal antibodies in cancer therapy. Oncoimmunol. 2012;1:28–37. doi: 10.4161/onci.1.1.17938. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Balachandran VP, Cavnar MJ, Zeng S, Bamboat ZM, Ocuin LM, Obaid H, et al. Imatinib potentiates antitumor T cell responses in gastrointestinal stromal tumor through the inhibition of Ido. Nat Med. 2011;17:1094–100. doi: 10.1038/nm.2438. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Delahaye NF, Rusakiewicz S, Martins I, Ménard C, Roux S, Lyonnet L, et al. Alternatively spliced NKp30 isoforms affect the prognosis of gastrointestinal stromal tumors. Nat Med. 2011;17:700–7. doi: 10.1038/nm.2366. [DOI] [PubMed] [Google Scholar]
  • 102.Anzano MA, Roberts AB, Meyers CA, Komoriya A, Lamb LC, Smith JM, et al. Synergistic interaction of two classes of transforming growth factors from murine sarcoma cells. Cancer Res. 1982;42:4776–8. [PubMed] [Google Scholar]
  • 103.Matzuk MM. Functional analysis of mammalian members of the transforming growth factor-beta superfamily. Trends Endocrinol Metab. 1995;6:120–7. doi: 10.1016/1043-2760(95)00032-D. [DOI] [PubMed] [Google Scholar]
  • 104.Miyazono K. TGF-beta receptors and signal transduction. Int J Hematol. 1997;65:97–104. doi: 10.1016/S0925-5710(96)00542-7. [DOI] [PubMed] [Google Scholar]
  • 105.Schmierer B, Hill CS. TGFbeta-SMAD signal transduction: molecular specificity and functional flexibility. Nat Rev Mol Cell Biol. 2007;8:970–82. doi: 10.1038/nrm2297. [DOI] [PubMed] [Google Scholar]
  • 106.Blobe GC, Schiemann WP, Lodish HF. Role of transforming growth factor beta in human disease. N Engl J Med. 2000;342:1350–8. doi: 10.1056/NEJM200005043421807. [DOI] [PubMed] [Google Scholar]
  • 107.Yang L, Anderson DE, Baecher-Allan C, Hastings WD, Bettelli E, Oukka M, et al. IL-21 and TGF-beta are required for differentiation of human T(H)17 cells. Nature. 2008;454:350–2. doi: 10.1038/nature07021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Towler DA. Bone morphogenetic proteins. Blood. 2009;114:2012–3. doi: 10.1182/blood-2009-06-228544. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Grütter C, Wilkinson T, Turner R, Podichetty S, Finch D, McCourt M, et al. A cytokine-neutralizing antibody as a structural mimetic of 2 receptor interactions. Proc Natl Acad Sci U S A. 2008;105:20251–6. doi: 10.1073/pnas.0807200106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Denton CP, Merkel PA, Furst DE, Khanna D, Emery P, Hsu VM, et al. Cat-192 Study Group. Scleroderma Clinical Trials Consortium Recombinant human anti-transforming growth factor beta1 antibody therapy in systemic sclerosis: a multicenter, randomized, placebo-controlled phase I/II trial of CAT-192. Arthritis Rheum. 2007;56:323–33. doi: 10.1002/art.22289. [DOI] [PubMed] [Google Scholar]
  • 111.Eriksson JE, Vandenabeele P. Workshop summary: cell death mechanisms controlled by the TNF family. Adv Exp Med Biol. 2011;691:585–8. doi: 10.1007/978-1-4419-6612-4_61. [DOI] [PubMed] [Google Scholar]
  • 112.Wang S, El-Deiry WS. TRAIL and apoptosis induction by TNF-family death receptors. Oncogene. 2003;22:8628–33. doi: 10.1038/sj.onc.1207232. [DOI] [PubMed] [Google Scholar]
  • 113.Locksley RM, Killeen N, Lenardo MJ. The TNF and TNF receptor superfamilies: integrating mammalian biology. Cell. 2001;104:487–501. doi: 10.1016/S0092-8674(01)00237-9. [DOI] [PubMed] [Google Scholar]
  • 114.Carswell EA, Old LJ, Kassel RL, Green S, Fiore N, Williamson B. An endotoxin-induced serum factor that causes necrosis of tumors. Proc Natl Acad Sci U S A. 1975;72:3666–70. doi: 10.1073/pnas.72.9.3666. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Kolb WP, Granger GA. Lymphocyte in vitro cytotoxicity: characterization of human lymphotoxin. Proc Natl Acad Sci U S A. 1968;61:1250–5. doi: 10.1073/pnas.61.4.1250. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Ruddle NH, Waksman BH. Cytotoxicity mediated by soluble antigen and lymphocytes in delayed hypersensitivity. 3. Analysis of mechanism. J Exp Med. 1968;128:1267–79. doi: 10.1084/jem.128.6.1267. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Pennica D, Nedwin GE, Hayflick JS, Seeburg PH, Derynck R, Palladino MA, et al. Human tumour necrosis factor: precursor structure, expression and homology to lymphotoxin. Nature. 1984;312:724–9. doi: 10.1038/312724a0. [DOI] [PubMed] [Google Scholar]
  • 118.Galluzzi L, Vitale I, Abrams JM, Alnemri ES, Baehrecke EH, Blagosklonny MV, et al. Molecular definitions of cell death subroutines: recommendations of the Nomenclature Committee on Cell Death 2012. Cell Death Differ. 2012;19:107–20. doi: 10.1038/cdd.2011.96. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Vandenabeele P, Galluzzi L, Vanden Berghe T, Kroemer G. Molecular mechanisms of necroptosis: an ordered cellular explosion. Nat Rev Mol Cell Biol. 2010;11:700–14. doi: 10.1038/nrm2970. [DOI] [PubMed] [Google Scholar]
  • 120.Bouillet P, O’Reilly LA. CD95, BIM and T cell homeostasis. Nat Rev Immunol. 2009;9:514–9. doi: 10.1038/nri2570. [DOI] [PubMed] [Google Scholar]
  • 121.Yang Y, Wilson JM. CD40 ligand-dependent T cell activation: requirement of B7-CD28 signaling through CD40. Science. 1996;273:1862–4. doi: 10.1126/science.273.5283.1862. [DOI] [PubMed] [Google Scholar]
  • 122.Kong YY, Yoshida H, Sarosi I, Tan HL, Timms E, Capparelli C, et al. OPGL is a key regulator of osteoclastogenesis, lymphocyte development and lymph-node organogenesis. Nature. 1999;397:315–23. doi: 10.1038/16852. [DOI] [PubMed] [Google Scholar]
  • 123.Schramek D, Leibbrandt A, Sigl V, Kenner L, Pospisilik JA, Lee HJ, et al. Osteoclast differentiation factor RANKL controls development of progestin-driven mammary cancer. Nature. 2010;468:98–102. doi: 10.1038/nature09387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Loser K, Mehling A, Loeser S, Apelt J, Kuhn A, Grabbe S, et al. Epidermal RANKL controls regulatory T-cell numbers via activation of dendritic cells. Nat Med. 2006;12:1372–9. doi: 10.1038/nm1518. [DOI] [PubMed] [Google Scholar]
  • 125.Wiley SR, Schooley K, Smolak PJ, Din WS, Huang CP, Nicholl JK, et al. Identification and characterization of a new member of the TNF family that induces apoptosis. Immunity. 1995;3:673–82. doi: 10.1016/1074-7613(95)90057-8. [DOI] [PubMed] [Google Scholar]
  • 126.Palladino MA, Bahjat FR, Theodorakis EA, Moldawer LL. Anti-TNF-alpha therapies: the next generation. Nat Rev Drug Discov. 2003;2:736–46. doi: 10.1038/nrd1175. [DOI] [PubMed] [Google Scholar]
  • 127.Brown JE, Coleman RE. Denosumab in patients with cancer-a surgical strike against the osteoclast. Nat Rev Clin Oncol. 2012;9:110–8. doi: 10.1038/nrclinonc.2011.197. [DOI] [PubMed] [Google Scholar]
  • 128.Lewiecki EM. New targets for intervention in the treatment of postmenopausal osteoporosis. Nat Rev Rheumatol. 2011;7:631–8. doi: 10.1038/nrrheum.2011.130. [DOI] [PubMed] [Google Scholar]
  • 129.Grünhagen DJ, de Wilt JH, ten Hagen TL, Eggermont AM. Technology insight: Utility of TNF-alpha-based isolated limb perfusion to avoid amputation of irresectable tumors of the extremities. Nat Clin Pract Oncol. 2006;3:94–103. doi: 10.1038/ncponc0426. [DOI] [PubMed] [Google Scholar]
  • 130.Deroose JP, Eggermont AM, van Geel AN, de Wilt JH, Burger JW, Verhoef C. 20 years experience of TNF-based isolated limb perfusion for in-transit melanoma metastases: TNF dose matters. Ann Surg Oncol. 2012;19:627–35. doi: 10.1245/s10434-011-2030-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Deroose JP, Grünhagen DJ, van Geel AN, de Wilt JH, Eggermont AM, Verhoef C. Long-term outcome of isolated limb perfusion with tumour necrosis factor-α for patients with melanoma in-transit metastases. Br J Surg. 2011;98:1573–80. doi: 10.1002/bjs.7621. [DOI] [PubMed] [Google Scholar]
  • 132.Halin C, Gafner V, Villani ME, Borsi L, Berndt A, Kosmehl H, et al. Synergistic therapeutic effects of a tumor targeting antibody fragment, fused to interleukin 12 and to tumor necrosis factor alpha. Cancer Res. 2003;63:3202–10. [PubMed] [Google Scholar]
  • 133.Clement JM, McDermott DF. The high-dose aldesleukin (IL-2) “select” trial: a trial designed to prospectively validate predictive models of response to high-dose IL-2 treatment in patients with metastatic renal cell carcinoma. Clin Genitourin Cancer. 2009;7:E7–9. doi: 10.3816/CGC.2009.n.014. [DOI] [PubMed] [Google Scholar]
  • 134.Halama N, Zoernig I, Jaeger D. Advanced malignant melanoma: immunologic and multimodal therapeutic strategies. J Oncol 2010; 2010:689893. [DOI] [PMC free article] [PubMed]

Articles from Oncoimmunology are provided here courtesy of Taylor & Francis

RESOURCES