• Open Access
    Review

    Type 1 and type 2 cytokine-mediated immune orchestration in the tumour microenvironment and their therapeutic potential

    Eric Jou 1,2*

    Explor Target Antitumor Ther. 2023;4:474–497 DOI: https://doi.org/10.37349/etat.2023.00146

    Received: February 27, 2023 Accepted: April 25, 2023 Published: June 30, 2023

    Academic Editor: Sufi Thomas, University of Kansas Medical Center, USA

    This article belongs to the special issue Therapeutic Targeting of the Tumor Microenvironment

    Abstract

    Cancer remains the second leading cause of death worldwide despite modern breakthroughs in medicine, and novel treatments are urgently needed. The revolutionary success of immune checkpoint inhibitors in the past decade serves as proof of concept that the immune system can be effectively harnessed to treat cancer. Cytokines are small signalling proteins with critical roles in orchestrating the immune response and have become an attractive target for immunotherapy. Type 1 immune cytokines, including interferon γ (IFNγ), interleukin-12 (IL-12), and tumour necrosis factor α (TNFα), have been shown to have largely tumour suppressive roles in part through orchestrating anti-tumour immune responses mediated by natural killer (NK) cells, CD8+ T cells and T helper 1 (Th1) cells. Conversely, type 2 immunity involving group 2 innate lymphoid cells (ILC2s) and Th2 cells are involved in tissue regeneration and wound repair and are traditionally thought to have pro-tumoural effects. However, it is found that the classical type 2 immune cytokines IL-4, IL-5, IL-9, and IL-13 may have conflicting roles in cancer. Similarly, type 2 immunity-related cytokines IL-25 and IL-33 with recently characterised roles in cancer may either promote or suppress tumorigenesis in a context-dependent manner. Furthermore, type 1 cytokines IFNγ and TNFα have also been found to have pro-tumoural effects under certain circumstances, further complicating the overall picture. Therefore, the dichotomy of type 1 and type 2 cytokines inhibiting and promoting tumours respectively is not concrete, and attempts of utilising these for cancer immunotherapy must take into account all available evidence. This review provides an overview summarising the current understanding of type 1 and type 2 cytokines in tumour immunity and discusses the prospects of harnessing these for immunotherapy in light of previous and ongoing clinical trials.

    Keywords

    Tumour microenvironment, cancer therapy, targeted therapy, preclinical models, immunotherapy, cytokines, oncology trials

    Introduction

    Despite modern medical breakthroughs, cancer is now the second leading cause of death worldwide [1], and global cancer incidence is projected to increase by a further 50% in the next 20 years [2]. The urgent development of novel cancer therapies is a key aspect of contemporary medicine and is more vital than ever. Whilst cancer immunotherapy, in particular immune checkpoint inhibitors which act by revitalising T cells, have completely revolutionised cancer treatment resulting in immunotherapy being erected as the fourth pillar of cancer treatment alongside traditional surgery, chemotherapy, and radiotherapy, not all cancer types or patients respond [35]. Identifying novel cancer immunotherapies to overcome therapeutic resistance is challenging yet crucial, and currently, two-thirds of ongoing clinical trials in oncology involve immunotherapy [5].

    Cytokines are small secreted proteins (6–70 kDa) produced by a broad range of immune and non-immune cell types allowing intercellular communication with key roles in orchestrating the immune response [6]. During protective immunity against pathogens, the specific types and combinations of cytokines released are tightly coordinated in order to elicit the optimal immune response for pathogen clearance [7, 8]. Similarly, the immune system is able to actuate cancer immunosurveillance and tumour rejection through an array of anti-tumour cytokines and effector immune cells [9]. However, established malignancies are able to create a tumour microenvironment that supports neoplastic growth through the release of immunomodulatory cytokines [10], and indeed, immunoevasion is a key hallmark of cancer as described by Hanahan and Weinberg [11].

    The immune system can be broadly classified into opposing type 1 and type 2 immunity each involving a unique array of cytokines with distinct functions [7, 12]. The type 1 immune cytokines interferon γ (IFNγ), interleukin-12 (IL-12), and tumour necrosis factor α (TNFα) orchestrate the protective type 1 immunity consisting of natural killer (NK) cells, CD8+ cytotoxic T cells, and CD4+ T helper 1 (Th1) cells, against intracellular pathogens such as viruses or intracellular bacteria [7, 13, 14]. On the other hand, the type 2 immune response has important roles in anti-helminth immunity, tissue regeneration, and wound repair, and is orchestrated by group 2 innate lymphoid cells (ILC2s) and Th2 cells with effector functions driven by the cytokines IL-4, IL-5, IL-9, and IL-13 [8, 15, 16]. Traditionally, type 1 immunity has been ascribed a predominantly anti-tumoural role showcased by many preclinical and clinical studies, while type 2 immunity harboured by the tumour microenvironment is associated with wound healing and repair thereby creating a pro-tumorigenic niche (Figure 1) [1725]. Nevertheless, recent studies indicate that the dichotomy between type 1 and type 2 cytokines in tumour immunity is not concrete as previously envisaged [2629]. Under certain contexts, type 1 immunity may promote cancer development through chronic inflammation, while type 2 cytokines may elicit anti-tumour immunity through blood vessel remodelling and macrophage recruitment [2629]. Therefore, any attempts to modify the tumour microenvironment through cytokine-based immunotherapy must be carefully tailored to the specific context in order to achieve the desired anti-tumoural effect.

    Overview of type 1 and type 2 cytokines in tumour immunity. M1 macrophages produce IL-12 which activates NK cells and CD8+ T cells and induces Th1 cell polarisation. Activated NK cells, CD8+ T cells, and Th1 cells in turn produce IFNγ which reciprocally activates M1 macrophages. T cells and M1 macrophages are also capable of producing TNFα. ILC2s and Th2 cells produce the type 2 cytokines IL-4, IL-5, IL-9, and IL-13. IL-4 and IL-13 induce M2 macrophages and immunosuppressive myeloid-derived suppressor cells (MDSCs), which inhibit anti-tumoural T cells. IL-5 and IL-9 promote eosinophils and mast cells respectively, and eosinophils may inhibit NK cell function. Overall, type 1 cytokines orchestrate an anti-tumour immune response leading to tumour cell apoptosis, while type 2 cytokines promote tumorigenesis. Sharp arrows indicate activation or stimulation while blunted arrows indicate inhibition. Figure was created in part using cartoon templates by Servier Medical Art (https://smart.servier.com/), licensed under a Creative Commons Attribution 3.0 Unported License

    This review provides an overview of the current understanding of the classical type 1 and type 2 cytokines in the tumour microenvironment, with particular focus on their role in tumour immunity. The type 2 immunity-related cytokines IL-25 and IL-33 with recently discovered roles in cancer are also explored in light of their potential as novel immunotherapeutic targets [30, 31]. Existing cytokine-based cancer immunotherapies and the broader prospects of harnessing cytokines for future cancer treatments are also discussed.

    Type 1 cytokines in tumour immunity

    In general, cytokines involved in type 1 immunity including IFNγ, IL-12, and TNFα are shown to orchestrate anti-tumour immune responses against a broad range of cancer types in numerous preclinical and clinical studies [32]. IFNγ is the main effector cytokine of type 1 immune response and is established as the prototypical anti-tumoural cytokine with critical roles in anti-cancer immunity and tumour rejection [33]. IL-12 mainly exerts function through inducing IFNγ expression [34, 35], while TNFα has conflicting roles and may either promote or suppress tumorigenesis [36].

    IFNγ

    IFNs are cytokines with important roles in antiviral immunity and consist of three major families, designated as type I, II, and III [37]. Whilst type I and type III IFNs predominantly protect the host from pathogens, IFNγ being the sole member of type II IFNs, has additional well-documented functions in cancer immunosurveillance [3841]. In the tumour microenvironment, sources of IFNγ include tumour-infiltrating NK cells, ILC1s, γδ T cells, CD8+ T cells, and Th1 cells [4244]. IFNγ signaling is mediated by the IFNγ receptor (IFNGR), leading to downstream activation of Janus kinase 1 (JAK1) and JAK2, respectively, and subsequent phosphorylation of the transcription factor signal transducer and activator of transcription 1 (STAT1) [45]. Nuclear translocation of the phosphorylated STAT1 homodimer then ensues, allowing subsequent binding to the corresponding gamma-activated sequence (GAS) DNA element and induction of IFN-stimulated gene (ISG) expression (Figure 2).

    Overview of IFNγ signalling and anti-tumoural properties. Sharp arrows indicate activation; circled P symbols indicate phosphorylation. Figure was created in part using cartoon templates by Servier Medical Art (https://smart.servier.com/), licensed under a Creative Commons Attribution 3.0 Unported License

    IFNγ signaling in cancer cells directly results in apoptosis through the induction of an array of pro-apoptotic genes including absent in melanoma 2 (AIM2) gene, IFN-induced transmembrane protein 2 (IFITM2) gene, inositol hexakisphosphate kinase 2 (IP6K2) gene, ISG12a, ISG54, phorbol-12-myristate-13-acetate-induced protein 1 (NOXA) gene, and neuronal precursor cell-expressed developmentally down-regulated protein 8 (NEDD8) ultimate buster 1 (NUB1), which ultimately results in a reduction in antiapoptotic B-cell lymphoma-2 (BCL-2) proteins and downstream activation of caspase 3 leading to cell death [46, 47], and IFNγ has also been shown to promote cell death via induction of nitric oxide (NO) or reactive oxygen species (ROS)-mediated pathways [48]. Conversely, IFNγ binding to receptors on anti-tumour type 1 immune cells typically results in activation and induction of effector function [49]. Accordingly, in mouse models of colorectal cancer (CRC), heterozygous genetic deletion of IFNγ (Ifng+/-), homozygous deletion of the corresponding receptor (Ifngr1-/-), or antibody-mediated neutralisation of IFNγ, increased adenocarcinoma development [24, 41, 50]. In humans, an IFNG-associated type 1 immune signature is associated with improved prognosis in human CRC patients [51], and similarly predicts prolonged patient survival across multiple other cancer types such as breast, liver, and ovarian cancers [5254]. Mechanistically, IFNγ induces the expression of cytokines IL-1β, IL-6, IL-12, IL-18, IL-23, TNFα, and proinflammatory factors such as ROS and NO by M1 macrophages, thereby contributing to M1 macrophage-mediated anti-tumour immunity [55, 56]. Both IFNγ and TNFα promote M1 macrophage polarisation [57, 58], and induction of M1 macrophages is associated with reduced tumour burden in mouse CRC [59]. Similarly in humans, increased tumour M1 macrophage infiltration is associated with improved survival in multiple cancer types including ovarian [60], lung [61], and hepatocellular carcinoma [62].

    Furthermore, IFNγ stimulates major histocompatibility complex class II (MHC-II) expression on professional antigen-presenting cells (pAPCs) such as dendritic cells (DCs) and macrophages through the transcriptional coactivator MHC class II transactivator (CIITA) [63], and together with M1 macrophage-derived IL-12, promotes T cell activation and subsequent polarisation towards a Th1 phenotype [64, 65]. As mentioned, Th1 cells are potent producers of IFNγ thus providing positive feedback, and both Th1 cells and IFNγ play critical roles in CD8+ cytotoxic T cell activation, the latter directly while the former indirectly through licensing XC chemokine receptor 1+ (XCR1+) DCs [43, 66]. XCR1+ DCs are specialised in cross-presentation of antigens, which is critical in efficient CD8+ T cell activation [67]. Whilst CD8+ T cells, like Th1 cells, are potent producers of IFNγ, they additionally exert direct cytotoxicity on tumour cells through Fas ligand (FasL) or perforin and granzyme which can be further enhanced in response to IFNγ signalling [68]. Cytotoxic CD8+ T cells have well-established roles in eliminating cancer cells in vivo [69], and are the most potent anti-tumour immune cell type known against human cancer [70].

    Various other mechanisms of IFNγ-mediated anti-tumour immunity have been proposed (Figure 2), for example by decreasing cancer stem cells in the 4T1 mouse model of breast cancer [43, 71]. IFNγ induces cancer cell senescence and arrest via activating p16Ink4a/p19Arf and p21Cip1 thereby inhibiting tumorigenesis in models of pancreatic islet cancer and Burkitt lymphoma [72]. Furthermore, in a heterotypic xenograft mouse model of human gallbladder cancer, treatment with IFNγ reduced tumorigenesis by inhibiting angiogenesis. This is via suppression of M2 macrophage vascular endothelial growth factor (VEGF) production and highlights the therapeutic potential of IFNγ [73]. Overall, IFNγ is capable of inducing anti-tumour immunity via direct action on tumour cells or indirectly through stimulating type 1 immunity.

    IL-12

    IL-12 is part of the IL-12 family of cytokines which includes IL-23, IL-27, IL-35, and IL-39, each with very distinct functions [74, 75]. For instance, whilst IL-12 is crucial in Th1 polarisation through the induction of STAT4 and T-box expressed in T cells (Tbet) thereby eliciting anti-tumour immunity [65], IL-23 participates in Th17 immune responses [76] while IL-35 is a recently characterised cytokine with immunosuppressive and pro-tumorigenic properties [77]. IL-12 is produced by the pAPCs macrophages and DCs, and signalling is mediated through binding to the IL-12 receptor (IL-12R), a heterodimer consisting of IL-12Rβ1 and IL-12Rβ2, on target cells [78]. Mechanistically, IL-12 exerts potent anti-tumoural functions by coordinating the aforementioned anti-tumour type 1 response and suppresses alternative T cell fates by inhibiting Th2, Th17, and regulatory T (Treg) cell differentiation [34, 35]. IL-12-induced Th1 cells in turn activate NK cells and CD8+ T cells, all of which are potent IFNγ producers [79]. IFNγ can further enhance IL-12 production by pAPCs thereby creating a positive feedback loop [80]. Mice genetically deficient in functional IL-12 are more susceptible to skin cancer and sarcoma development [81, 82], while treatment with recombinant IL-12 delayed tumorigenesis, reduced peritoneal seeding, and prolonged survival in mice with metastatic ovarian cancer [83]. IL-12 can also induce IFNγ-independent CRC tumour rejection via activation of CD8+ T cells, a process that is dependent on CD4+ T cells and granulocyte-macrophage colony-stimulating factor (GM-CSF) [84]. Similarly, in humans, polymorphisms in the IL-12 genes IL-12B (encoding the p40 subunit of IL-12, A > C, rs3212227), are associated with increased risk of breast cancer when harbouring the A allele [odds ratio (OR) = 1.68, 95% confidence interval (CI) 1.09–2.59] [85], while the IL-12A rs568408 GA/AA variant is associated with increased risk of cervical cancer (OR = 1.43, 95% CI 1.06–1.93) [86].

    TNFα

    On the other hand, unlike IL-12 and IFNγ with predominant anti-tumoural properties, the role of TNFα in tumorigenesis is context-dependent. Whilst initially discovered as an anti-tumoural cytokine capable of inducing necrosis of sarcomas in mice [87], more recent studies have uncovered pro-tumoural aspects of TNFα [36]. TNFα is produced by macrophages (similar to IL-12), and additionally by NK cells and activated T cells. TNFα elicits antigen-independent killing of tumour cells by CD8+ T cells and NK cells [88], and can directly induce target cell death through signalling [89, 90]. Conversely, genetic deficiency of TNFα enhanced tumour burden in a mouse model of skin cancer [91], and in humans elevated TNFα expression is associated with breast cancer recurrence [92] and poor prognosis in ovarian cancer patients [93]. TNFα may promote tumours through various proposed mechanisms. TNFα may promote breast cancer cell stemness through the nuclear factor kappa B (NF-κB) pathway [94, 95], or through inducing dedifferentiation of melanoma cells resulting in downregulation of tumour antigens thereby promoting immunoevasion [96]. Furthermore, TNFα may directly promote Treg activation through the TNF receptor type 2 (TNFR2) thereby suppressing anti-tumour immunity [97].

    Type 1 cytokines in immunotherapy

    Type 1 immune cytokines are detected in a broad range of cancer types with well-established prognostic roles and are therefore attractive therapeutic targets for cancer treatment [36, 98100]. Due to its potent anti-tumoural functions, single agent IFNγ therapy was tested in some clinical trials in the 1990s for cancer treatment however showed largely inconsistent results [101]. Given that the tumour microenvironment is highly immunosuppressive with complex interactions between tumour cells, stroma, and infiltrating immune cells, single-agent therapies are unlikely to be adequate in overcoming tumour immune evasion [102]. Indeed, most modern clinical trials involving IFNγ largely focus on combinational therapies to overcome therapeutic resistance. In phase III clinical trial of 148 patients with the International Federation of Gynecology and Obstetrics (FIGO) stage Ic–IIIc ovarian cancer, the addition of subcutaneous IFNγ treatment to the combined cisplatin and cyclophosphamide regime improved progression-free survival (PFS) to 51% in the treatment arm compared to 38% in the control group [103]. This trial serves as a proof of concept demonstrating that IFNγ can be utilised in the treatment of cancer. Furthermore, in a phase II trial of patients with ovarian, fallopian tube, and primary peritoneal cancer, the addition of IFNγ and GM-CSF to standard carboplatin treatment led to patient-reported improvements in quality of life [104].

    However, a phase II trial (NCT00786643) adding IFNγ to the 5-fluorouracil, leucovorin, and bevacizumab regimen for metastatic CRC showed no additional benefits. Importantly, recent studies have uncovered potential pro-tumoural functions of IFNγ, with proposed mechanisms including induction of immunosuppressive programmed death-ligand 1 (PD-L1) and indoleamine 2,3-dioxygenase (IDO) expression, along with induction of CD8+ T cell apoptosis [105]. Further studies are required to understand the specific conditions that render IFNγ pro-tumoural. Critically, immunotherapeutic treatment with IFNγ must be carefully designed to maximise efficacy and to prevent iatrogenic harm through tumour induction, especially as all upcoming and ongoing clinical trials continue to utilise IFNγ as a potential treatment agent for cancer (Table 1). Furthermore, IFNγ treatment may be associated with toxicities for example through overactivation of macrophages. In a pilot trial where two patients with synovial sarcoma were treated with an addition of IFNγ and IL-2 to the combined cyclophosphamide and adoptive T cell transfer therapy, although one patient showed significant tumour regression, the other suffered fatal histiocytic myocarditis [106]. Studies on human melanoma biopsies found an IFNγ signature to be associated with a high response to checkpoint inhibitor therapy in patients, and in vitro exposure of 58 distinct human cell lines to IFNγ induced a similar signature [107]. This indicates that combinational treatment with IFNγ may potentiate immune checkpoint inhibitor efficacy, and this exciting prospect will be tested in upcoming trials (Table 1).

    A comprehensive list of ongoing oncology clinical trials involving IFNγ and IL-12, either alone or in combination with other cytokine-based therapies

    CytokineClinical trial IDPhaseNumber of patientsCancer typeTreatmentStatusEstimated study completion date
    IFNγNCT03112590I/II51HER2+ breast cancerIFNγ plus paclitaxel, trastuzumab and pertuzumabActiveJune 2023
    IFNγNCT03063632II28Mycosis fungoides, Sezary syndrome, and advanced synovial sarcomaIFNγ plus pertuzumabActiveApril 2023
    IFNγNCT05268172I40Malignant pleural effusion tumoursIFNγ plus T cellsRecruitingDecember 2023
    IFNγNCT03747484I/II16Metastatic or unresectable Merkel cell cancerIFNγ plus T cells and avelumab or pembrolizumabRecruitingDecember 2025
    IL-12NCT02555397I15Prostate cancerAdenovirus-mediated cytotoxic and IL-12 gene therapyActiveFebruary 2023
    IL-12NCT04491955II23Small bowel cancer and CRCNHS-IL-12 plus CV301 vaccine, M7824 (anti-PD-L1/TGFβ Trap fusion protein) and N-803 (IL-15 superagonist) combination immunotherapyActiveJuly 2024
    IL-12NCT03439085II77HPV associated malignanciesDNA plasmid-encoding IL-12/HPV DNA plasmids vaccine (MEDI0457) plus durvalumabActiveDecember 2022
    IL-12NCT04015700I9GlioblastomaPlasmid encoded IL-12 with personalised neoantigen DNA vaccineActiveApril 2023
    IL-12NCT04911166I16Metastatic non-small cell lung cancerAdenovirus-mediated IL-12 gene therapy plus atezolimumabRecruitingJune 2024
    IL-12NCT05162118I/II51Advanced pancreatic cancerVG161 (IL-12/IL-15/PD-L1 blocking peptide) oncolytic HSV1 injection plus nivolumabRecruitingDecember 2025
    IL-12NCT04287868I/II51Advanced HPV-associated malignanciesNHS-IL-12 plus PDS0101 and M7824 (anti-PD-L1/TGFβ Trap fusion protein)RecruitingJanuary 2024
    IL-12NCT05392699I60Advanced solid tumoursHuman IL-12 mRNARecruitingJanuary 2027
    IL-12NCT04806464I44Primary liver cancerIL-12/IL-15/PD-L1 blocking peptide oncolytic HSV1 injectionRecruitingDecember 2022
    IL-12NCT04708470I/II90HPV-associated malignancies, small bowel cancer, and CRCNHS-IL-12 plus bintrafusp alfa and entinostatRecruitingDecember 2024
    IL-12NCT04471987I94Advanced or metastatic solid tumoursIL-12-L19L19RecruitingDecember 2023
    IL-12NCT04388033I/II10GlioblastomaIL-12 plus DC tumour vaccine and temozolomideRecruitingDecember 2023
    IL-12NCT02483312I9Acute myeloid leukemiaIL-12RecruitingFebruary 2022
    IL-12NCT04758897I18Advanced malignant solid tumoursVG161 (IL-12/IL-15/PD-L1 blocking peptide) oncolytic HSV1 injectionRecruitingDecember 2022
    IL-12NCT05477849I30Advanced malignant solid tumoursIL-12/IL-15 dual-regulated oncolytic HSV1 injectionRecruitingDecember 2024
    IL-12NCT04303117I/II64Kaposi sarcomaNHS-IL-12 with or without M7824 (anti-PD-L1/TGFβ Trap fusion protein)RecruitingDecember 2025
    IL-12NCT05717712I18GliomaAdenovirus-mediated-non-secretory-IL-12RecruitingJanuary 2028
    Display full size

    HER2: human epidermal growth factor receptor 2; TGFβ: transforming growth factor β; HPV: human papillomavirus; HSV1: herpes simplex virus 1; mRNA: messenger RNA

    IL-12 has been considered a strong candidate for cytokine-based immunotherapy due to its potent properties in promoting anti-tumour type 1 immunity and IFNγ responses [108]. However, direct systemic administration of IL-12 is associated with significant toxicities. In a phase 2 study of 17 patients with advanced renal cell carcinoma, systemic administration of IL-12 led to severe toxicities resulting in 12 patients being hospitalised and two deaths likely due to overwhelming IFNγ release [109]. In phase I/II trial of 33 patients with metastatic melanoma, autologous transfer of tumour infiltration lymphocytes transduced with the gene encoding IL-12 showed enhanced anti-tumour efficacy, but was associated with significant toxicities including high fevers, liver dysfunction, and life-threatening haemodynamic instability [110]. Potential alternative methods to deliver IL-12 are being developed through preclinical models in an attempt to improve safety, for example through nanoparticles [111]. In an animal model of pancreatic cancer, oncolytic virus-mediated delivery of IL-12 which lacks the signalling peptide enhanced survival while inducing minimal toxicities [112]. Altogether, overcoming toxicities in upcoming clinical trials remains the major challenge of implementing IL-12 for cancer treatment (Table 1).

    Methods to further improve the efficacy of IL-12-based therapies are being tested. As both IFNγ and IL-12 activate T cells, adoptive transfer of T cells with co-administration of these cytokines may enhance anti-tumour immunity and prevent T cell exhaustion as seen in mouse models of melanoma [113, 114]. The T cell activating properties of IL-12 may further allow synergy with checkpoint inhibitors. NHS-IL-12, a recombinant fusion protein consisting of IL-12 fused to the human monoclonal immunoglobulin G1 (IgG1) antibody NHS76, has been shown to have improved tumour targeting abilities and longer plasma half-life compared to IL-12, resulting in further enhanced activation of pAPCs and reduction in tumour growth in a mouse model of bladder cancer [115]. These are all being investigated in ongoing clinical trials with exciting prospects (Table 1).

    Finally, whilst TNFα has well-established roles in promoting or inhibiting tumorigenesis, the majority of existing trials to date focussed on enhancing TNFα signalling in cancer treatment. This is based on the rationale that high levels of TNFα promote tumour rejection, while chronic low levels of sustained TNFα expression, as seen in the tumour microenvironment, instead promotes tumorigenesis [36]. Indeed, preclinical studies support the notion of exogenous TNFα inhibiting tumorigenesis. In a mouse xenograft model of human breast cancer, TNFα enhanced the cytotoxicity exerted by combined chemotherapy with docetaxel, 5-fluorouracil, and cisplatin [116]. However, the majority of clinical trials involving systemic TNFα treatment were limited by sepsis-like symptoms and showed no efficacy in tumour rejection at the maximally tolerated dose in phase I and phase II trials [117119]. A potential alternative to boost TNFα signalling and improve the safety profile is through the removal of the inhibitory plasma soluble TNFR rather than direct administration of TNFα, and ongoing clinical trials that explore this are listed in Table 2. Similar to IL-12-based therapies, reducing therapeutic toxicities is vital for the future success of utilising TNFα in cancer treatment.

    All ongoing oncology clinical trials involving the removal of soluble TNFR

    Cytokine Clinical trial IDPhaseNumber of patientsCancer typeTreatmentStatusEstimated study completion date
    TNFαNCT04004910I/II170Advanced breast cancerThree-way comparison between plasma soluble TNFR pulldown with or without chemotherapy, and chemotherapy aloneRecruitingJuly 2023
    TNFαNCT04142931I30Stage IV non-small cell lung cancer, stage IV melanoma, triple-negative breast cancer, or stage IV renal cell carcinomaReduction of soluble TNFR, with or without nivolumabRecruitingDecember 2022
    TNFαNCT04690686II24Non-small cell lung cancerReduction of soluble TNFR alone, or with atezolizumab or paclitaxelRecruitingDecember 2022
    Display full size

    Type 2 cytokines in tumour immunity

    The key cytokines that partake in type 2 immunity include IL-4, IL-5, IL-9, and IL-13 with critical roles in protective anti-helminth immunity, wound healing, and tissue regeneration [8, 15, 16]. ILC2s are recently discovered tissue-resident cells found mainly at epithelial and mucosal barriers that rapidly respond to acute immune perturbations [120] and represent the dominant early innate source of these cytokines [13, 121, 122]. Within the innate immune system, mast cells and eosinophils may also contribute to IL-4 and IL-13 production [123125]. In adaptive immunity, Th2 cells are the major source of IL-4, IL-5, and IL-13 while the closely related Th9 cells produce IL-9, thereby amplifying the type 2 immune response [122, 126]. T cell polarisation towards a Th2 or Th9 phenotype is driven by IL-4 [127], with Th9 requiring additional TGFβ signals [126]. In addition, ILC2s play critical roles in Th2 cell polarisation in both IL-4-dependent and independent manners thus shaping the overall type 2 inflammatory environment [122]. In the context of cancer, the type 2 cytokines IL-4, IL-5, IL-9, and IL-13 along with ILC2s, Th2, and Th9 cells may either promote or inhibit tumorigenesis in a context-dependent manner that involves complex interactions with cancer cells and the constituents of the tumour microenvironment.

    IL-4 and IL-13

    IL-4, the prototypical type 2 cytokine, exerts its function through binding to IL-4R, which is comprised of the IL-4Rα subunit and the common gamma chain (γc, Figure 3) [128]. Early studies of IL-4 soon after its discovery found that overexpression of IL-4 by tumour cell lines through genetic approaches led to tumour rejection after in vivo implantation, consistent with an anti-tumoural role [129131]. This was observed in a wide variety of tumour cell line cancer models such as renal cell tumour [129], CRC [131], myeloma, and breast cancer [130]. However, the anti-tumoural functions of IL-4 may only occur at supraphysiological levels, as most subsequent studies found that endogenously expressed IL-4 instead exert a predominantly pro-tumoural role. In a mouse model of colitis-associated cancer, genetic IL-4-deficiency (thus removing endogenous IL-4 signals) reduced tumour burden [132], and genetic deletion of STAT6, the downstream signaling mediator for IL-4, similarly reduced tumorigenesis in a model of adenomatous polyposis coli (APC)-mutation-mediated CRC [133]. IL-4 may promote tumorigenesis through multiple proposed mechanisms, including through the induction of Th2 cells which produce other pro-tumoural cytokines such as IL-13 [24]. IL-4-signalling induces alternative activation of macrophages to an M2 tumour-associated macrophage (TAM) phenotype with well-documented pro-tumoural functions in the tumour microenvironment [59, 134]. Tumour cells express elevated levels of IL-4R, and IL-4 signaling reduces cancer cell apoptosis through the upregulation of anti-apoptotic genes [135] while also directly promoting tumour cell proliferation [136]. Together with IL-13, IL-4 may also promote MDSCs to inhibit anti-tumour immunity as shown recently where MDSC-mediated CD8+ T cell suppression is critically dependent on IL-4 and IL-13 signalling [24].

    Overview of IL-4 and IL-13 signalling. TYK2: tyrosine kinase 2; AP-1: activator protein 1. Sharp arrows indicate activation; circled P symbols indicate phosphorylation. Figure was created in part using cartoon templates by Servier Medical Art (https://smart.servier.com/), licensed under a Creative Commons Attribution 3.0 Unported License

    Like IL-4, IL-13-signalling is mediated by STAT6, through binding to IL-13R which is comprised of an IL-4Rα subunit shared with the IL-4R, coupled with an IL-13Rα1 subunit that provides specificity (Figure 3) [128]. IL-13 similarly promotes M2 TAM polarisation and MDSC-mediated T cell suppression thereby contributing to tumorigenesis [29, 59, 134]. In a model of sporadic CRC, genetic IL-13 deficiency significantly reduced tumour burden, and adoptive transfer of IL-13+ ILC2s compared to IL-13 ILC2s led to enhanced MDSC activation and downstream suppression of anti-tumoural IFNγ+ CD8+ T cells and Th1 cells [24]. Importantly, when culturing MDSCs with ILC2-derived factors which enhance MDSC-mediated T cell suppression, neutralisation of ILC2-derived IL-4 and IL-13 reprogrammed MDSCs to an anti-tumoural phenotype characterised by enhanced CD8+ T cell activation and IFNγ+ and granzyme expression. Similarly in a study on bladder cancer, detection of IL-13 in patient urine samples correlated with increased ILC2 and MDSCs prevalence and reduced T cells [137]. IL-13 may also directly promote cancer cell proliferation and metastasis through an alternative IL-13Rα2 (Figure 3) for example in pancreatic cancer [138, 139]. In humans, IL-13R expression is increased in a broad range of solid tumours including CRC, glioblastoma, breast cancer, and pancreatic cancer, and is associated with poor prognosis [140].

    IL-5 and IL-9

    IL-5 contributes to type 2 immunity mainly through eosinophils. Eosinophil development is IL-5-dependent, and genetic ablation of the IL-5R subunit IL-5Rα in mice results in defective eosinophils [141]. Likewise, human eosinophil progenitors also express IL-5Rα [142]. Furthermore, IL-5 directly recruits and activates mature eosinophils, and both IL-5 and eosinophils have been implicated in cancer [143, 144]. The majority of animal and human studies seem to suggest an anti-tumoural role of IL-5 and eosinophils, suggesting them as the dominant effectors of anti-tumoural responses by type 2 immunity. Eosinophils can directly induce cancer cell lysis through the release of cytotoxic granules in a human CRC cell line model [145], or indirectly through promoting CD8+ T cell-mediated tumour rejection in mouse melanoma [146]. In human patients, increased eosinophils have been associated with improved prognosis across several cancer types including CRC, melanoma, and breast cancer [147150].

    Nevertheless, IL-5 may have pro-tumorigenic roles similar to other type 2 cytokines such as IL-4 and IL-13. A recent study found that IL-5-activated eosinophils upregulated glycolysis, which resulted in depletion of glucose in the tumour microenvironment [151]. This resulted in increased cancer metastasis to the lungs, due to impaired NK cell effector function which is critically dependent on glucose. Similarly, others have found IL-5 and eosinophils to promote lung metastasis by CRC MC38 cells, however instead through the recruitment of Tregs in response to the eosinophil-derived chemokine C-C motif chemokine 22 (CCL22) [144]. In this model, genetic deficiency or antibody-mediated neutralisation of IL-5 reduced metastasis, while adoptive transfer of eosinophils enhanced tumour spread, consistent with a pro-tumoural role. Therefore, the role of IL-5 and eosinophils in tumour development is likely context-dependent and may vary depending on the specific cancer subtype. Indeed, other prognostic studies in humans have found eosinophils to be associated with poor survival in patients with Hodgekin’s lymphoma, leukaemia, and cervical cancer [152154].

    IL-9 is produced by ILC2s and Th9 cells, and like IL-5, has been shown to either promote or inhibit tumorigenesis. In nude mice, exogenous IL-9 treatment inhibited gastric cancer growth [155]. Interestingly, in this model, IL-9 treatment is associated with a reduction in serum IL-4, indicating potential opposing functions. In a mouse model of colitis-associated cancer, IL-9 suppressed tumour growth through CD8+ T cell activation [156]. Similarly, findings were observed in melanoma where IL-9 supports anti-tumoural CD4+CD8+ double-positive T cells directly inducing their proliferation and preventing apoptosis via signaling through the IL-9R [157]. Overexpression of IL-9 by the CT26 CRC cell line genetically led to enhanced CD4+ and CD8+ T cell recruitment resulting in heightened serum INFγ and tumour lysis [158]. Therefore, although considered as part of type 2 immunity, studies are consistent with IL-9 inducing anti-tumour immunity through cross-activating type 1 immune INFγ producing cells such as CD8+ T cells and Th1 cells [155, 158]. Conversely, some studies have found that IL-9 may instead promote tumorigenesis as seen in a heterotypic CT26 implant model of CRC where genetic IL-9 deficiency reduced tumour growth [159]. A potential explanation may be that IL-9 is also a prototypical mast cell growth factor and may exert tumour-promoting properties through mast cells [160]. Mast cells are associated with poor prognosis in a broad range of human cancer types, including melanoma, pancreatic, and prostate cancer [161163], and can promote cancer angiogenesis through tryptase-mediated extracellular matrix remodelling [164]. Therefore, the role of IL-9 in tumorigenesis is likely dependent on the specific immune cell landscape of the tumour microenvironment depending on the cancer type.

    IL-25 and IL-33: novel players of type 2 immunity in tumorigenesis

    IL-25 and IL-33 are well-established type 2 immune-related cytokines with recently discovered roles in cancer immunity. Both are potent inducers of type 2 immunity with some degree of overlap in function, despite belonging to distinct cytokine families [165]. Specifically, IL-33 and IL-25 directly activate ILC2s and stimulate the release of type 2 cytokines, and the differential sensitivities of tissue-specific ILC2s to these cytokines underlie their functional differences and relative importance in inducing type 2 immunity at different organs [166]. For example, ILC2s in the intestines express high levels of IL-25R, while expression of IL-33R component suppression of tumorigenicity 2 (ST2) is minimal [167]. IL-33 also has additional roles in stimulating Tregs and Th17 cells, the latter likely in a mast cell-dependent manner [168171]. These may underlie the differential roles of IL-33 and IL-25 in different cancer types depending on the tissue site.

    The roles of IL-33 and IL-25 in CRC have recently been reviewed [31]. Briefly, IL-25 is produced by rare tuft cells in the intestines and may promote CRC tumorigenesis through directly inducing tumour stemness or indirectly via stimulation of ILC2s which activates MDSCs [24, 172, 173]. Critically, therapeutic blockade of the IL-25-ILC2-MDSC axis increased IFNγ+ expressing CD8+ T cells and Th1 cells and significantly reduced CRC burden. In human CRC patients, increased tumour IL-25 expression is associated with poor prognosis, indicating that blocking IL-25 signalling may be a potential novel therapeutic option. Conversely, others have found that IL-25 treatment can reduce subcutaneous tumour growth across a broad range of implanted cancer cell lines in immunodeficient mice [174], most likely through directly inducing apoptosis as shown in breast cancer cells [175]. Therefore, any therapeutic attempts utilising IL-25 to treat cancer will need to be directed to tumour cells and must be used with caution given the potential to elicit pro-tumoural immune responses in the tumour microenvironment.

    IL-33 may similarly promote or inhibit cancer in a context-dependent manner. Genetic deficiency of ST2 enhanced NK cell effector function resulting in reduced 4T1 breast cancer growth and metastasis in mice [176]. Others have found IL-33 to stimulate ILC2-mediated NK cell suppression in lung cancer [151]. Similarly, IL-33 administration increased tumour-infiltrating IL-13+ ILC2s, MDSCs, and Tregs, and correlated with reduced NK cell cytotoxicity and accelerated tumour growth and metastasis to the lung and liver [177]. Mechanistically, IL-33 can promote metastasis via the induction of desmoplastic reactions by activating cancer-associated fibroblasts [178]. Furthermore, both IL-33 and IL-25 have been shown to promote angiogenesis which also contributes to metastasis, and genetic deletion of ST2 reduced VEGF expression and tumour burden in mouse breast cancer [179181]. IL-33 can also promote tumorigenesis through stimulating mast cells and Tregs, as seen in mouse models of CRC [182, 183].

    Currently, there are no clinical trials related to IL-33 or IL-25 for cancer treatment. Nevertheless, given their overarching effects on tumour immunity, acting upstream of the main effector type 2 cytokines IL-4, IL-5, IL-9, and IL-13 and T cells, the prospect of therapeutically targeting these cytokines is promising. For example, MDSCs have been shown to exert therapeutic resistance across all cancer treatment modalities including surgery, chemotherapy, radiotherapy, and immunotherapy across a broad range of cancer types [184188]. Combination therapies through concomitant IL-25-signalling blockade may effectively deplete MDSCs by inhibiting the IL-25-ILC2-MDSC axis thereby potentiating existing cancer therapies [24].

    Type 2 cytokines in immunotherapy

    Unlike type 1 immune cytokines, there are few clinical trials exploring the therapeutic efficacy of targeting type 2 cytokines. This is likely due to the less consistent roles of type 2 immunity in cancer where they may either promote or suppress tumorigenesis depending on specific contexts. In the case of IL-4, preclinical studies as discussed above indicate that supraphysiologic levels of exogenous IL-4 may induce tumour rejection, while endogenous IL-4 typically promotes tumour progression [129133]. Therefore, potential strategies for cancer treatment may involve injecting high levels of IL-4 or blocking existing IL-4-signalling in patients. Human cancer studies have found IL-4R expression to be elevated in many tumours types such as bladder, lung, breast, liver, and prostate cancer [135, 189, 190], suggesting that IL-4-signalling can be selectively targeted therapeutically for cancer treatment. A phase I clinical trial of 17 patients with solid tumours found that IL-4 treatment is limited by toxicities, with symptoms including weight gain, effusions, rash, peripheral oedema, and oliguria [191]. Importantly, whether the maximally tolerated dose of IL-4 injected in patients is sufficient to elicit anti-tumour properties is unclear. In a phase II trial of advanced renal cell carcinoma where 49 patients were treated with subcutaneous injection of IL-4 at 5 mcg/kg per day for 28 days, no complete or partial responses were observed [192]. In a similar phase II trial by the same group but instead in melanoma patients, only 1 patient responded to IL-4 therapy out of the 34 treated patients [193]. Another phase II study also concluded no benefit of IL-4 treatment at the maximum tolerated dose for melanoma or renal cell carcinoma [194]. Currently, there are no ongoing oncology clinical trials involving IL-4. Future clinical trials involving IL-4 administration will have to overcome the hurdle of therapeutic toxicities in order to achieve a clinically effective dose, or alternatively through monoclonal antibodies neutralising endogenous pro-tumoural IL-4 or blocking the IL-4R.

    Similar to IL-4, currently there are no ongoing clinical trials involving IL-5 or IL-9 in cancer treatment, likely a reflection of our current inadequate understanding of whether they would promote or inhibit cancer under different contexts. Furthermore, there are few studies that directly assessed the expression of IL-5, IL-9, or their receptors in human cancer. IL-9-producing Tregs have been detected in human non-small cell lung cancer samples, while others have found that IL-9R expression is significantly enhanced in endometrial cancer compared to other cancer types such as renal or breast cancers [195, 196]. Conversely, increased breast cancer IL-5 expression is associated with metastasis and poor prognosis [197], while others have found that tumour Th2 cell-derived IL-5 may instead enhance response to immune checkpoint inhibitors in breast cancer [198]. Future trials targeting these cytokines should therefore focus on cancer types where these cytokines and respective receptors are readily detected. On the other hand, studies consistently showed IL-13 to exert pro-tumoural properties, and in humans, particularly, through the alternative IL-13Rα2 [138, 140]. Inhibiting IL-13 has many theoretical benefits, including reducing MDSCs and M2 TAMs, while promoting anti-tumour IFNγ+ and T cell infiltration [24, 59, 134]. This can be achieved through monoclonal antibody-mediated neutralisation of IL-13 or blocking the IL-13R, or alternatively through recently developed IL-13Rα2-targeting chimeric antigen receptor (CAR)-T cells [199]. CARs are genetically engineered receptors that redirect T cells to recognise and eliminate cells expressing the target antigen, in this case, IL-13Rα2. In gliomas, IL-13Rα2 expression is only detected in tumours and not normal brain tissue [200]. Currently, there are six ongoing clinical trials utilising IL-13Rα2-targeting CAR-T cells, either alone or in combination with chemotherapy or checkpoint inhibitor immunotherapy (Table 3). Chemotherapy can induce antigen release while checkpoint inhibitors revitalise exhausted T cells [201], theoretically potentiating CAR-T cell therapy. Furthermore, in one study, the IL-2 cytokine is also added which has a major role in inducing T cell proliferation [202] and may further enhance IL-13Rα2-targeting CAR-T cell efficacy. In addition to directly killing tumour cells, IL-13Rα2-targeting CAR-T cells may eliminate other IL-13-responsive cells in the tumour microenvironment contributing to therapeutic efficacy.

    A comprehensive list of ongoing oncology clinical trials involving IL-13

    CytokineClinical trial IDPhaseNumber of patientsCancer typeTreatmentStatusEstimated study completion date
    IL-13NCT02208362I82Recurrent or refractory gliomaIL-13Rα2-specific Hinge-optimized 4-1BB-co-stimulatory CAR/truncated CD19-expressing autologous T cellsActiveJune 2023
    IL-13NCT04119024I24MelanomaIL-13Rα2-specific Hinge-optimized 4-1BB-co-stimulatory CAR/truncated CD19-expressing autologous naive and memory T cells plus cyclophosphamide, fludarabine, and IL-2RecruitingOctober 2025
    IL-13NCT04661384I30Ependymoma, glioblastoma, medulloblastoma, leptomeninges cancerIL-13Rα2-specific Hinge-optimized 4-1BB-co-stimulatory CAR truncated CD19-expressing autologous T-lymphocytesRecruitingNovember 2025
    IL-13NCT04510051I18Brain cancerIL-13Rα2-specific Hinge-optimized 4-1BB-co-stimulatory CAR truncated CD19-expressing autologous T-lymphocytes plus cyclophosphamide and fludarabineRecruiting September 2023
    IL-13NCT04003649I60Recurrent or refractory glioblastomaIL-13Rα2-specific Hinge-optimized 4-1BB-co-stimulatory CAR/truncated CD19-expressing autologous naive and memory T cells plus ipilimumab and nivolumabRecruitingDecember 2024
    IL-13NCT05168423I18GlioblastomaCAR-T-EGFR-IL-13Rα2 plus cyclophosphamide and fludarabineNot yet recruitingDecember 2039
    Display full size

    EGFR: epidermal growth factor receptor

    Conclusions

    Cytokines play overarching roles in shaping tumour immunity with broad influence on the tumour microenvironment and are attractive targets for cancer immunotherapy. Understanding how each cytokine reacts and influences the tumour microenvironment of different cancer types is critical in order to achieve beneficial therapeutic outcomes, especially as the same cytokine may have entirely opposing functions in different cancers. Overall, exogenous administration of type 1 immune cytokines for cancer treatments, particularly IFNγ and IL-12, are promising but may be limited by toxicities. Ongoing clinical trials will continue to investigate different modalities of administration to minimise iatrogenic harm while maximising treatment efficacy. A better contextual understanding of type 2 immune cytokines, in particular IL-5 and IL-9, will open new avenues for treatment testing in clinical trials. The roles of cytokines in cancer pathways are continuingly being discovered and will have great impacts on future cancer therapy.

    Abbreviations

    CAR:

    chimeric antigen receptor

    CRC:

    colorectal cancer

    DCs:

    dendritic cells

    IFNGR:

    interferon γ receptor

    IFNγ:

    interferon γ

    IL-12:

    interleukin-12

    IL-12R:

    interleukin-12 receptor

    ILC2s:

    group 2 innate lymphoid cells

    ISG:

    interferon-stimulated gene

    JAK1:

    Janus kinase 1

    MDSCs:

    myeloid-derived suppressor cells

    NK:

    natural killer

    pAPCs:

    professional antigen-presenting cells

    PD-L1:

    programmed death-ligand 1

    ST2:

    suppression of tumorigenicity 2

    STAT1:

    signal transducer and activator of transcription 1

    TAM:

    tumour-associated macrophage

    TGFβ:

    transforming growth factor β

    Th1:

    T helper 1

    TNFR2:

    tumour necrosis factor receptor type 2

    TNFα:

    tumour necrosis factor α

    Treg:

    regulatory T

    Declarations

    Author contributions

    EJ: Conceptualization, Methodology, Writing—original draft, Writing—review & editing.

    Conflicts of interest

    The author declares that he has no conflicts of interest.

    Ethical approval

    Not applicable.

    Consent to participate

    Not applicable.

    Consent to publication

    Not applicable.

    Availability of materials and data

    Not applicable.

    Funding

    Not applicable.

    Copyright

    © The Author(s) 2023.

    References

    GBD 2015 Mortality and Causes of Death Collaborators. Global, regional, and national life expectancy, all-cause mortality, and cause-specific mortality for 249 causes of death, 1980–2015: a systematic analysis for the Global Burden of Disease Study 2015. Lancet. 2016;388:1459544. Erratum in: Lancet. 2017;389:E1. [DOI] [PubMed] [PMC]
    Sung H, Ferlay J, Siegel RL, Laversanne M, Soerjomataram I, Jemal A, et al. Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J Clin. 2021;71:20949. [DOI] [PubMed]
    Jenkins RW, Barbie DA, Flaherty KT. Mechanisms of resistance to immune checkpoint inhibitors. Br J Cancer. 2018;118:916. [DOI] [PubMed] [PMC]
    Hunter P. The fourth pillar. EMBO Rep. 2017;18:188992. [DOI]
    Robert C. A decade of immune-checkpoint inhibitors in cancer therapy. Nat Commun. 2020;11:3801. [DOI] [PubMed] [PMC]
    Stenken JA, Poschenrieder AJ. Bioanalytical chemistry of cytokines–a review. Anal Chim Acta. 2015;853:95115. [DOI] [PubMed] [PMC]
    Annunziato F, Romagnani C, Romagnani S. The 3 major types of innate and adaptive cell-mediated effector immunity. J Allergy Clin Immunol. 2015;135:62635. [DOI] [PubMed]
    Walker JA, McKenzie ANJ. Development and function of group 2 innate lymphoid cells. Curr Opin Immunol. 2013;25:14855. [DOI] [PubMed] [PMC]
    Ribatti D. The concept of immune surveillance against tumors: the first theories. Oncotarget. 2017;8:717580. [DOI] [PubMed] [PMC]
    Kartikasari AER, Huertas CS, Mitchell A, Plebanski M. Tumor-induced inflammatory cytokines and the emerging diagnostic devices for cancer detection and prognosis. Front Oncol. 2021;11:692142. [DOI] [PubMed] [PMC]
    Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. Cell. 2011;144:64674. [DOI] [PubMed]
    Muhammad Yusoff F, Wong KK, Mohd Redzwan N. Th1, Th2, and Th17 cytokines in systemic lupus erythematosus. Autoimmunity. 2020;53:820. [DOI] [PubMed]
    Walker JA, Barlow JL, McKenzie AN. Innate lymphoid cells — how did we miss them? Nat Rev Immunol. 2013;13:7587. [DOI] [PubMed]
    Lucey DR, Clerici M, Shearer GM. Type 1 and type 2 cytokine dysregulation in human infectious, neoplastic, and inflammatory diseases. Clin Microbiol Rev. 1996;9:53262. [DOI] [PubMed] [PMC]
    Rodriguez-Rodriguez N, Gogoi M, McKenzie ANJ. Group 2 innate lymphoid cells: team players in regulating asthma. Annu Rev Immunol. 2021;39:16798. [DOI] [PubMed] [PMC]
    Gieseck RL 3rd, Wilson MS, Wynn TA. Type 2 immunity in tissue repair and fibrosis. Nat Rev Immunol. 2018;18:6276. [DOI] [PubMed]
    Ellyard JI, Simson L, Parish CR. Th2-mediated anti-tumour immunity: friend or foe? Tissue Antigens. 2007;70:111. [DOI] [PubMed]
    Ikeda H, Chamoto K, Tsuji T, Suzuki Y, Wakita D, Takeshima T, et al. The critical role of type-1 innate and acquired immunity in tumor immunotherapy. Cancer Sci. 2004;95:697703. [DOI] [PubMed]
    Disis ML. Immune regulation of cancer. J Clin Oncol. 2010;28:45318. [DOI] [PubMed] [PMC]
    Pagès F, Kirilovsky A, Mlecnik B, Asslaber M, Tosolini M, Bindea G, et al. In situ cytotoxic and memory T cells predict outcome in patients with early-stage colorectal cancer. J Clin Oncol. 2009;27:594451. [DOI] [PubMed]
    Galon J, Costes A, Sanchez-Cabo F, Kirilovsky A, Mlecnik B, Lagorce-Pagès C, et al. Type, density, and location of immune cells within human colorectal tumors predict clinical outcome. Science. 2006;313:19604. [DOI] [PubMed]
    Mlecnik B, Tosolini M, Kirilovsky A, Berger A, Bindea G, Meatchi T, et al. Histopathologic-based prognostic factors of colorectal cancers are associated with the state of the local immune reaction. J Clin Oncol. 2011;29:6108. [DOI] [PubMed]
    Pagès F, Galon J, Dieu-Nosjean MC, Tartour E, Sautès-Fridman C, Fridman WH. Immune infiltration in human tumors: a prognostic factor that should not be ignored. Oncogene. 2010;29:1093102. [DOI]
    Jou E, Rodriguez-Rodriguez N, Ferreira AF, Jolin HE, Clark PA, Sawmynaden K, et al. An innate IL-25–ILC2–MDSC axis creates a cancer-permissive microenvironment for Apc mutation–driven intestinal tumorigenesis. Sci Immunol. 2022;7:eabn0175. [DOI] [PubMed] [PMC]
    Grivennikov SI, Greten FR, Karin M. Immunity, inflammation, and cancer. Cell. 2010;140:88399. [DOI] [PubMed] [PMC]
    Liu M, Kuo F, Capistrano KJ, Kang D, Nixon BG, Shi W, et al. TGF-β suppresses type 2 immunity to cancer. Nature. 2020;587:11520. [DOI] [PubMed] [PMC]
    Lorvik KB, Hammarström C, Fauskanger M, Haabeth OA, Zangani M, Haraldsen G, et al. Adoptive transfer of tumor-specific Th2 cells eradicates tumors by triggering an in situ inflammatory immune response. Cancer Res. 2016;76:686476. [DOI] [PubMed]
    Mojic M, Takeda K, Hayakawa Y. The dark side of IFN-γ: its role in promoting cancer immunoevasion. Int J Mol Sci. 2017;19:89. [DOI] [PubMed] [PMC]
    Singh N, Baby D, Rajguru JP, Patil PB, Thakkannavar SS, Pujari VB. Inflammation and cancer. Ann Afr Med. 2019;18:1216. [DOI] [PubMed] [PMC]
    Fournié JJ, Poupot M. The pro-tumorigenic IL-33 involved in antitumor immunity: a yin and yang cytokine. Front Immunol. 2018;9:2506. [DOI] [PubMed] [PMC]
    Jou E, Rodriguez-Rodriguez N, McKenzie ANJ. Emerging roles for IL-25 and IL-33 in colorectal cancer tumorigenesis. Front Immunol. 2022;13:981479. [DOI] [PubMed] [PMC]
    Shen J, Xiao Z, Zhao Q, Li M, Wu X, Zhang L, et al. Anti-cancer therapy with TNFα and IFNγ: a comprehensive review. Cell Prolif. 2018;51:e12441. [DOI] [PubMed] [PMC]
    Castro F, Cardoso AP, Gonçalves RM, Serre K, Oliveira MJ. Interferon-gamma at the crossroads of tumor immune surveillance or evasion. Front Immunol. 2018;9:847. [DOI] [PubMed] [PMC]
    Djuretic IM, Levanon D, Negreanu V, Groner Y, Rao A, Ansel KM. Transcription factors T-bet and Runx3 cooperate to activate Ifng and silence Il4 in T helper type 1 cells. Nat Immunol. 2007;8:14553. [DOI] [PubMed]
    Prochazkova J, Pokorna K, Holan V. IL-12 inhibits the TGF-β-dependent T cell developmental programs and skews the TGF-β-induced differentiation into a Th1-like direction. Immunobiology. 2012;217:7482. [DOI] [PubMed]
    Wang X, Lin Y. Tumor necrosis factor and cancer, buddies or foes? Acta Pharmacol Sin. 2008;29:127588. [DOI] [PubMed] [PMC]
    Wack A, Terczyńska-Dyla E, Hartmann R. Guarding the frontiers: the biology of type III interferons. Nat Immunol. 2015;16:8029. [DOI] [PubMed] [PMC]
    Shankaran V, Ikeda H, Bruce AT, White JM, Swanson PE, Old LJ, et al. IFNγ and lymphocytes prevent primary tumour development and shape tumour immunogenicity. Nature. 2001;410:110711. [DOI]
    Kaplan DH, Shankaran V, Dighe AS, Stockert E, Aguet M, Old LJ, et al. Demonstration of an interferon γ-dependent tumor surveillance system in immunocompetent mice. Proc Natl Acad Sci U S A. 1998;95:755661. [DOI] [PubMed] [PMC]
    Street SE, Trapani JA, MacGregor D, Smyth MJ. Suppression of lymphoma and epithelial malignancies effected by interferon γ. J Exp Med. 2002;196:12934. [DOI] [PubMed] [PMC]
    Wang L, Wang Y, Song Z, Chu J, Qu X. Deficiency of interferon-gamma or its receptor promotes colorectal cancer development. J Interferon Cytokine Res. 2015;35:27380. [DOI] [PubMed]
    Erick TK, Brossay L. Phenotype and functions of conventional and non-conventional NK cells. Curr Opin Immunol. 2016;38:6774. [DOI] [PubMed] [PMC]
    Bhat P, Leggatt G, Waterhouse N, Frazer IH. Interferon-γ derived from cytotoxic lymphocytes directly enhances their motility and cytotoxicity. Cell Death Dis. 2017;8:e2836. [DOI] [PubMed] [PMC]
    Weizman OE, Adams NM, Schuster IS, Krishna C, Pritykin Y, Lau C, et al. ILC1 confer early host protection at initial sites of viral infection. Cell. 2017;171:795808.E12. [DOI] [PubMed] [PMC]
    Ivashkiv LB. IFNγ: signalling, epigenetics and roles in immunity, metabolism, disease and cancer immunotherapy. Nat Rev Immunol. 2018;18:54558. [DOI] [PubMed] [PMC]
    Kotredes KP, Gamero AM. Interferons as inducers of apoptosis in malignant cells. J Interferon Cytokine Res. 2013;33:16270. [DOI] [PubMed] [PMC]
    Su Q, Wang F, Dong Z, Chen M, Cao R. IFN‑γ induces apoptosis in human melanocytes by activating the JAK1/STAT1 signaling pathway. Mol Med Rep. 2020;22:31116. [DOI] [PubMed] [PMC]
    Rakshit S, Chandrasekar BS, Saha B, Victor ES, Majumdar S, Nandi D. Interferon-gamma induced cell death: regulation and contributions of nitric oxide, cJun N-terminal kinase, reactive oxygen species and peroxynitrite. Biochim Biophys Acta. 2014;1843:264561. [DOI] [PubMed]
    Su X, Yu Y, Zhong Y, Giannopoulou EG, Hu X, Liu H, et al. Interferon-γ regulates cellular metabolism and mRNA translation to potentiate macrophage activation. Nat Immunol. 2015;16:83849. [DOI] [PubMed] [PMC]
    Zhang C, Hou D, Wei H, Zhao M, Yang L, Liu Q, et al. Lack of interferon-γ receptor results in a microenvironment favorable for intestinal tumorigenesis. Oncotarget. 2016;7:42099109. [DOI] [PubMed] [PMC]
    Pagès F, Berger A, Camus M, Sanchez-Cabo F, Costes A, Molidor R, et al. Effector memory T cells, early metastasis, and survival in colorectal cancer. N Engl J Med. 2005;353:265466. [DOI] [PubMed]
    Heimes AS, Härtner F, Almstedt K, Krajnak S, Lebrecht A, Battista MJ, et al. Prognostic significance of interferon-γ and its signaling pathway in early breast cancer depends on the molecular subtypes. Int J Mol Sci. 2020;21:7178. [DOI] [PubMed] [PMC]
    Lee IC, Huang YH, Chau GY, Huo TI, Su CW, Wu JC, et al. Serum interferon gamma level predicts recurrence in hepatocellular carcinoma patients after curative treatments. Int J Cancer. 2013;133:2895902. [DOI] [PubMed]
    Duncan TJ, Rolland P, Deen S, Scott IV, Liu DT, Spendlove I, et al. Loss of IFNγ receptor is an independent prognostic factor in ovarian cancer. Clin Cancer Res. 2007;13:413945. [DOI]
    Atri C, Guerfali FZ, Laouini D. Role of human macrophage polarization in inflammation during infectious diseases. Int J Mol Sci. 2018;19:1801. [DOI] [PubMed] [PMC]
    Pan Y, Yu Y, Wang X, Zhang T. Tumor-associated macrophages in tumor immunity. Front Immunol. 2020;11:583084. Erratum in: Front Immunol. 2021;12:775758. [DOI] [PubMed] [PMC]
    Mantovani A, Sica A, Sozzani S, Allavena P, Vecchi A, Locati M. The chemokine system in diverse forms of macrophage activation and polarization. Trends Immunol. 2004;25:67786. [DOI] [PubMed]
    Wang N, Liang H, Zen K. Molecular mechanisms that influence the macrophage M1–M2 polarization balance. Front Immunol. 2014;5:614. [DOI] [PubMed] [PMC]
    Nakanishi Y, Nakatsuji M, Seno H, Ishizu S, Akitake-Kawano R, Kanda K, et al. COX-2 inhibition alters the phenotype of tumor-associated macrophages from M2 to M1 in Apc Min/+ mouse polyps. Carcinogenesis. 2011;32:13339. [DOI] [PubMed]
    Zhang M, He Y, Sun X, Li Q, Wang W, Zhao A, et al. A high M1/M2 ratio of tumor-associated macrophages is associated with extended survival in ovarian cancer patients. J Ovarian Res. 2014;7:19. [DOI] [PubMed] [PMC]
    Jackute J, Zemaitis M, Pranys D, Sitkauskiene B, Miliauskas S, Vaitkiene S, et al. Distribution of M1 and M2 macrophages in tumor islets and stroma in relation to prognosis of non-small cell lung cancer. BMC Immunol. 2018;19:3. [DOI] [PubMed] [PMC]
    Dong P, Ma L, Liu L, Zhao G, Zhang S, Dong L, et al. CD86⁺/CD206⁺, diametrically polarized tumor-associated macrophages, predict hepatocellular carcinoma patient prognosis. Int J Mol Sci. 2016;17:320. [DOI] [PubMed] [PMC]
    Steimle V, Siegrist CA, Mottet A, Lisowska-Grospierre B, Mach B. Regulation of MHC class II expression by interferon-γ mediated by the transactivator gene CIITA. Science. 1994;265:1069. [DOI]
    Biswas SK, Mantovani A. Macrophage plasticity and interaction with lymphocyte subsets: cancer as a paradigm. Nat Immunol. 2010;11:88996. [DOI] [PubMed]
    Ylikoski E, Lund R, Kyläniemi M, Filén S, Kilpeläinen M, Savolainen J, et al. IL-12 up-regulates T-bet independently of IFN-γ in human CD4+ T cells. Eur J Immunol. 2005;35:3297306. Erratum in: Eur J Immunol. 2006;36:1058. [DOI] [PubMed]
    Eickhoff S, Brewitz A, Gerner MY, Klauschen F, Komander K, Hemmi H, et al. Robust anti-viral immunity requires multiple distinct T cell-dendritic cell interactions. Cell. 2015;162:132237. [DOI] [PubMed] [PMC]
    Bachem A, Hartung E, Güttler S, Mora A, Zhou X, Hegemann A, et al. Expression of XCR1 characterizes the Batf3-dependent lineage of dendritic cells capable of antigen cross-presentation. Front Immunol. 2012;3:214. [DOI] [PubMed] [PMC]
    Martínez-Lostao L, Anel A, Pardo J. How do cytotoxic lymphocytes kill cancer cells? Clin Cancer Res. 2015;21:504756. [DOI] [PubMed]
    McLane LM, Abdel-Hakeem MS, Wherry EJ. CD8 T cell exhaustion during chronic viral infection and cancer. Annu Rev Immunol. 2019;37:45795. [DOI] [PubMed]
    Raskov H, Orhan A, Christensen JP, Gögenur I. Cytotoxic CD8+ T cells in cancer and cancer immunotherapy. Br J Cancer. 2021;124:35967. [DOI] [PubMed] [PMC]
    Zhuang X, Shi G, Hu X, Wang H, Sun W, Wu Y. Interferon-gamma inhibits aldehyde dehydrogenasebright cancer stem cells in the 4T1 mouse model of breast cancer. Chin Med J (Engl). 2021;135:194204. [DOI] [PubMed] [PMC]
    Brenner E, Schörg BF, Ahmetlić F, Wieder T, Hilke FJ, Simon N, et al. Cancer immune control needs senescence induction by interferon-dependent cell cycle regulator pathways in tumours. Nat Commun. 2020;11:1335. [DOI] [PubMed] [PMC]
    Sun T, Yang Y, Luo X, Cheng Y, Zhang M, Wang K, et al. Inhibition of tumor angiogenesis by interferon-γ by suppression of tumor-associated macrophage differentiation. Oncol Res. 2014;21:22735. [DOI] [PubMed]
    Vignali DAA, Kuchroo VK. IL-12 family cytokines: immunological playmakers. Nat Immunol. 2012;13:7228. [DOI] [PubMed] [PMC]
    Lu Z, Xu K, Wang X, Li Y, Li M. Interleukin 39: a new member of interleukin 12 family. Cent Eur J Immunol. 2020;45:2147. [DOI] [PubMed] [PMC]
    Toussirot E. The IL23/Th17 pathway as a therapeutic target in chronic inflammatory diseases. Inflamm Allergy Drug Targets. 2012;11:15968. [DOI] [PubMed]
    Liu K, Huang A, Nie J, Tan J, Xing S, Qu Y, et al. IL-35 regulates the function of immune cells in tumor microenvironment. Front Immunol. 2021;12:683332. [DOI] [PubMed] [PMC]
    Zheng H, Ban Y, Wei F, Ma X. Regulation of interleukin-12 production in antigen-presenting cells. Adv Exp Med Biol. 2016;941:11738. [DOI] [PubMed]
    Lasek W, Zagożdżon R, Jakobisiak M. Interleukin 12: still a promising candidate for tumor immunotherapy? Cancer Immunol Immunother. 2014;63:41935. [DOI] [PubMed] [PMC]
    Grohmann U, Belladonna ML, Vacca C, Bianchi R, Fallarino F, Orabona C, et al. Positive regulatory role of IL-12 in macrophages and modulation by IFN-γ1. J Immunol. 2001;167:2217. [DOI] [PubMed]
    Meeran SM, Mantena SK, Meleth S, Elmets CA, Katiyar SK. Interleukin-12-deficient mice are at greater risk of UV radiation–induced skin tumors and malignant transformation of papillomas to carcinomas. Mol Cancer Ther. 2006;5:82532. [DOI] [PubMed]
    Smyth MJ, Taniguchi M, Street SE. The anti-tumor activity of IL-12: mechanisms of innate immunity that are model and dose dependent1. J Immunol. 2000;165:266570. [DOI] [PubMed]
    Cohen CA, Shea AA, Heffron CL, Schmelz EM, Roberts PC. Interleukin-12 immunomodulation delays the onset of lethal peritoneal disease of ovarian cancer. J Interferon Cytokine Res. 2016;36:6273. [DOI] [PubMed] [PMC]
    Zilocchi C, Stoppacciaro A, Chiodoni C, Parenza M, Terrazzini N, Colombo MP. Interferon γ-independent rejection of interleukin 12-transduced carcinoma cells requires CD4+ T cells and granulocyte/macrophage colony-stimulating factor. J Exp Med. 1998;188:13343. [DOI] [PubMed] [PMC]
    Kaarvatn MH, Vrbanec J, Kulic A, Knezevic J, Petricevic B, Balen S, et al. Single nucleotide polymorphism in the interleukin 12B gene is associated with risk for breast cancer development. Scand J Immunol. 2012;76:32935. [DOI] [PubMed]
    Chen X, Han S, Wang S, Zhou X, Zhang M, Dong J, et al. Interactions of IL-12A and IL-12B polymorphisms on the risk of cervical cancer in Chinese women. Clin Cancer Res. 2009;15:4005. [DOI] [PubMed]
    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;72:366670. [DOI] [PubMed] [PMC]
    Kearney CJ, Vervoort SJ, Hogg SJ, Ramsbottom KM, Freeman AJ, Lalaoui N, et al. Tumor immune evasion arises through loss of TNF sensitivity. Sci Immunol. 2018;3:eaar3451. [DOI] [PubMed]
    Rath PC, Aggarwal BB. TNF-induced signaling in apoptosis. J Clin Immunol. 1999;19:35064. [DOI] [PubMed]
    Idriss HT, Naismith JH. TNFα and the TNF receptor superfamily: structure-function relationship(s). Microsc Res Tech. 2000;50:18495. [DOI] [PubMed]
    Moore RJ, Owens DM, Stamp G, Arnott C, Burke F, East N, et al. Mice deficient in tumor necrosis factor-α are resistant to skin carcinogenesis. Nat Med. 1999;5:82831. Erratum in: Nat Med. 1999;5:1087. [DOI] [PubMed]
    Mercogliano MF, Bruni S, Elizalde PV, Schillaci R. Tumor necrosis factor α blockade: an opportunity to tackle breast cancer. Front Oncol. 2020;10:584. [DOI] [PubMed] [PMC]
    Naylor MS, Stamp GW, Foulkes WD, Eccles D, Balkwill FR. Tumor necrosis factor and its receptors in human ovarian cancer. Potential role in disease progression. J Clin Invest. 1993;91:2194206. [DOI] [PubMed] [PMC]
    Liu W, Lu X, Shi P, Yang G, Zhou Z, Li W, et al. TNF-α increases breast cancer stem-like cells through up-regulating TAZ expression via the non-canonical NF-κB pathway. Sci Rep. 2020;10:1804. [DOI] [PubMed] [PMC]
    Storci G, Sansone P, Mari S, D’Uva G, Tavolari S, Guarnieri T, et al. TNFalpha up-regulates SLUG via the NF-kappaB/HIF1alpha axis, which imparts breast cancer cells with a stem cell-like phenotype. J Cell Physiol. 2010;225:68291. [DOI] [PubMed] [PMC]
    Landsberg J, Kohlmeyer J, Renn M, Bald T, Rogava M, Cron M, et al. Melanomas resist T-cell therapy through inflammation-induced reversible dedifferentiation. Nature. 2012;490:4126. [DOI] [PubMed]
    Salomon BL, Leclerc M, Tosello J, Ronin E, Piaggio E, Cohen JL. Tumor necrosis factor α and regulatory T cells in oncoimmunology. Front Immunol. 2018;9:444. [DOI] [PubMed] [PMC]
    Yao B, Wang L, Wang H, Bao J, Li Q, Yu F, et al. Seven interferon gamma response genes serve as a prognostic risk signature that correlates with immune infiltration in lung adenocarcinoma. Aging (Albany NY). 2021;13:11381410. [DOI] [PubMed] [PMC]
    Tao Y, Tao T, Gross N, Peng X, Li Y, Huang Z, et al. Combined effect of IL-12Rβ2 and IL-23R expression on prognosis of patients with laryngeal cancer. Cell Physiol Biochem. 2018;50:104154. [DOI] [PubMed]
    Xu L, Pelosof L, Wang R, McFarland HI, Wu WW, Phue JN, et al. NGS evaluation of colorectal cancer reveals interferon gamma dependent expression of immune checkpoint genes and identification of novel IFNγ induced genes. Front Immunol. 2020;11:224. [DOI] [PubMed] [PMC]
    Korentzelos D, Wells A, Clark AM. Interferon-γ increases sensitivity to chemotherapy and provides immunotherapy targets in models of metastatic castration-resistant prostate cancer. Scientific Reports. 2022;12:6657. [DOI] [PubMed] [PMC]
    Bayat Mokhtari R, Homayouni TS, Baluch N, Morgatskaya E, Kumar S, Das B, et al. Combination therapy in combating cancer. Oncotarget. 2017;8:3802243. [DOI] [PubMed] [PMC]
    Windbichler GH, Hausmaninger H, Stummvoll W, Graf AH, Kainz C, Lahodny J, et al. Interferon-gamma in the first-line therapy of ovarian cancer: a randomized phase III trial. Br J Cancer. 2000;82:113844. [DOI] [PubMed] [PMC]
    Schmeler KM, Vadhan-Raj S, Ramirez PT, Apte SM, Cohen L, Bassett RL, et al. A phase II study of GM-CSF and rIFN-γ1b plus carboplatin for the treatment of recurrent, platinum-sensitive ovarian, fallopian tube and primary peritoneal cancer. Gynecol Oncol. 2009;113:2105. [DOI] [PubMed] [PMC]
    Jorgovanovic D, Song M, Wang L, Zhang Y. Roles of IFN-γ in tumor progression and regression: a review. Biomark Res. 2020;8:49. [DOI] [PubMed] [PMC]
    Schroeder BA, Black RG, Spadinger S, Zhang S, Kohli K, Cao J, et al. Histiocyte predominant myocarditis resulting from the addition of interferon gamma to cyclophosphamide-based lymphodepletion for adoptive cellular therapy. J Immunother Cancer. 2020;8:e000247. [DOI] [PubMed] [PMC]
    Grasso CS, Tsoi J, Onyshchenko M, Abril-Rodriguez G, Ross-Macdonald P, Wind-Rotolo M, et al. Conserved interferon-γ signaling drives clinical response to immune checkpoint blockade therapy in melanoma. Cancer Cell. 2020;38:50015.E3. Erratum in: Cancer Cell. 2021;39:122. [DOI] [PubMed] [PMC]
    Mirlekar B, Pylayeva-Gupta Y. IL-12 family cytokines in cancer and immunotherapy. Cancers (Basel). 2021;13:167. [DOI] [PubMed] [PMC]
    Leonard JP, Sherman ML, Fisher GL, Buchanan LJ, Larsen G, Atkins MB, et al. Effects of single-dose interleukin-12 exposure on interleukin-12–associated toxicity and interferon-γ production. Blood. 1997;90:25418. [PubMed]
    Zhang L, Morgan RA, Beane JD, Zheng Z, Dudley ME, Kassim SH, et al. Tumor-infiltrating lymphocytes genetically engineered with an inducible gene encoding interleukin-12 for the immunotherapy of metastatic melanoma. Clin Cancer Res. 2015;21:227888. [DOI] [PubMed] [PMC]
    Li J, Lin W, Chen H, Xu Z, Ye Y, Chen M. Dual-target IL-12-containing nanoparticles enhance T cell functions for cancer immunotherapy. Cell Immunol. 2020;349:104042. [DOI] [PubMed]
    Wang P, Li X, Wang J, Gao D, Li Y, Li H, et al. Re-designing interleukin-12 to enhance its safety and potential as an anti-tumor immunotherapeutic agent. Nat Commun. 2017;8:1395. Erratum in: Nat Commun. 2018;9:203. [DOI] [PubMed] [PMC]
    Tucker CG, Mitchell JS, Martinov T, Burbach BJ, Beura LK, Wilson JC, et al. Adoptive T cell therapy with IL-12-preconditioned low-avidity T cells prevents exhaustion and results in enhanced T cell activation, enhanced tumor clearance, and decreased risk for autoimmunity. J Immunol. 2020;205:144960. [DOI] [PubMed] [PMC]
    Donia M, Hansen M, Sendrup SL, Iversen TZ, Ellebæk E, Andersen MH, et al. Methods to improve adoptive T-cell therapy for melanoma: IFN-γ enhances anticancer responses of cell products for infusion. J Invest Dermatol. 2013;133:54552. [DOI] [PubMed]
    Greiner JW, Morillon YM 2nd, Schlom J. NHS-IL12, a tumor-targeting immunocytokine. Immunotargets Ther. 2021;10:15569. [DOI] [PubMed] [PMC]
    Wu X, Wu MY, Jiang M, Zhi Q, Bian X, Xu MD, et al. TNF-α sensitizes chemotherapy and radiotherapy against breast cancer cells. Cancer Cell Int. 2017;17:13. [DOI] [PubMed] [PMC]
    Roberts NJ, Zhou S, Diaz LA Jr, Holdhoff M. Systemic use of tumor necrosis factor alpha as an anticancer agent. Oncotarget. 2011;2:73951. [DOI] [PubMed] [PMC]
    Creaven PJ, Plager JE, Dupere S, Huben RP, Takita H, Mittelman A, et al. Phase I clinical trial of recombinant human tumor necrosis factor. Cancer Chemother Pharmacol. 1987;20:13744. [DOI] [PubMed]
    Budd GT, Green S, Baker LH, Hersh EP, Weick JK, Osborne CK. A Southwest Oncology Group phase II trial of recombinant tumor necrosis factor in metastatic breast cancer. Cancer. 1991;68:16945. Erratum in: Cancer. 1992;69:2866. [DOI] [PubMed]
    Gasteiger G, Fan X, Dikiy S, Lee SY, Rudensky AY. Tissue residency of innate lymphoid cells in lymphoid and nonlymphoid organs. Science. 2015;350:9815. [DOI] [PubMed] [PMC]
    Neill DR, Wong SH, Bellosi A, Flynn RJ, Daly M, Langford TKA, et al. Nuocytes represent a new innate effector leukocyte that mediates type-2 immunity. Nature. 2010;464:136770. [DOI] [PubMed] [PMC]
    Oliphant CJ, Hwang YY, Walker JA, Salimi M, Wong SH, Brewer JM, et al. MHCII-mediated dialog between group 2 innate lymphoid cells and CD4+ T cells potentiates type 2 immunity and promotes parasitic helminth expulsion. Immunity. 2014;41:28395. [DOI] [PubMed] [PMC]
    McLeod JJ, Baker B, Ryan JJ. Mast cell production and response to IL-4 and IL-13. Cytokine. 2015;75:5761. [DOI] [PubMed] [PMC]
    Doran E, Cai F, Holweg CTJ, Wong K, Brumm J, Arron JR. Interleukin-13 in asthma and other eosinophilic disorders. Front Med (Lausanne). 2017;4:139. [DOI] [PubMed] [PMC]
    Piehler D, Stenzel W, Grahnert A, Held J, Richter L, Köhler G, et al. Eosinophils contribute to IL-4 production and shape the T-helper cytokine profile and inflammatory response in pulmonary cryptococcosis. Am J Pathol. 2011;179:73344. [DOI] [PubMed] [PMC]
    Veldhoen M, Uyttenhove C, van Snick J, Helmby H, Westendorf A, Buer J, et al. Transforming growth factor-β ‘reprograms’ the differentiation of T helper 2 cells and promotes an interleukin 9-producing subset. Nat Immunol. 2008;9:13416. [DOI] [PubMed]
    Forbes E, van Panhuys N, Min B, Le Gros G. Differential requirements for IL-4/STAT6 signalling in CD4 T-cell fate determination and Th2-immune effector responses. Immunol Cell Biol. 2010;88:2403. [DOI] [PubMed]
    Junttila IS. Tuning the cytokine responses: an update on interleukin (IL)-4 and IL-13 receptor complexes. Front Immunol. 2018;9:888. [DOI] [PubMed] [PMC]
    Golumbek PT, Lazenby AJ, Levitsky HI, Jaffee LM, Karasuyama H, Baker M, et al. Treatment of established renal cancer by tumor cells engineered to secrete interleukin-4. Science. 1991;254:7136. [DOI] [PubMed]
    Tepper RI, Pattengale PK, Leder P. Murine interleukin-4 displays potent anti-tumor activity in vivo. Cell. 1989;57:50312. [DOI] [PubMed]
    Stoppacciaro A, Paglia P, Lombardi L, Parmiani G, Baroni C, Colombo MP. Genetic modification of a carcinoma with the IL-4 gene increases the influx of dendritic cells relative to other cytokines. Eur J Immunol. 1997;27:237582. [DOI] [PubMed]
    Osawa E, Nakajima A, Fujisawa T, Kawamura YI, Toyama-Sorimachi N, Nakagama H, et al. Predominant T helper type 2-inflammatory responses promote murine colon cancers. Int J Cancer. 2006;118:22326. [DOI] [PubMed]
    Jayakumar A, Bothwell ALM. Stat6 promotes intestinal tumorigenesis in a mouse model of adenomatous polyposis by expansion of MDSCs and inhibition of cytotoxic CD8 response. Neoplasia. 2017;19:595605. [DOI] [PubMed] [PMC]
    McClellan JL, Steiner JL, Day SD, Enos RT, Davis MJ, Singh UP, et al. Exercise effects on polyp burden and immune markers in the ApcMin/+ mouse model of intestinal tumorigenesis. Int J Oncol. 2014;45:8618. [DOI] [PubMed] [PMC]
    Li Z, Jiang J, Wang Z, Zhang J, Xiao M, Wang C, et al. Endogenous interleukin-4 promotes tumor development by increasing tumor cell resistance to apoptosis. Cancer Res. 2008;68:868794. [DOI] [PubMed]
    Nappo G, Handle F, Santer FR, McNeill RV, Seed RI, Collins AT, et al. The immunosuppressive cytokine interleukin-4 increases the clonogenic potential of prostate stem-like cells by activation of STAT6 signalling. Oncogenesis. 2017;6:e342. [DOI] [PubMed] [PMC]
    Chevalier MF, Trabanelli S, Racle J, Salomé B, Cesson V, Gharbi D, et al. ILC2-modulated T cell-to-MDSC balance is associated with bladder cancer recurrence. J Clin Invest. 2017;127:291629. [DOI] [PubMed] [PMC]
    Shi J, Song X, Traub B, Luxenhofer M, Kornmann M. Involvement of IL-4, IL-13 and their receptors in pancreatic cancer. Int J Mol Sci. 2021;22:2998. [DOI] [PubMed] [PMC]
    Suzuki A, Leland P, Joshi BH, Puri RK. Targeting of IL-4 and IL-13 receptors for cancer therapy. Cytokine. 2015;75:7988. [DOI] [PubMed]
    Knudson KM, Hwang S, McCann MS, Joshi BH, Husain SR, Puri RK. Recent advances in IL-13Rα2-directed cancer immunotherapy. Front Immunol. 2022;13:878365. [DOI] [PubMed] [PMC]
    Uhm TG, Kim BS, Chung IY. Eosinophil development, regulation of eosinophil-specific genes, and role of eosinophils in the pathogenesis of asthma. Allergy Asthma Immunol Res. 2012;4:6879. [DOI] [PubMed] [PMC]
    Mori Y, Iwasaki H, Kohno K, Yoshimoto G, Kikushige Y, Okeda A, et al. Identification of the human eosinophil lineage-committed progenitor: revision of phenotypic definition of the human common myeloid progenitor. J Exp Med. 2009;206:18393. [DOI] [PubMed] [PMC]
    Takatsu K, Nakajima H. IL-5 and eosinophilia. Curr Opin Immunol. 2008;20:28894. [DOI] [PubMed]
    Zaynagetdinov R, Sherrill TP, Gleaves LA, McLoed AG, Saxon JA, Habermann AC, et al. Interleukin-5 facilitates lung metastasis by modulating the immune microenvironment. Cancer Res. 2015;75:162434. [DOI] [PubMed] [PMC]
    Gatault S, Delbeke M, Driss V, Sarazin A, Dendooven A, Kahn JE, et al. IL-18 is involved in eosinophil-mediated tumoricidal activity against a colon carcinoma cell line by upregulating LFA-1 and ICAM-1. J Immunol. 2015;195:248392. [DOI] [PubMed]
    Carretero R, Sektioglu IM, Garbi N, Salgado OC, Beckhove P, Hämmerling GJ. Eosinophils orchestrate cancer rejection by normalizing tumor vessels and enhancing infiltration of CD8+ T cells. Nat Immunol. 2015;16:60917. Erratum in: Nat Immunol. 2016;176:214. [DOI] [PubMed]
    Prizment AE, Anderson KE, Visvanathan K, Folsom AR. Inverse association of eosinophil count with colorectal cancer incidence: atherosclerosis risk in communities study. Cancer Epidemiol Biomarkers Prev. 2011;20:18614. [DOI] [PubMed] [PMC]
    Harbaum L, Pollheimer MJ, Kornprat P, Lindtner RA, Bokemeyer C, Langner C. Peritumoral eosinophils predict recurrence in colorectal cancer. Mod Pathol. 2015;28:40313. [DOI] [PubMed]
    Onesti CE, Josse C, Boulet D, Thiry J, Beaumecker B, Bours V, et al. Blood eosinophilic relative count is prognostic for breast cancer and associated with the presence of tumor at diagnosis and at time of relapse. Oncoimmunology. 2020;9:1761176. [DOI] [PubMed] [PMC]
    Moreira A, Leisgang W, Schuler G, Heinzerling L. Eosinophilic count as a biomarker for prognosis of melanoma patients and its importance in the response to immunotherapy. Immunotherapy. 2017;9:11521. [DOI] [PubMed]
    Schuijs MJ, Png S, Richard AC, Tsyben A, Hamm G, Stockis J, et al. ILC2-driven innate immune checkpoint mechanism antagonizes NK cell antimetastatic function in the lung. Nat Immunol. 2020;21:9981009. [DOI] [PubMed] [PMC]
    van Driel WJ, Kievit-Tyson P, van den Broek LC, Zwinderman AH, Trimbos BJ, Fleuren GJ. Presence of an eosinophilic infiltrate in cervical squamous carcinoma results from a type 2 immune response. Gynecol Oncol. 1999;74:18895. [DOI] [PubMed]
    Pinto A, Aldinucci D, Gloghini A, Zagonel V, Degan M, Perin V, et al. The role of eosinophils in the pathobiology of Hodgkin’s disease. Ann Oncol. 1997;8:S8996. [PubMed]
    Utsunomiya A, Ishida T, Inagaki A, Ishii T, Yano H, Komatsu H, et al. Clinical significance of a blood eosinophilia in adult T-cell leukemia/lymphoma: a blood eosinophilia is a significant unfavorable prognostic factor. Leuk Res. 2007;31:91520. [DOI] [PubMed]
    Cai L, Zhang Y, Zhang Y, Chen H, Hu J. Effect of Th9/IL-9 on the growth of gastric cancer in nude mice. Onco Targets Ther. 2019;12:222534. [DOI] [PubMed] [PMC]
    Wan J, Wu Y, Huang L, Tian Y, Ji X, Abdelaziz MH, et al. ILC2-derived IL-9 inhibits colorectal cancer progression by activating CD8+ T cells. Cancer Lett. 2021;502:3443. [DOI] [PubMed]
    Parrot T, Allard M, Oger R, Benlalam H, Raingeard de la Blétière D, Coutolleau A, et al. IL-9 promotes the survival and function of human melanoma-infiltrating CD4+CD8+ double-positive T cells. Eur J Immunol. 2016;46:177082. [DOI] [PubMed]
    Do Thi VA, Park SM, Lee H, Kim YS. Ectopically expressed membrane-bound form of IL-9 exerts immune-stimulatory effect on CT26 colon carcinoma cells. Immune Netw. 2018;18:e12. [DOI] [PubMed] [PMC]
    Hoelzinger DB, Dominguez AL, Cohen PA, Gendler SJ. Inhibition of adaptive immunity by IL9 can be disrupted to achieve rapid T-cell sensitization and rejection of progressive tumor challenges. Cancer Res. 2014;74:684555. [DOI] [PubMed] [PMC]
    Goswami R, Kaplan MH. A brief history of IL-9. J Immunol. 2011;186:32838. [DOI] [PubMed] [PMC]
    Ribatti D, Ennas MG, Vacca A, Ferreli F, Nico B, Orru S, et al. Tumor vascularity and tryptase-positive mast cells correlate with a poor prognosis in melanoma. Eur J Clin Invest. 2003;33:4205. [DOI] [PubMed]
    Nonomura N, Takayama H, Nishimura K, Oka D, Nakai Y, Shiba M, et al. Decreased number of mast cells infiltrating into needle biopsy specimens leads to a better prognosis of prostate cancer. Br J Cancer. 2007;97:9526. [DOI] [PubMed] [PMC]
    Strouch MJ, Cheon EC, Salabat MR, Krantz SB, Gounaris E, Melstrom LG, et al. Crosstalk between mast cells and pancreatic cancer cells contributes to pancreatic tumor progression. Clin Cancer Res. 2010;16:225765. [DOI] [PubMed] [PMC]
    Blair RJ, Meng H, Marchese MJ, Ren S, Schwartz LB, Tonnesen MG, et al. Human mast cells stimulate vascular tube formation. Tryptase is a novel, potent angiogenic factor. J Clin Invest. 1997;99:2691700. [DOI] [PubMed] [PMC]
    Barlow JL, Peel S, Fox J, Panova V, Hardman CS, Camelo A, et al. IL-33 is more potent than IL-25 in provoking IL-13–producing nuocytes (type 2 innate lymphoid cells) and airway contraction. J Allergy Clin Immunol. 2013;132:93341. [DOI] [PubMed]
    Roan F, Obata-Ninomiya K, Ziegler SF. Epithelial cell–derived cytokines: more than just signaling the alarm. J Clin Invest. 2019;129:144151. [DOI] [PubMed] [PMC]
    Ricardo-Gonzalez RR, Van Dyken SJ, Schneider C, Lee J, Nussbaum JC, Liang HE, et al. Tissue signals imprint ILC2 identity with anticipatory function. Nat Immunol. 2018;19:10939. [DOI] [PubMed] [PMC]
    Smithgall MD, Comeau MR, Yoon BR, Kaufman D, Armitage R, Smith DE. IL-33 amplifies both Th1- and Th2-type responses through its activity on human basophils, allergen-reactive Th2 cells, iNKT and NK cells. Int Immunol. 2008;20:101930. [DOI] [PubMed]
    Cho KA, Suh JW, Sohn JH, Park JW, Lee H, Kang JL, et al. IL-33 induces Th17-mediated airway inflammation via mast cells in ovalbumin-challenged mice. Am J Physiol Lung Cell Mol Physiol. 2012;302:L42940. [DOI] [PubMed]
    Schiering C, Krausgruber T, Chomka A, Fröhlich A, Adelmann K, Wohlfert EA, et al. The alarmin IL-33 promotes regulatory T-cell function in the intestine. Nature. 2014;513:5648. [DOI] [PubMed] [PMC]
    Pascual-Reguant A, Bayat Sarmadi J, Baumann C, Noster R, Cirera-Salinas D, Curato C, et al. TH17 cells express ST2 and are controlled by the alarmin IL-33 in the small intestine. Mucosal Immunol. 2017;10:143142. [DOI] [PubMed]
    Liu J, Qian B, Zhou L, Shen G, Tan Y, Liu S, et al. IL25 enhanced colitis-associated tumorigenesis in mice by upregulating transcription factor GLI1. Front Immunol. 2022;13:837262. [DOI] [PubMed] [PMC]
    Goto N, Fukuda A, Yamaga Y, Yoshikawa T, Maruno T, Maekawa H, et al. Lineage tracing and targeting of IL17RB+ tuft cell-like human colorectal cancer stem cells. Proc Natl Acad Sci U S A. 2019;116:129963005. [DOI] [PubMed] [PMC]
    Benatar T, Cao MY, Lee Y, Lightfoot J, Feng N, Gu X, et al. IL-17E, a proinflammatory cytokine, has antitumor efficacy against several tumor types in vivo. Cancer Immunol Immunother. 2010;59:80517. [DOI] [PubMed]
    Furuta S, Jeng YM, Zhou L, Huang L, Kuhn I, Bissell MJ, et al. IL-25 causes apoptosis of IL-25R–expressing breast cancer cells without toxicity to nonmalignant cells. Sci Transl Med. 2011;3:78ra31. [DOI] [PubMed] [PMC]
    Jovanovic I, Radosavljevic G, Mitrovic M, Juranic VL, McKenzie AN, Arsenijevic N, et al. ST2 deletion enhances innate and acquired immunity to murine mammary carcinoma. Eur J Immunol. 2011;41:190212. [DOI] [PubMed] [PMC]
    Jovanovic IP, Pejnovic NN, Radosavljevic GD, Pantic JM, Milovanovic MZ, Arsenijevic NN, et al. Interleukin-33/ST2 axis promotes breast cancer growth and metastases by facilitating intratumoral accumulation of immunosuppressive and innate lymphoid cells. Int J Cancer. 2014;134:166982. [DOI] [PubMed]
    Shin N, Son GM, Shin DH, Kwon MS, Park BS, Kim HS, et al. Cancer-associated fibroblasts and desmoplastic reactions related to cancer invasiveness in patients with colorectal cancer. Ann Coloproctol. 2019;35:3646. [DOI] [PubMed] [PMC]
    Milosavljevic MZ, Jovanovic IP, Pejnovic NN, Mitrovic SLJ, Arsenijevic NN, Simovic Markovic BJ, et al. Deletion of IL-33R attenuates VEGF expression and enhances necrosis in mammary carcinoma. Oncotarget. 2016;7:1810615. [DOI] [PubMed] [PMC]
    Corrigan CJ, Wang W, Meng Q, Fang C, Wu H, Reay V, et al. T-helper cell type 2 (Th2) memory T cell-potentiating cytokine IL-25 has the potential to promote angiogenesis in asthma. Proc Natl Acad Sci U S A. 2011;108:157984. [DOI] [PubMed] [PMC]
    Yao X, Wang W, Li Y, Huang P, Zhang Q, Wang J, et al. IL-25 induces airways angiogenesis and expression of multiple angiogenic factors in a murine asthma model. Respir Res. 2015;16:39. [DOI] [PubMed] [PMC]
    Maywald RL, Doerner SK, Pastorelli L, De Salvo C, Benton SM, Dawson EP, et al. IL-33 activates tumor stroma to promote intestinal polyposis. Proc Natl Acad Sci U S A. 2015;112:E248796. [DOI] [PubMed] [PMC]
    He Z, Chen L, Souto FO, Canasto-Chibuque C, Bongers G, Deshpande M, et al. Epithelial-derived IL-33 promotes intestinal tumorigenesis in Apc Min/+ mice. Sci Rep. 2017;7:5520. [DOI] [PubMed] [PMC]
    Bruchard M, Mignot G, Derangère V, Chalmin F, Chevriaux A, Végran F, et al. Chemotherapy-triggered cathepsin B release in myeloid-derived suppressor cells activates the Nlrp3 inflammasome and promotes tumor growth. Nat Med. 2013;19:5764. [DOI] [PubMed]
    Kang C, Jeong SY, Song SY, Choi EK. The emerging role of myeloid-derived suppressor cells in radiotherapy. Radiat Oncol J. 2020;38:110. [DOI] [PubMed] [PMC]
    Kawano M, Mabuchi S, Matsumoto Y, Sasano T, Takahashi R, Kuroda H, et al. The significance of G-CSF expression and myeloid-derived suppressor cells in the chemoresistance of uterine cervical cancer. Sci Rep. 2015;5:18217. [DOI] [PubMed] [PMC]
    Liao W, Overman MJ, Boutin AT, Shang X, Zhao D, Dey P, et al. KRAS-IRF2 axis drives immune suppression and immune therapy resistance in colorectal cancer. Cancer Cell. 2019;35:55972.E7. [DOI] [PubMed] [PMC]
    Weber R, Fleming V, Hu X, Nagibin V, Groth C, Altevogt P, et al. Myeloid-derived suppressor cells hinder the anti-cancer activity of immune checkpoint inhibitors. Front Immunol. 2018;9:1310. [DOI] [PubMed] [PMC]
    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;8:350311. [PubMed]
    Guo C, Ouyang Y, Cai J, Xiong L, Chen Y, Zeng X, et al. High expression of IL-4R enhances proliferation and invasion of hepatocellular carcinoma cells. Int J Biol Markers. 2017;32:e38490. [DOI] [PubMed]
    Sosman JA, Fisher SG, Kefer C, Fisher RI, Ellis TM. A phase I trial of continuous infusion interleukin-4 (IL-4) alone and following interleukin-2 (IL-2) in cancer patients. Ann Oncol. 1994;5:44752. [DOI] [PubMed]
    Whitehead RP, Lew D, Flanigan RC, Weiss GR, Roy V, Glode ML, et al. Phase II trial of recombinant human interleukin-4 in patients with advanced renal cell carcinoma: a Southwest Oncology Group study. J Immunother. 2002;25:3528. [DOI] [PubMed]
    Whitehead RP, Unger JM, Goodwin JW, Walker MJ, Thompson JA, Flaherty LE, et al. Phase II trial of recombinant human interleukin-4 in patients with disseminated malignant melanoma: a Southwest Oncology Group study. J Immunother. 1998;21:4406. [DOI] [PubMed]
    Margolin K, Aronson FR, Sznol M, Atkins MB, Gucalp R, Fisher RI, et al. Phase II studies of recombinant human interleukin-4 in advanced renal cancer and malignant melanoma. J Immunother Emphasis Tumor Immunol. 1994;15:14753. [DOI] [PubMed]
    Tong H, Feng H, Hu X, Wang MF, Song YF, Wen XL, et al. Identification of interleukin-9 producing immune cells in endometrial carcinoma and establishment of a prognostic nomogram. Front Immunol. 2020;11:544248. [DOI] [PubMed] [PMC]
    Heim L, Yang Z, Tausche P, Hohenberger K, Chiriac MT, Koelle J, et al. IL-9 producing tumor-infiltrating lymphocytes and Treg subsets drive immune escape of tumor cells in non-small cell lung cancer. Front Immunol. 2022;13:859738. [DOI] [PubMed] [PMC]
    Eiró N, González L, González LO, Fernandez-Garcia B, Lamelas ML, Marín L, et al. Relationship between the inflammatory molecular profile of breast carcinomas and distant metastasis development. PLoS One. 2012;7:e49047. [DOI] [PubMed] [PMC]
    Blomberg OS, Spagnuolo L, Garner H, Voorwerk L, Isaeva OI, van Dyk E, et al. IL-5-producing CD4+ T cells and eosinophils cooperate to enhance response to immune checkpoint blockade in breast cancer. Cancer Cell. 2023;41:10623.E10. [DOI] [PubMed]
    Brown CE, Aguilar B, Starr R, Yang X, Chang WC, Weng L, et al. Optimization of IL13Rα2-targeted chimeric antigen receptor T cells for improved anti-tumor efficacy against glioblastoma. Mol Ther. 2018;26:3144. [DOI] [PubMed] [PMC]
    Zeng J, Zhang J, Yang YZ, Wang F, Jiang H, Chen HD, et al. IL13RA2 is overexpressed in malignant gliomas and related to clinical outcome of patients. Am J Transl Res. 2020;12:470214. [PubMed] [PMC]
    Emens LA, Middleton G. The interplay of immunotherapy and chemotherapy: harnessing potential synergies. Cancer Immunol Res. 2015;3:43643. [DOI] [PubMed] [PMC]
    Bachmann MF, Oxenius A. Interleukin 2: from immunostimulation to immunoregulation and back again. EMBO Rep. 2007;8:11428. [DOI] [PubMed] [PMC]