Next Article in Journal
Shell Distribution of Vitamin K3 within Reinforced Electrospun Nanofibers for Improved Photo-Antibacterial Performance
Previous Article in Journal
CXCL12 Gene Polymorphisms and Serum Levels: Associations with Multiple Sclerosis Prevalence and Clinical Parameters in Lithuania
Previous Article in Special Issue
Oleanolic Acid Dimers with Potential Application in Medicine—Design, Synthesis, Physico-Chemical Characteristics, Cytotoxic and Antioxidant Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Tumor Microenvironment in Pancreatic Cancer Immunotherapy: Current Status and Future Perspectives

by
Fotini Poyia
1,
Christiana M. Neophytou
2,
Maria-Ioanna Christodoulou
3 and
Panagiotis Papageorgis
1,*
1
Tumor Microenvironment, Metastasis and Experimental Therapeutics Laboratory, Basic and Translational Cancer Research Center, Department of Life Sciences, European University Cyprus, Nicosia 2404, Cyprus
2
Apoptosis and Cancer Chemoresistance Laboratory, Basic and Translational Cancer Research Center, Department of Life Sciences, European University Cyprus, Nicosia 2404, Cyprus
3
Tumor Immunology and Biomarkers Laboratory, Basic and Translational Cancer Research Center, Department of Life Sciences, European University Cyprus, Nicosia 2404, Cyprus
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(17), 9555; https://doi.org/10.3390/ijms25179555
Submission received: 26 July 2024 / Revised: 29 August 2024 / Accepted: 30 August 2024 / Published: 3 September 2024
(This article belongs to the Special Issue Recent Advances in Anti-Cancer Drugs)

Abstract

:
Pancreatic cancer comprises different subtypes, where most cases include ductal adenocarcinoma (PDAC). It is one of the deadliest tumor types, with a poor prognosis. In the majority of patients, the disease has already spread by the time of diagnosis, making full recovery unlikely and increasing mortality risk. Despite developments in its detection and management, including chemotherapy, radiotherapy, and targeted therapies as well as advances in immunotherapy, only in about 13% of PDAC patients does the overall survival exceed 5 years. This may be attributed, at least in part, to the highly desmoplastic tumor microenvironment (TME) that acts as a barrier limiting perfusion, drug delivery, and immune cell infiltration and contributes to the establishment of immunologically ‘cold’ conditions. Therefore, there is an urgent need to unravel the complexity of the TME that promotes PDAC progression and decipher the mechanisms of pancreatic tumors’ resistance to immunotherapy. In this review, we provide an overview of the major cellular and non-cellular components of PDAC TME, as well as their biological interplays. We also discuss the current state of PDAC therapeutic treatments and focus on ongoing and future immunotherapy efforts and multimodal treatments aiming at remodeling the TME to improve therapeutic efficacy.

1. Introduction

Pancreatic cancer (PC) is the fourth leading cause of cancer-associated deaths worldwide both in men and women [1]. The most common form of PC is pancreatic ductal adenocarcinoma (PDAC), accounting for 90% of all cases [2]. With persistently escalating incidence and minimal change in mortality rates, PDAC is predicted to become the second most frequent cause of cancer death within the next six years [3]. The average lifetime risk of developing PC is around 1.5%, which translates to 1 in 64 people. The aggressive behavior and the rapid development of metastases explain why PC patients have a 5-year overall survival (OS) rate of less than 13% [4,5]. The progression from stage I to stage IV is estimated to last just over a year. Moreover, PDAC symptoms, such as type-2 diabetes outbreak, abdominal and back discomfort, lack of appetite, and weight loss, are not specific and are often misinterpreted, leading to poor and late diagnosis. The risk of developing PC increases with age, with the average age at the time of diagnosis to be 70 years [6,7]. Furthermore, only 10% of PC incidents are hereditary, and approximately 90% are sporadic. Among patients with PC, 90% carry a KRAS mutation, which is considered a driver gene for PC progression, and 50–80% have inactivating mutations in TP53, CDKN2A, and SMAD4 [8]. While novel treatments have significantly improved the OS rate in other cancers, PC still constitutes one of the deadliest forms of malignancy, with a median survival of 15.5 months after surgery. Only up to 10% of patients who receive a timely diagnosis become disease-free after treatment. Consistent with this, patients who are diagnosed before metastatic growth have an average survival time of just 3 to 3.5 years, whereas patients diagnosed with final stage of PC have a life expectancy of about 3–5 months [2,9]. Surgical resection is the mainstay of curative treatment for patients with localized pancreatic tumors. In cases of borderline resectable PC (BRPC) or unresectable locally advanced PC (LAPC), pre-operative neoadjuvant therapy using FOLFIRINOX (5-fluorouracil, leucovorin, oxaliplatin, and irinotecan) or gemcitabine/nab-paclitaxel), with or without radiation therapy, is used to downsize the tumor to facilitate surgical resection. These are also administered as the standard-of-care treatment for patients with metastatic PDAC, whereas pembrolizumab is considered for <1% of metastatic patients with high microsatellite instability (MSI-H) or mismatch repair deficiency (dMMR). The poly (ADP-ribose) polymerase (PARP) inhibitor Olaparib is used as maintenance therapy in patients with BRCA1/2 mutations after initial chemotherapy response [10].
This review summarizes the current state of PC treatment, the main characteristics of pancreatic cancer TME, and their implications in exploring new potential therapeutic targets and combination approaches. Specifically, we focus on immunotherapy and the roles of PC TME in hindering its efficacy and discuss the latest pre-clinical and clinical evidence regarding multimodal treatment strategies to improve immunotherapy outcomes in PC patients.

2. Current Therapies for Pancreatic Cancer

Different types of treatments are available for people with PC, with the only curative approach being surgical management. Importantly, if PC has spread, palliative treatment can only improve patients’ quality of life by controlling the symptoms of this disease. In this section, we discuss the current types of standard-of-care treatment for PC.

2.1. Surgery

While surgery is the only available treatment with curative potential, 85% of newly diagnosed pancreatic tumors are considered unresectable due to late diagnosis and locally advanced disease or metastasis to distal organs [11,12,13]. However, up to 80% of the patients that are eligible for surgery may relapse and die after the operation [9,14]. In cases where cancer has spread throughout the pancreas but is still resectable, a total pancreatectomy is performed, and post-operative patients receive pancreatic enzymes for life [12]. Therefore, even though surgical management plays a vital role in PC cure, few patients can benefit long-term due to compromised quality of life and other surgery-related risks, including post-operative infections [10]. Hence, administration of chemotherapy after surgical management of PDAC has been shown to lower the risk of recurrence and improve survival rates.

2.2. Chemotherapy

Chemotherapy has been typically the standard-of-care treatment for primary and metastatic PDAC. Gemcitabine is the reference treatment as anti-cancer chemotherapy in PC patients that are not eligible for combination chemotherapy [15]. Additional treatment protocols, including neoadjuvant chemotherapy and adjuvant chemotherapy for resected PC, are often combined with surgery to increase the rate of successful resection and extent of survival. Unfortunately, even with cytotoxic chemotherapy, the OS of locally advanced PC patients with metastatic disease is only 5–6 months, with a response rate of 5.4% [16,17]. NALIRIFOX, an irinotecan liposome (ONIVYDE) with oxaliplatin, fluorouracil, and leucovorin, was approved by the FDA (Food and Drug Administration) as a first-line treatment for metastatic pancreatic adenocarcinoma. The NAPOLI 3 clinical trial (NCT04083235) demonstrated a statistically significant improvement in overall survival for the NALIRIFOX arm over the Nab-paclitaxel plus gemcitabine arm (11.1 months vs. 9.2 months). The application of adjuvant chemotherapy was supported by the phase III CONKO-001 randomized trial (ISRCTN34802808), which showed a significant benefit of adjuvant gemcitabine after PC resection against surgery alone. A prolonged disease-free survival (DFS) (13.4 vs. 6.7 months) and 5-year OS (20.7% vs. 10.4%) was observed, including patients with R0 or R1 resected tumors. Long-term follow-up also displayed an increased 10-year OS of 5%. Furthermore, the ESPAC randomized trials aimed to identify the most effective chemotherapy scheme. The most recent ESPAC-4 trial (ISRCTN96397434) of Neoptolemos et al. underlined the advantage of gemcitabine plus capecitabine in the adjuvant setting after surgery, resulting in median survival of 26 months and 5-year survival of 30% [18].
Progress has also been achieved with FOLFIRINOX based on the PRODIGE-24/CCTG clinical trial (NCT01526135), which compared the outcomes of FOLFIRINOX against gemcitabine in patients with resected PDAC. The results indicated a clear improvement in OS using FOLFIRINOX against the gemcitabine group (54.4 vs. 35 months). The 5-year disease-free survival rate was 26.1% for FOLFIRINOX-treated patients and 19.0% for gemcitabine [19]. However, FOLFIRINOX administration was correlated with an increased risk of complications. Improved survival was observed in metastatic patients after combination treatment of gemcitabine and nab-paclitaxel, with fewer side effects than FOLFIRINOX [20].
Although adjuvant treatment may provide a survival benefit, about 74% of patients still relapse within two years [21]. A meta-analysis of 13 trials demonstrated downstaging of unresectable tumors after neoadjuvant FOLFIRINOX treatment, achieving an R0 resection rate of 40% [11]. However, resectable disease may become unresectable upon receiving neoadjuvant therapy due to complications which can prevent surgical management [12]. Further clinical studies are needed to reach optimal treatment protocols for administering chemotherapy in the neoadjuvant setting.
Up to 80% of patients do not respond to neoadjuvant chemotherapy to become eligible for surgery. The standard treatment of these tumors using systemic chemotherapy, commonly FOLFIRINOX, gemcitabine, and/or nab-paclitaxel, aims to control the disease which has already metastasized [22]. However, the OS of locally advanced PDAC (LAPC) patients remains below one year [23]. Interestingly, Shelemey et al. reported a case of shrinkage of an adenocarcinoma mass on the pancreatic tail and of liver metastases after FOLFIRINOX administration. Importantly, upon completion of 37 cycles of FOLFIRI (FOLFIRINOX without oxaliplatin), the pancreatic mass disappeared, the liver metastasis decreased, and no recurrence was observed [24].

2.3. Radiotherapy

External radiation therapy may be a treatment option for PC patients depending on the stage of the disease; specific tumor characteristics, such as size and location; and the patient’s overall health. Although the aim of radiation is to control cancer development and relieve patient symptoms, its use has been limited due to the inability to administer effective radiation doses in the pancreatic tumor. The inherent resistance of the pancreatic tumor, the intraperitoneal tumor location, and the neighboring organs impose barriers on the effective and targeted application of radiation therapy [25]. To address these problems, therapies have been developed to decrease tumor volume using the FDA-approved stereotactic body radiotherapy (SBRT). It has been shown that SBRT is beneficial as an adjuvant treatment in high-risk PDAC patients with affected tumor margins after surgery [26]. Additionally, chemoradiotherapy, the use of radiotherapy as neoadjuvant treatment together with chemotherapy, has been assessed in patients with borderline resectable PC. The combination of chemotherapy with photon radiotherapy has demonstrated improved local control rates, but no OS benefit if no surgical resection is followed [27]. A phase I/II study (NCT00438256) by Hong et al. showed that preoperative chemoradiation along with proton radiotherapy and capecitabine followed by early surgery is feasible with low toxicity levels [28,29]. Several clinical trials are currently underway to evaluate long-term survival benefits from the combination of radiotherapy with other anti-cancer treatments.

2.4. Targeted Therapies

Developments in whole-genome sequencing approaches have aided the design of personalized targeted therapies through mapping of key genetic alterations that drive PC progression. Mutations that are produced by genomic instability frequently produce cancer cell vulnerabilities that could be key for effective anti-cancer therapies. Diverse targeted agents have been assessed either alone or in combination with chemotherapeutic drugs against PDAC [30]. Unfortunately, most of these approaches have failed to prolong patient survival, mainly due to the hypovascular and desmoplastic nature of PC’s impenetrable stroma, as is described in more detail below [31]. A series of phase III clinical trials (NCT00088894, NCT01214720, NCT00471146, NCT00541021) in patients with advanced or metastatic PC failed to enhance OS after treatment with VEGF inhibitors, including bevacizumab (humanized anti-VEGF-A monoclonal antibody), sorafenib (VEGF-R inhibitor), and axitinib (tyrosine kinase inhibitor, TKI), along with gemcitabine or gemcitabine/erlotinib [32,33,34,35,36]. Furthermore, therapies targeting key signaling pathways in PDAC, including the anti-insulin-like growth factor 1 (IGF-1) receptor using the antibodies ganitumab and cixutumumab, the multi-kinase inhibitor masitinib, and the phosphoinositide 3-kinase (PI3K) inhibitor rigosertib, have been proven ineffective in randomized clinical trials [30]. Erlotinib is the only targeted agent that has exhibited a statistically significant, yet clinically modest, effect in patient survival. In a randomized trial (NCIC CTG PA.3) by Moore et al., it was reported that a combination of erlotinib with gemcitabine rendered a survival benefit of 2 weeks compared to gemcitabine alone [37].

3. Pancreatic Tumor Microenvironment and Opportunities for Therapeutic Interventions

Understanding the complex molecular composition and cellular interactions between cancer cells and the TME is of paramount importance for improving outcomes of existing therapeutic approaches and for designing novel personalized and precision therapeutic strategies for PC patients. Here, we describe the major components of the PDAC microenvironment, analyze the interactions of malignant cells with stromal and immune cells during PC formation and progression, and discuss their therapeutic implications.

3.1. Non-Cellular Components and Desmoplasia in the Pancreatic TME

Throughout all stages of PDAC growth, tumor cells are not only in physical, but also in biological contact with the stroma via secreted factors mediating cell-to-cell communication. These continuous interactions affect the TME during oncogenesis and the accompanying stromagenesis [38,39,40]. Tumor desmoplasia is a phenomenon which refers to the growth of dense connective tissue or stroma around a tumor mass. It is characterized by the proliferation of fibroblasts and the production of extracellular matrix (ECM) components, such as collagen, resulting in a fibrous or hard tissue environment around the tumor, which is a hallmark of primary and metastatic PC [41]. ECM is a high-density network made up mainly of matrix proteins, including I, III, and IV collagens, which are secreted by cellular components of the TME, such as fibroblasts and pancreatic stellate cells (PSCs), along with hyaluronic acid, fibronectin, and glycosaminoglycan, and can collectively represent up to 90% of the tumor mass [13]. Specifically, type IV collagen has been proposed as a potential serum biomarker in predicting PC patient survival following a surgical operation [42]. Excessive deposition of ECM components in the TME results in increased tumor stiffness, solid stress, and interstitial fluid pressure (IFP), as well as application of mechanical forces by the surrounding stroma which compress blood vessels, leading to hypoperfusion, hypoxia, and decreased infiltration of cytotoxic immune cells [43,44]. Importantly, this highly dense ECM tissue also behaves as a physical barrier diminishing the effective penetration of anti-cancer drugs [45,46,47,48]. Losartan, an angiotensin II receptor blocker which suppresses TGF-β activity in PC, was shown to reduce the levels of collagen and hyaluronan in PDAC models. As a result, losartan improved vascular perfusion and, thus, the delivery and efficacy of cytotoxic agents, such as 5-FU and Doxorubicin. Interestingly, tumors were significantly smaller in mice administrated with losartan combined with either 5-FU or Doxorubicin instead of pancreatic tumors treated with one of them as a monotherapy [49,50]. Furthermore, Anup et al. indicated that an analysis of metastatic PDAC patients treated with FOLFIRINOX plus losartan revealed a longer progression-free survival (PFS) than the control group, although no statistical significance was observed [51]. Overall, it is unambiguously accepted that pancreatic tumor stiffness and desmoplasia play crucial roles in promoting disease aggressiveness, therapeutic resistance, and poor prognosis of PC patients [45].

3.2. Non-Immune Cellular Components in the TME

The excessive production of ECM components resulting in the formation of the extremely dense PC stroma is mainly mediated by PSCs which represent the main fibroblastic cell type in PDAC, along with cancer-associated fibroblasts (CAFs). CAFs are a heterogeneous population of fibroblasts found in the pancreatic tumor stroma. They can originate from various sources, including resident fibroblasts, mesenchymal stem cells, and even PSCs that have been activated and further modified in the tumor microenvironment. They can be further subcategorized in myofibroblastic CAFs (myCAFs) and inflammatory CAFs (iCAFs) [52,53]. MyCAFs highly express α smooth muscle actin (α-SMA) and are often localized close to PC cell clusters, whereas inflammatory CAFs (iCAFs) are located more distantly from the tumor cells in the desmoplastic stroma [54].

3.2.1. Pancreatic Stellate Cells (PSCs)

In the healthy pancreas, PSCs remain in a quiescent state. Upon pancreatic injury, inflammation, or tumor formation, PSCs become activated, transforming into a myofibroblast-like phenotype, which is associated with increased production of ECM components. Activated PSCs within and surrounding the tumor produce collagen and other subcomponents of the ECM, which contribute to desmoplasia, increased solid stress, and poor vascularity due to vessel compression, a characteristic feature of PDAC [55]. It has also been suggested that PSCs are involved in cancer initiation, angiogenesis, epithelial-to-mesenchymal transition (EMT), local invasion, and metastasis of PC cells [56] by expressing paracrine molecules, such as transforming growth factor-β (TGF-β) and platelet-derived growth factors (PDGFs) [57,58]. Moreover, PSCs have been shown to support PC progression by increasing the number of immunosuppressive cells and inhibiting infiltration of cytotoxic CD8+ T cells [54,59,60].

3.2.2. Myofibroblastic and Inflammatory CAFs

Based on initial evidence that the desmoplastic stroma acts as a physical barrier to compromise efficient cytotoxic drug delivery in pancreatic tumors [61,62], subsequent efforts were aimed to eliminate or target myofibroblastic cancer-associated fibroblasts (myCAFs) from pancreatic cancers. However, these studies suggested that local depletion or inhibition of α-SMA+ myCAFs in a murine PDAC model reduced desmoplasia, but were associated with increased tumor aggressiveness, immunosuppression, and shorter survival instead of promoting anti-tumor effects [63,64]. Collectively, these studies have shown that myCAFs could restrain tumor growth and that therapeutic strategies targeting them may only be considered in the context of combination therapies.
Inflammatory CAFs (iCAFs) also originate from activated PSCs and other fibroblasts within the tumor microenvironment, but are distinguished by their inflammatory profile based on the expression of inflammatory cytokines and chemokines. For example, secretion of IL-6 by iCAFs acts synergistically with IL-10 and TGF-β to inhibit dendritic cell proliferation, therefore inhibiting tumor-antigen presentation [65,66,67]. They may also express fibroblast-specific markers like fibroblast activation protein (FAP) and PDGFRβ, but are primarily defined by their secretion profile. By promoting inflammation, they influence the recruitment and activation of M2-type tumor-associated macrophages (TAMs), myeloid-derived suppressor cells (MDSCs), and regulatory T cells (Tregs), which can contribute to immunosuppression and tumor immune evasion [54,68].

3.2.3. Endothelial Cells

Moreover, endothelial cells play crucial roles in the development and structure of blood vessels in pancreatic tumors [69]. Studies have shown that angiogenesis in PDAC demonstrates abundant production of the vascular endothelial growth factor (VEGF) by endothelial as well as PC cells under the control of hypoxia-inducible factor 1 subunit alpha (HIF1α) and signal transducer and activator of transcription 3 (STAT3) under hypoxia conditions [70,71,72]. Unfortunately, so far, no anti-angiogenic therapy has been clinically effective in PDAC.

3.3. Immune Cell Populations in the PDAC TME

It becomes increasingly evident that the functional network of interactions between tumor, stromal, and immune cells supports the progression of PC. During pancreatic tumorigenesis, the immune system may act as a double-edged sword; certain immune components can suppress tumor growth or progression by recognizing and eliminating mutated cells, while others can promote an immunosuppressive and pro-tumorigenic environment. Although immune cell populations account for up to 50% of the total cell number in PDAC, only a small subset are tumoricidal cells [73]. The major immune cell types in the PDAC TME that play crucial roles in these processes include macrophages, myeloid-derived suppressor cells (MDSCs), natural killer cells (NKs), neutrophils, dendritic cells (DC), and T lymphocytes.

3.3.1. Tumor-Associated Macrophages (TAMs)

Tumor-associated macrophages (TAMs) are one of the most abundant immune cell types in the pancreatic TME; they originate from circulating monocytes and are recruited to the pancreatic tumor site by various chemokines and growth factors secreted locally [74,75,76]. They are considered a heterogenous population due to their plasticity and ability to switch between the anti-tumoral M1 and pro-tumoral M2 phenotypes depending on the conditions in the TME and activation signals [77,78]. The majority of TAMs in PDAC display the M2-polarized phenotype, characterized by the surface markers CD163 and CD206, and secrete IL-10 and TGF-β [79]. Importantly, Ino et al. reported that the tumor-infiltrating % of M1high/M2low may act as an independent prognostic factor for OS in PDAC patients [80]. Macrophage depletion was found to reduce liver and lung metastasis in an orthotopic PDAC mouse model [81]. Several findings support that TAMs could also regulate PDAC metastasis through secretion of exosomes containing miRNAs that promote tumor cell migration, EMT, and ECM remodeling [82,83]. Finally, in an in vivo PDAC mouse model study, cytidine deaminase (CDA), a key metabolizer of gemcitabine, was found to be upregulated by TAMs, resulting in an anti-inflammatory macrophage phenotype and chemotherapy resistance [84].

3.3.2. Myeloid-Derived Suppressor Cells

Myeloid-derived suppressor cells represent a mixture of immature myeloid cells with a critical role in immunosuppression in PC. They are abundantly found in PDAC and are dispersed throughout the tumor. Their accumulation is associated with the stage of the disease [73,85]. High levels of granulocyte macrophage colony-stimulating factor (GM-CSF) produced by tumor cells are associated with MDSC development and migration through the bloodstream [86]. Pylayeva et al. demonstrated that the oncogenic KRASG12D mutation, present in more than 90% of PC cases, is responsible for the upregulation of GM-CSF [87]. It is also known that MDSC differentiation is triggered by the STAT3 signaling pathway upon IL-6 release from activated PSCs [88]. Furthermore, MDSCs can suppress CD4+ and CD8+ T cell responses via several mechanisms, first by upregulating PD-L1 and inhibiting T cell activation and tumor tolerance [89]. Moreover, MDSCs were shown to stimulate expansion of immunosuppressive Tregs [90] which, in turn, induce MDSCs to release reactive oxygen species (ROS), causing oxidative stress in T cells to further inhibit antigen-presenting proliferation [91]. In addition, MDSCs downregulate L-selectin in CD4+ and CD8+ T cells impairing T cell homing to lymph nodes [92].

3.3.3. Natural Killer Cells

Natural killer (NK) cells account for 5–20% of human peripheral blood mononuclear cells (PBMCs). They are characterized by the expression of the natural cytotoxicity receptor (NCR) NKp46 and the neural cell adhesion molecule (NCAM/CD56) [93,94]. Upon activation, NKs secrete IFN-γ, GM-CSF, tumor necrosis factor-α (TNF-α), and chemokines that regulate the functions of other innate and adaptive immune cells [95]. The number of circulating NKs in PDAC is positively correlated with median patient survival [96]. Increasing evidence suggests that interactions within the pancreatic TME can regulate the phenotype and function of NKs. It was recently proposed that tumor cell-derived extracellular vesicles from the TME can functionally change NK cells by inhibiting the recognition and killing of cancer cells [97]. The activity of NKs was shown to be reduced in PDAC compared to peripheral blood leukocytes of healthy blood donors based on the production of lower levels of granzyme B and perforin, which are crucial for the elimination of cancer cells [98]. Human PC cells express Fas ligand, leading to apoptosis of tumor-infiltrating lymphocytes, including NKs [99]. Moreover, NKs’ recognition and killing abilities are impaired by IL-10, TGF-β, Indoleamine 2,3-dioxygenase (IDO) and metalloproteinases produced by PDAC cells [100]. Reduced levels of the activating receptor NKp46 are correlated with PC progression [101]. Finally, C-X-C motif chemokine receptor 2 (CXCR2) has been found to be essential for the recruitment of NKs into the TME. This chemokine receptor is downregulated in PDAC patients, resulting in limited NK cell infiltration [13].

3.3.4. Neutrophils

Neutrophils, an essential component of the innate immune system, have evident anticancer activity and can induce phagocytosis as well as direct cytotoxic elimination of malignant cells. They infiltrate the TME upon interaction of CXCR2⁺ neutrophils with CXCL1/2 ligands [102]. During early stages of cancer development, tumor-associated neutrophils (TANs) can be distinguished according to their cytokine status, activation, and effects on cancer cells. N1 TANs regulated by IFN-α can exert anti-tumor effects. They exert cytotoxic action against tumor cells and prevent immunosuppression within the TME mainly by recruiting and activating CD8⁺ T cells [13,103]. On the contrary, N2 TANs induced by TGF-β undergo a phenotypic switch to a pro-tumoral phenotype, promoting tumor progression by remodeling the TME, whereas TGF-β blockade was able to reverse this effect in colorectal cancer in an in vitro study [90,104]. In addition, neutrophils contribute to tumor invasion and metastasis through secretion of VEGF and the metalloproteinase-9 (MMP-9), which are related to angiogenesis [13]. Therefore, high levels of TANs could provide a survival advantage for tumors, resulting in relapse and poor clinical outcomes for PC patients [105]. Blockade of CXCR2 in vivo hindered neutrophils’ entrance into PDAC stroma to significantly expand mouse survival [102]. Moreover, prevention of neutrophil maturation and migration by the tyrosine kinase inhibitor (TKI) lorlatinib was shown to abrogate PDAC development and metastasis in pre-clinical mouse models [106]. It was also shown that the high neutrophil/lymphocyte ratio (NLR) prior to therapy is associated with development of metastatic disease, and it could be used as a prognostic marker for OS in PC patients [107]. Additionally, neutrophils can form neutrophil extracellular traps (NETs). These structures consist of extracellular DNA released together with proteolytic enzymes that can enclose the tumor and inhibit the penetration of other anti-tumorigenic agents. NETs can also stimulate metastasis by attracting cancer cells from distant sites [13]. Treatment of PDAC mouse models with DNase I, a NET inhibitor, decreased the number of CAFs in the metastatic liver environment and thus suppressed metastasis [108]. Altogether, current evidence suggests that neutrophils can work synergistically with other cellular components to remodel the TME primarily in favor of pancreatic tumor growth.

3.3.5. Dendritic Cells

Dendritic cells (DCs) are trained antigen-presenting cells able to regulate anti-tumor immune responses by activating CD8+ and CD4+ T cells via MHC class I and II molecules, respectively [109]. DCs infiltrate pancreatic tumor lesions, and their abundance is associated with inhibition of disease progression [110]. The CD86 costimulatory marker expressed on DCs provides signals necessary for T cell activation and survival by binding to CD28 on the surfaces of T cells. However, cytotoxic T-lymphocyte-associated protein 4 (CTLA-4) produced by Tregs binds to CD86 with higher affinity than CD28, thus affecting CD8+ T cell activation and the recruitment of additional DCs [111]. The capacity of DCs is further impaired by PC cells through inhibition of their recruitment, maturation, and survival. The binding of CD154 (CD40L) on T helper cells to CD40 activates antigen-presenting cells. When DCs interact with cancer cells, immunosuppressive cytokines and chemokines, such as IL-10, TGF-β, and GM-CSF, are secreted, decreasing CD40 expression and keeping DCs in an immature state [112,113]. In contrary, pro-inflammatory cytokines and chemokines secreted by tumor cells, which are required for DCs activation, were downregulated by the activation of STAT3 in a melanoma mouse model [114]. A clinical study by Kobayashi et al. suggested that standard chemotherapy along with peptide-pulsed DC vaccines can act synergistically to improve PC patient survival [115].

3.3.6. T Lymphocytes

T lymphocytes are mainly classified as CD8+ cytotoxic T cells (CTLs) and CD4+ helper T (Th) cells, which include Th1, Th2, Th17, and regulatory T cells (Tregs) [80]. Th1 cells promote cellular type I immunity against intracellular pathogens and tumors and Th2 cells are involved in the regulation of humoral type II immunity, whereas Th17 cells are the defense against extracellular pathogens. Finally, Tregs provide suppressive inflammatory responses to control autoimmunity, and they are able to diminish antitumor responses to promote tumor progression [13].
Pancreatic tumors are considered immunologically ‘cold’, displaying low infiltration of CD8+ CTLs that are localized along the invasive margin of the tumor border or in the surrounding fibrotic tissue [116]. PDAC patient tumors are usually abundant in Tregs that are inversely correlated with the presence of CD8+ CTLs and associated with poor clinical outcomes [117]. Furthermore, in the progression of different types of tumors, infiltrating CD8+ CTLs exhibit minimal activation and become exhausted. These non-functional CD8+ CTLs are characterized by impaired effector function, metabolism dysregulation, and less proliferative activity [118]. Additionally, TGF-β secretion in the TME inhibits CD8+ CTLs from producing cytolytic proteins, while PC cells often downregulate MHC-I expression, preventing recognition and cytotoxic activity by CD8+ CTLs [119].
In contrast to CD8+ CTLs, CD4+ helper T cells are a prominent feature of the infiltrated immune cells in the pancreatic TME. Th1 cells promote cell-mediated immune responses and are responsible for the activation of CD8+ T cells, NK cells, and M1-type macrophages [120]. On the other hand, Th2 lymphocytes assist humoral immune responses by producing a plethora of cytokines including IL-4, which contribute to the formation of the dense tumor stroma and the polarization of macrophages to M2 stage and promote PDAC progression [13]. The shift from Th1 to Th2 cells is a common characteristic in PDAC, correlated with decreased patient survival. Th2 skewing in PDAC is driven by the CAFs in stroma and cytokines such as IL-10 and TGF-β secreted by PC cells [117,119]. Furthermore, elevated numbers of Th17 cells in the tumor are associated with disease progression, and serum IL-17 levels are increased in PC patients and connected with disease severity [13,90]. Finally, Tregs are often found in higher numbers in pancreatic tumors, and usually contribute to an immunosuppressive environment that allows the tumor to evade the immune system and promote T cell exhaustion [89,121]. However, recent evidence in PDAC mouse models showed that Tregs depletion may not diminish immunosuppression but may promote tumor progression due to reduction in Tregs-mediated TGF-β secretion and subsequent loss of tumor-restraining fibroblasts. Moreover, upon Tregs reduction, chemokines CCL3, CCL6, and CCL8 were increased, resulting in restoration of immune responses [122,123]. Therefore, the role of Tregs in PC may be more complicated, and further studies are required in order to elucidate their detailed biological roles, depending on the cellular context.

3.3.7. B Lymphocytes

Through the expression of B cell receptors on their surface, B lymphocytes bind to foreign antigens and initiate an antibody response. Tumor-infiltrating B cells (TIL-Bs) complement T cell-mediated antitumor immunity [124]. In PDAC, elevated B cell infiltration is generally correlated with better prognosis, especially when those B cells cluster in tertiary lymphoid structures (TLS) [125]. However, the role of B lymphocytes in PDAC tumorigenesis remains controversial. This contradiction could be explained as B cells acquire different phenotypes during tumor progression. Their function is also determined by their localization in the TME, either scattered at the periphery of tumor or forming complexes with CD8+ T cells [126]. The immunosuppressive B cells (Bregs), which are responsible for restricting ongoing immune responses and reestablishing immune homeostasis, represent only a small fraction of the entire B cell population in PDAC [127,128]. TIL-Bs have been shown to be involved in PDAC initiation, progression, and fibrogenesis. B1 cells constitute a unique B cell subset with abnormal receptor signaling, an unusual resting location, stimulation of T cell expansion, induction of Th17 cell differentiation, and production of immunomodulatory IL-10 [129]. Upon pancreas-specific HIF1-α depletion, fibro-inflammatory stroma secretes CXCL13, leading to an influx of B1 regulatory B cells into the tumor and thus promoting carcinogenesis [87]. Treatment of HIF1-α-deficient mice with B cell-depleting αCD20 monoclonal antibodies prevented pancreatic intraepithelial neoplasia (PanIN) progression and development of invasive carcinomas [130]. Lastly, targeting Bruton tyrosine kinase (BTK) using the BTK inhibitor ibrutinib was shown to inhibit B cell- and M2 macrophage-mediated T cell suppression to decrease PC growth [131].
In summary, as discussed above, the highly desmoplastic nature of PC TME results in elevated mechanical forces which compress blood vessels, leading to hypoperfusion, hypoxia, and decreased infiltration of cytotoxic immune cells. Even the small number of anti-tumor immune cells present in TME is either exhausted or with an immature phenotype [132]. Immune evasion perpetrated by the tumor cells involves aberrant expression of immune and cancer cell surface markers, secretion of immunosuppressive cytokine and chemokine molecules in the TME, and activation of immune checkpoint pathways, as described below [133]. Therefore, considering the highly immunosuppressive pancreatic TME, new combinatorial therapeutic approaches are urgently needed to overcome PC immune tolerance (Figure 1).

4. Current and Future Immunotherapy Strategies for Pancreatic Cancer

Cancer cells can survive and give rise to tumor development by, among other means, escaping immune surveillance either directly or indirectly via cells in the TME. Even though immunotherapy has revolutionized the treatment of various solid tumors during the last few years, it remains largely ineffective in PDAC patients by providing only a negligible improvement in patient survival [25,134,135]. The lack of PDAC responses to immunotherapies could be attributed, at least in part, to the low tumor mutation burden (TMB) in the vast majority of cases and to the highly desmoplastic TME, which collectively contribute to the development of an immunologically ‘cold’ environment [132]. PDAC has a highly desmoplastic TME with extensive fibrosis and extensive immunosuppression, which significantly compromises cytotoxic immune cell infiltration [58,136,137].
Under physiological conditions, immune system responses are regulated by various immune checkpoint pathways. Immune checkpoints are crucial modulators of the immune system, often exploited by cancer cells to evade immune surveillance. For example, programmed death protein–ligand 1 (PD-L1) or PD-L2 expressed by PC cells, MDCSs, or TAMs bind to PD-1 receptors on the surfaces of activated T cells, leading to T cell anergy or death [138,139]. Similarly, expression of CTLA-4 on the surface of T cells and binding to B7 molecules on DCs delivers an inhibitory signal that reduces T cell proliferation and activation and suppresses immune responses against tumor cells [140]. The upregulation of these inhibitory molecules and chronic antigen exposure leads to T cell exhaustion [13].
Over the last decade, several monoclonal antibodies targeting immune checkpoint molecules, such as PD-1 and CTLA-4, have been developed and granted FDA approval for the treatment of various solid tumors by reversing T cell dysfunction, leading to tumor killing [141,142]. Several studies have investigated the effects of anti-PD-1 (pembrolizumab, nivolumab), anti-PD-L1 (durvalumab), and anti-CTLA-4 (ipilimumab, tremelimumab) in PDAC as monotherapy or in combination with other approaches (Table 1). Favorable results regarding inhibition of tumor progression and improvement of patients’ survival observed in many cancer types were not observed in PDAC [134,135,143,144]. For the majority of PC patients, monotherapy treatment using PD-1 or CTLA-4 blockade has failed to produce any objective response, since some of the subjects experienced grade 3 or 4 adverse events related to treatment, and there were no responders [143,144]. The failure of ICB in pancreatic tumors is thought to be attributed, at least in part, to the low proportion of tumor-infiltrating T cells and the low tumor mutation burden (TMB) in PCs [73,145,146]. However, in a minor group of PDAC patients (~1%) with mismatch repair deficiency (dMMR) or microsatellite instability high (MSI-H), PD-1 blockade by pembrolizumab was shown to be effective and, currently, it is the only FDA-approved immunotherapy for patients with advanced PDAC. In this study, 8 out of the 86 patients had PC, and the objective response rate (ORR) among them was 62% (two patients had complete responses, three patients had partial responses, one patient had stable disease, and two patients were not evaluable) [147,148,149,150]. In the KEYNOTE-158 multi-cohort phase II study evaluating pembrolizumab, the median OS was 4 months in the PC subgroup, although the median duration of response was 13.4 months [151]. The next most promising outcomes of ICB were reported in PDAC patients after receiving anti-PD-1 together with gemcitabine/nab-paclitaxel, leading to PFS of 9.1 months and OS of 15 months [152]. Immunogenic cancer cell death includes the secretion of damage-associated molecular patterns from dying tumor cells that lead in the activation of tumor-specific immune responses, thus inducing long-term efficacy of anticancer drugs [153]. Hence, cytotoxic drugs may improve immunotherapeutic efficacy by stimulating immunogenic cancer cell death, decreasing tumor-induced immunosuppression, and enhancing effector T cell function and intra-tumoral infiltration [4].
A clinical trial (NCT01473940) examined the efficacy of ipilimumab combined with gemcitabine in patients with advanced PDAC, demonstrating a median PFS of 2.5 months and OS of 6.9 months, similar to gemcitabine treatment alone (6.8 months) [154]. Combinatorial treatment of tremelimumab with durvalumab or durvalumab monotherapy in PDAC patients yielded similar poor patient outcomes and no effect on disease progression [155]. Moreover, based on pre-clinical evidence showing the potential of Losartan as a stromal modifier able to reduce desmoplasia and enhance the intratumoral penetration and effectiveness of therapeutics in patients with PC [156], phase II clinical trials (NCT01821729, NCT05077800, and NCT03563248) are ongoing for the evaluation of losartan with ICB (nivolumab), chemotherapy (FOLFIRINOX), and radiotherapy in PC patients. Currently, there are various ongoing trials that explore the amalgamation of monoclonal antibodies with other therapies against PDAC (NCT05187338, NCT06353646, NCT05014776, NCT04117087, NCT05088889, NCT03816358, NCT03755739, NCT02834013) (Table 2).
Another immunotherapeutic strategy proposed in recent years for PC treatment is the use of vaccines. A specific anti-tumor immune response may be induced by presenting tumor antigens to the immune system in the form of a tumor-based vaccine. The vaccines clinically pursued in PDAC treatment mainly consist of whole-tumor cells, peptides, proteins, or recombinant constructs [13]. They may contain tumor-associated antigens (TAAs) or mutated tumor-specific antigens (TSAs) or neoantigens [157]. However, their mechanisms of action and efficacy depend on the TMB, which is limited in PDAC [158]. In a phase I trial, a variety of vaccines indicated no long-term survival benefit, although they broke tolerance and generated T cell immunity without any short-term adverse effects [159,160,161]. One of the best-studied therapeutic vaccines in PDAC is GVAX (granulocyte-macrophage colony-stimulating factor (GM-CSF) gene-modified tumor vaccine), which contains irradiated whole pancreatic tumor cells unable to grow that have been genetically modified to secrete GM-CSF. The results, so far, have demonstrated that GVAX is safe and able to induce antigen-specific T cell responses, with or without complementary treatments [4]. The preliminary results of NCT03161379 and NCT02648282 clinical trials showed that, in combination with immunochemoradiotherapy, GVAX is safe and can induce antigen-specific T cell responses [162]. However, GVAX has failed to provide long-term survival benefits and does result in improved treatment efficacy compared to standard-of-care chemotherapy [4]. Various agents, including viruses and bacteria, are currently being used to explore novel mechanisms that expose tumor antigens to the immune system using vaccines [163].
Furthermore, the use of cytokines and chemokines as immunomodulators is another approach in the immunotherapeutic armamentarium against cancer [164]. Agonistic CD40 therapy has been shown to polarize macrophages into a more tumoricidal M1 phenotype, leading to short-term survival benefits [165]. In a pre-clinical model of PDAC, a combination of agonist anti-CD40 with ICB improved survival by inducing T cell immunity and regression of subcutaneous tumors. This treatment combination almost doubled the survival of mice with spontaneous tumors, although they were not cured [13]. The most clinically important target of myeloid cells is C-C motif chemokine receptor 2 (CCR2), which controls the recruitment of inhibitory macrophages in the TME, and it is correlated with poor prognosis [81]. Mitchem et al. demonstrated that CCR2 blockade in PDAC enhances responses to chemotherapy, inhibits metastasis, and blocks monocyte access in the TME, which elevates the infiltration of T cells [166]. It was also shown that CD40 activation along with gemcitabine led to a partial response in PDAC patients, possibly by influencing the immune reaction via TAMs [165]. In a phase I trial, the CCR2 inhibitor PF-04136309 indicated a clinical response in half of PDAC patients pre-treated with FOLFIRINOX [167]. In addition, the pre-clinical analysis of colony-stimulating factor 1 receptor (CSF-1R), a key cytokine regulator of MDSCs and TAMs in PDAC, in combination with checkpoint inhibition using PD1 and CTLA-4, displayed favorable results [168]. Lastly, after CXCR2 inhibition, which is responsible for neutrophil and MDSC migration, T cell infiltration was enhanced. When the CXCR2 blockade was combined with either CSF-1R inhibition or checkpoint blockade, tumor responses were improved [169,170]. Moreover, a dual inhibitor targeting both the Bromo- and extra-terminal Motif (BET) protein BRD4 and histone acetyltransferase EP300/CBP was found to inhibit oncogenic Ras signaling and enhance the efficacy of anti-PD-1 ICB in PC mouse models [171]. Interestingly, a recent study showed that combining anti-4-1BB and anti-LAG3 ICBs with a CXCR1/2 inhibitor targeting myeloid cells can overcome immunotherapy resistance and result in durable therapeutic responses in genetically engineered and syngeneic mouse models for PDAC [172]. On the other hand, in a cohort study of 69 patients with resected PDAC, it was demonstrated that although T cells were associated with prolonged DFS, lymphocyte-activation gene 3 (LAG-3) expression by PDAC-infiltrating T cells was correlated with reduced DFS [173].
Adoptive cell therapies constitute an alternative approach against pancreatic tumors by using autologous or allogeneic immune effector cells, including T cells, DCs, or NKs, to eradicate cancer cells [4]. After being isolated from the patient’s blood or a tumor, these immune cells are genetically modified ex vivo to stimulate an anti-tumor immune response and injected back into the patient [4,30]. Expanding tumor-infiltrating T cells that can migrate from the vasculature into the TME may be beneficial [174]. Immune cells could also be engineered in vitro to replace non-functional lymphocytes. Specifically, chimeric antigen receptor (CAR)-T cells select a cancer cell-specific surface protein as a target to generate effective therapy by removing T cells and reducing competition for stimulatory cytokines [175]. CAR-T cells are typically infused systemically to target malignant cells and exert anti-tumor activity [176]. Several trials for treating solid tumors including PDAC, such as NCT02850536 and NCT02349724, using CAR-T cells have been completed, with preliminary results exhibiting the safety of regionally infused CAR-T cells, but limited biological efficacy [177]. A pre-clinical study in a murine PDAC model demonstrated promising data after treatment with mesothelin-directed CAR-T cells together with oncolytic viruses expressing IL-2 and tumor necrosis factor-α (TNF-α) [178]. Clinically, CAR-T cell therapy has not yet been extensively explored in PDAC. Although CAR-engineered cells show early promise in hematological malignancies, their efficacy in solid tumors is still limited, but is under active investigation [176]. Importantly, this approach is not risk-free, as it may cause immune hyperstimulation, which can be fatal in some cases [179]. A multifaceted approach involving this cellular monotherapy is required in order to identify the appropriate personalized treatment for each PDAC patient.
Table 1. Summary of completed clinical trials in pancreatic cancer immunotherapy.
Table 1. Summary of completed clinical trials in pancreatic cancer immunotherapy.
Drug(s)MechanismPopulationClinical Trial/Year of CompletionPhase of TrialTrial Design/Number of PatientsPrimary Endpoint
CP-870,893
in combination with gemcitabine
CD40 agonist antibody with chemotherapyChemotherapy-naïve, surgically incurable PCNCT00711191
01/2011
ISingle group assignment/22Dose-limiting toxicities (DLTs), adverse events [165]
Ipilimumab (BMS-734016) and PC vaccineCTLA-4 inhibitor with allogeneic pancreatic tumor cells transfected with a GM-CSF geneLocally advanced, unresectable, or metastatic pancreatic adenocarcinomaNCT00836407
07/2012
ITwo-arm trial:
Ibilimumab alone
and Ibilimumab plus PC vaccine/30
Unacceptable toxicity [180]
MEDI4736 in combination with nab-paclitaxel and gemcitabine or with AZD5069Anti-PD-L1 monoclonal antibody with chemotherapy or CXCR2 antagonistMetastatic PDACNCT02583477
07/2018
I/IIParallel assignment/23Adverse events, DLTs, ORR
TG01/GM-CSF with gemcitabineTG01 and granulocyte macrophage colony-stimulating factor with chemotherapyResected adenocarcinoma of the pancreasNCT02261714
05/2019
I/IISingle group assignment/32Immune responses, adverse events [181]
Ipilimumab, vaccine, FOLFIRINOXCTLA-4 inhibitor with allogeneic GM-CSF-transfected pancreatic tumor vaccine against chemotherapyMetastatic PCNCT01896869
05/2019
IITwo-arm trial:
Ibilimumab plus vaccine or FOLFIRINOX alone/83
Overall survival (OS)
FOLFIRINOX, Losartan, Proton Beam RadiationCombination of 5-fluorouracil, leucovorin and oxaliplatin with proton beam therapyLocally advanced diseaseNCT01821729
09/2021
IISingle group assignment/50Number of participants with R0 resection
Cyclophosphamide, GVAX PC vaccine, Pembrolizumab, radiationChemotherapy, GM-CSF-secreting allogeneic PC vaccine with PD-1 blockade antibody and stereotactic body radiation therapyLocally advanced adenocarcinoma of the pancreasNCT02648282
01/2022
IISingle group assignment/58Distant metastasis-free survival
Oleclumab (MEDI9447), Durvalumab, Gemcitabine, Nab-paclitaxel, Oxaliplatin, Folinic acid, 5-FUAnti-CD73 monoclonal antibody, anti-PD-L1 monoclonal antibody in combination with chemotherapyMetastatic PDACNCT03611556
07/2022
I/IIParallel assignment/213TEAEs, TESAEs, DLTs, abnormal vital signs, electrocardiogram and clinical laboratory parameters, ORR
Plerixafor and CemiplimabCXCR4 inhibitor (small molecule) and PD-1 blocking antibodyMetastatic PCNCT04177810
05/2023
IISingle group assignment/25ORR
Cyclophosphamide, Nivolumab, Ipilimumab, GVAX Pancreas Vaccine, CRS-207Chemotherapy plus immunotherapy (anti-PD-1 and anti-CTLA-4) plus PC vaccinePreviously treated metastatic pancreatic adenocarcinomaNCT03190265
08/2023
IIParallel assignment with or without PC vaccine/61Objective response rate (ORR)
Table 2. Summary of ongoing clinical trials in pancreatic cancer immunotherapy.
Table 2. Summary of ongoing clinical trials in pancreatic cancer immunotherapy.
Drug(s)MechanismType of DiseaseIDPhaseTrial Design/Number of PatientsPrimary Endpoint
Ipilimumab in combination with pembrolizumab and durvalumabCombination of three antibodies against PD1, PDL1, and CTLA4Advanced solid tumorsNCT05187338I/IISingle group assignment/100
(estimated)
Drug safety, progression-free survival (PFS), disease control rate, duration of remission
XH001 combination with Ipilimumab and gemcitabine /capecitabinemRNA neoantigen cancer vaccine in combination with
immunotherapy that targets CTLA-4 protein on T cells plus chemotherapy
Resected PCNCT06353646N/ASingle-center, open-label, single-arm/12
(estimated)
Efficacy and safety trial
Tadalafil, Pembrolizumab, Ipilimumab, CRS-207PD-1/CTLA-4-blocking antibodies in combination with immunotherapyPreviously treated metastatic PDACNCT05014776IISingle group assignment/17 (actual)Objective response rate (irORR) using immune response evaluation criteria for solid tumors (iRECIST)
KRAS peptide vaccine
Nivolumab
Ipilimumab
KRAS peptide vaccine with poly-ICLC adjuvant in combination with immunotherapyResected mismatch repair protein (MMR-p), Colorectal and PCNCT04117087IParallel assignment/30 (estimated)Number of participants experiencing study drug-related toxicities
Fold change in interferon-producing mutant-KRAS-specific cytotoxic (CD8) and helper (CD4) T cells at 16 weeks
Nivolumab
Ipilimumab
Stereotactic body radiation therapy
Low-dose irradiation
Cytotoxic chemotherapy followed by hypofractionated radiotherapy to sensitize pancreatic cancer to immunotherapy consisting of combined PD-1 and CTLA4 blockadeFirst-line treatment stage IV pancreatic cancerNCT05088889IStudy arm/10 (estimated)Objective tumor response rate 1 (ORR1) in study patients, assessed by RECIST v1.1
Objective tumor response rate 2 (ORR2) after first progression, assessed by RECIST v1.1
Combination of anetumab ravtansine with Either nivolumab, nivolumab and ipilimumab, or gemcitabine and nivolumabMonoclonal antibody linked to a chemotherapy drug called DM4 attaches to mesothelin-positive cancer cells and delivers DM4 to kill them. Immunotherapy chemotherapyAdvanced PCNCT03816358IParallel assignment/ 74 (estimated)Maximum tolerated dose (MTD)
Checkpoint inhibitor (CPI) such as Pembrolizumab plus chemotherapyTrans-artery/intra-tumor infusion of PD1/PDL1 antibody and/or CTLA4 antibody ipilimumab plus chemotherapeutic drug and comparison of their differences.Advanced solid tumorsNCT03755739II/IIIParallel assignment/200 (estimated)Overall survival
Complete response (CR) rate before or at month 6
Nivolumab and IpilimumabImmunotherapy with monoclonal antibodiesRare pancreatic tumors including acinar cell carcinoma, mucinous cystadenocarcinoma, or serous cystadenocarcinomaNCT02834013IIParallel assignment/818 (estimated)ORR
9-ING-41
Losartan
Ferumoxytol
FOLFIRINOX
Blocking of GSK-3β activity using 9-ING-41 and blocking of TGF-β function using Losartan to inhibit cancer cell resistance to FOLFIRINOX chemotherapyMetastatic PDAC without prior therapyNCT05077800IISingle group assignment/70 (estimated)PFS
Losartan and nivolumab in Combination with FOLFIRINOX and SBRTCombination chemotherapy. Losartan is used to lower blood pressure. Nivolumab is an antibody that may cause apoptosis. Radiation by stereotactic body radiation therapyLocalized PDAC; borderline/potentially resectable or locally advanced.NCT03563248IIParallel assignment/168 (actual)Proportion of participants with R0 resection
Motixafortide
Cemiplimab Gemcitabine
Nab-paclitaxel
Combination chemotherapy (gemcitabine and nab-paclitaxel), chemokine (C-X-C) motif receptor 4 inhibitor (BL-8040), and immune checkpoint blockade (Cemiplimab)Metastatic treatment-naïve PDACNCT04543071IISingle group assignment/10 (estimated)Overall response rate (complete response (CR) + partial response (PR))
Durvalumab
Rintatolimod
Combination therapy of ICI therapy with a toll-like receptor 3 (TLR-3) agonistMetastatic PCNCT05927142I/IISequential assignment/43 (estimated)Phase Ib: Determine safety of combination therapy with durvalumab and rintatolimod
Phase II: Determine the clinical benefit rate of combination therapy with durvalumab and rintatolimod.
PT199
Tislelizumab
PT199 counters the adenosine-mediated immunosuppressive TME, rendering anti-tumor immune cells more responsive to checkpoint immunotherapies, such as PD-1/PD-L1 inhibitorsAdvanced solid tumorsNCT05431270ISequential assignment/40 (estimated)MTD,
DLT,
Safety of PT199
IMM-101, Pembrolizumab, GemcitabineCombination of a heat-inactivated mycobacterium, immune modulator with chemotherapyMetastatic PCNCT06498518IISingle group assignment/50 (estimated)ORR
Ipilimumab
Nivolumab
Radiation Therapy
Ipilimumab inhibits cancer cell growth. Nivolumab induces apoptosis. Radiation therapy may increase the likelihood of response to interventions.Metastatic, microsatellite-stable PCNCT04361162IISingle group assignment/30 (estimated)ORR
Pembrolizumab With OlaparibPD-1 inhibitor in combination with PARP inhibitorMetastatic PDAC with mismatch repair deficiency or tumor mutation burden > 4 mutations/MbNCT05093231IISingle group assignment/20 (estimated)ORR
Nivolumab
Ipilimumab
SBRT
TGFβ-B-15 peptide
Combination of checkpoint-blocking antibodies with immunomodulation of the TME; TGFβ-15 immune response is correlated to clinical benefitRefractory PCNCT05721846ISingle group assignment/20 (estimated)Adverse effects
Pembrolizumab
Folfirinox
SABR
Combination of chemotherapy, PD-1 inhibition, and radiotherapyBorderline resectable PCNCT06384560I/IISingle group assignment/66 (estimated)Percentage of patients with progression free survival at 18 months (RECIST 1.1)
Lenvatinib
Pembrolizumab
Inhibition of cell growth, PD-1 inhibitionAdvanced unresectable PCNCT04887805IISingle group assignment/28 (estimated)PFS
Pembrolizumab With Olaparib/Olaparib alonePD-1 inhibitor in combination with PARP inhibitorAdvanced PC with germline BRCA1 or BRCA2 mutationsNCT04548752IIParallel assignment /88 (estimated)PFS
Pembrolizumab
Defactinib
Reprogramming the TME by targeting FAK following chemotherapy to potentiate anti-PD-1 antibodyResectable PDACNCT03727880IIParallel assignment /36 (estimated)Pathologic complete response (pCR) rate
SBRT
Nivolumab
CCR2/CCR5 dual antagonist
GVAX
Combination therapy to enhance the infiltration of CD8+CD137+ cells in PDACLocally advanced PDACNCT03767582I/IISequential assignment /30 (estimated)Number of Participants experiencing study drug-related toxicities
Percentage of participants treated with immunotherapy who achieve an immune response
Nivolumab
Ipilimumab
SBRT
Low dose irradiation
Cytotoxic chemotherapy followed by hypofractionated radiotherapy to sensitize PC to combined PD-1 and CTLA4 blockadeStage IV PDACNCT05088889ISingle group assignment/10 (estimated)Objective tumor response rate 1 (ORR1) in study patients, assessed by RECIST v1.1
Objective tumor response rate 2 (ORR2) after first progression, assessed by RECIST v1.1
Nivolumab
Ipilimumab
Hydroxychloroquine (HCQ)
Immunotherapy in combination with standard-of-care chemotherapyPreviously untreated PDACNCT04787991IParallel assignment/45 (estimated)Incidence and severity of adverse events
Tumor-Infiltrating Lymphocytes
Pembrolizumab
Autologous tumor-infiltrating lymphocytes following a lymphodepleting regimen plus PD-1 inhibitionMetastatic PCNCT01174121IIParallel assignment/332 (estimated)Response rate
Epacadostat
Pembrolizumab
CRS-207
IDO-1 inhibition, PD-1 inhibition, stimulation of immune response to mesothelinMetastatic PCNCT03006302IIParallel assignment/ 41 (estimated)MTD
6-month survival
Atezolizumab
Tivozanib
Inhibition of VEGF, inhibition of PD-1Immunogenically cold PDACNCT05000294I/IISequential assignment/29 (estimated)ORR
Gemcitabine, Nab-paclitaxel, Capecitabine, Cisplatin, Irinotecan Olaparib and PembrolizumabLow-dose chemotherapy followed by maintenance with PD-1 and PARP inhibitionMetastatic untreated PDACNCT04753879IISingle group assignment/38 (estimated)PFS
mFOLFIRINOX
Verteporfin Pembrolizumab
Photoradiation with verteporfin, PD-1 inhibition plus standard-of-care chemotherapyLocally Advanced or Metastatic PCNCT06381154IISingle group assignment /25 (estimated)ORR
Nivolumab
Ipilimumab
Radiation Therapy
Blocking PD-1/PD-L1 and CTLA-4 pathwaysPCNCT03104439IISingle group assignment /80 (estimated)Disease control rate
Gemcitabine
Nab-paclitaxel
Nivolumab
ICIs with chemotherapyBorderline resectable, locally advanced or mPCNCT03970252I/IISingle group assignment /40 (estimated)Incidence of treatment-related AEs, SAEs, AEs leading to discontinuation, death, and laboratory abnormalities
Futibatinib
Pembrolizumab
Cisplatin
ICIs with chemotherapyPDACNCT05945823IIParallel group assignment /40 (estimated)ORR
LGK974
PDR001
PORCN inhibitor
PD-1 inhibitor
PC that has progressed despite standard therapy or for which no effective standard therapy existsNCT01351103ISingle group assignment/185 (actual)Maximum tolerated dose or recommended dose for expansion of LGK974 as a single agent or in combination with PDR001 in treated patients

5. Conclusions

Pancreatic cancer remains one of the deadliest diseases, with a dismal prognosis. Currently, prevention or early diagnosis at a curable stage is extremely difficult. Failure of approved PDAC therapies has dramatic effects on the quality of life of cancer patients. One of the emerging interests in PDAC therapy is immunotherapy due to the capability of cancer cells to escape immune surveillance through various mechanisms, including their densely packed TME that is depleted of effector T cells. Numerous evidence suggests the importance of unraveling the complexity of TME components and their multifaceted interactions in tumor suppression or progression. Findings from clinical trials encompassing immunotherapeutic strategies have not been encouraging regarding a cure for PDAC. Moreover, some treatments have shown partial effectiveness against PDAC, such as FOLFIRINOX, but most of them have failed to provide a significant survival improvement without serious side effects. Hence, future treatment approaches should aim to comprise therapies that target multiple characteristics of the TME simultaneously. To this end, ongoing efforts to evaluate the efficacy of remodeling the pancreatic TME using multimodal strategies might be proven beneficial for a subset of PDAC patients to provide an increased survival rate and/or better quality of life. However, an increased number of more sizeable studies are needed to clarify the effects of different treatment options and the optimal therapeutic protocols to combine them. Various trials have failed, but the progress in our knowledge about pancreatic TME gives substantial hope for the future development of successful therapies. An improved understanding of the crosstalk between tumor, stromal, and immune cells may soon lead to discoveries able to reverse the innate resistance of PDAC to immunotherapies.

Author Contributions

Conceptualization, P.P. writing—original draft preparation, F.P., C.M.N., M.-I.C. and P.P.; writing—review and editing, F.P., C.M.N., M.-I.C. and P.P.; visualization, F.P and P.P.; supervision, P.P.; project administration, P.P.; funding acquisition, P.P. All authors have read and agreed to the published version of the manuscript.

Funding

The project EXCELLENCE/0421/0179 is co-financed by the European Regional Development Fund and the Republic of Cyprus through the Research and Innovation Foundation.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created as part of this review paper.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Vincent, A.; Herman, J.; Schulick, R.; Hruban, R.H.; Goggins, M. Pancreatic cancer. Lancet 2011, 378, 607–620. [Google Scholar] [CrossRef]
  2. Rawla, P.; Sunkara, T.; Gaduputi, V. Epidemiology of Pancreatic Cancer: Global Trends, Etiology and Risk Factors. World J. Oncol. 2019, 10, 10–27. [Google Scholar] [CrossRef]
  3. Rahib, L.; Smith, B.D.; Aizenberg, R.; Rosenzweig, A.B.; Fleshman, J.M.; Matrisian, L.M. Projecting cancer incidence and deaths to 2030: The unexpected burden of thyroid, liver, and pancreas cancers in the United States. Cancer Res. 2014, 74, 2913–2921. [Google Scholar] [CrossRef] [PubMed]
  4. Timmer, F.E.F.; Geboers, B.; Nieuwenhuizen, S.; Dijkstra, M.; Schouten, E.A.C.; Puijk, R.S.; de Vries, J.J.J.; van den Tol, M.P.; Bruynzeel, A.M.E.; Streppel, M.M.; et al. Pancreatic Cancer and Immunotherapy: A Clinical Overview. Cancers 2021, 13, 4138. [Google Scholar] [CrossRef]
  5. Siegel, R.L.; Giaquinto, A.N.; Jemal, A. Cancer statistics, 2024. CA Cancer J. Clin. 2024, 74, 12–49. [Google Scholar] [CrossRef]
  6. Strobel, O.; Neoptolemos, J.; Jager, D.; Buchler, M.W. Optimizing the outcomes of pancreatic cancer surgery. Nat. Rev. Clin. Oncol. 2019, 16, 11–26. [Google Scholar] [CrossRef]
  7. Wolfgang, C.L.; Herman, J.M.; Laheru, D.A.; Klein, A.P.; Erdek, M.A.; Fishman, E.K.; Hruban, R.H. Recent progress in pancreatic cancer. CA Cancer J. Clin. 2013, 63, 318–348. [Google Scholar] [CrossRef]
  8. Jones, S.; Zhang, X.; Parsons, D.W.; Lin, J.C.; Leary, R.J.; Angenendt, P.; Mankoo, P.; Carter, H.; Kamiyama, H.; Jimeno, A.; et al. Core signaling pathways in human pancreatic cancers revealed by global genomic analyses. Science 2008, 321, 1801–1806. [Google Scholar] [CrossRef] [PubMed]
  9. Kleeff, J.; Korc, M.; Apte, M.; La Vecchia, C.; Johnson, C.D.; Biankin, A.V.; Neale, R.E.; Tempero, M.; Tuveson, D.A.; Hruban, R.H.; et al. Pancreatic cancer. Nat. Rev. Dis. Primers 2016, 2, 16022. [Google Scholar] [CrossRef] [PubMed]
  10. Conroy, T.; Pfeiffer, P.; Vilgrain, V.; Lamarca, A.; Seufferlein, T.; O’Reilly, E.M.; Hackert, T.; Golan, T.; Prager, G.; Haustermans, K.; et al. Pancreatic cancer: ESMO Clinical Practice Guideline for diagnosis, treatment and follow-up. Ann. Oncol. 2023, 34, 987–1002. [Google Scholar] [CrossRef]
  11. Petrelli, F.; Coinu, A.; Borgonovo, K.; Cabiddu, M.; Ghilardi, M.; Lonati, V.; Aitini, E.; Barni, S.; Gruppo Italiano per lo Studio dei Carcinomi dell’Apparato Digerente (GISCAD). FOLFIRINOX-based neoadjuvant therapy in borderline resectable or unresectable pancreatic cancer: A meta-analytical review of published studies. Pancreas 2015, 44, 515–521. [Google Scholar] [CrossRef] [PubMed]
  12. Lopez, N.E.; Prendergast, C.; Lowy, A.M. Borderline resectable pancreatic cancer: Definitions and management. World J. Gastroenterol. 2014, 20, 10740–10751. [Google Scholar] [CrossRef]
  13. Robatel, S.; Schenk, M. Current Limitations and Novel Perspectives in Pancreatic Cancer Treatment. Cancers 2022, 14, 985. [Google Scholar] [CrossRef] [PubMed]
  14. Cid-Arregui, A.; Juarez, V. Perspectives in the treatment of pancreatic adenocarcinoma. World J. Gastroenterol. 2015, 21, 9297–9316. [Google Scholar] [CrossRef] [PubMed]
  15. Springfeld, C.; Jager, D.; Buchler, M.W.; Strobel, O.; Hackert, T.; Palmer, D.H.; Neoptolemos, J.P. Chemotherapy for pancreatic cancer. Presse Med. 2019, 48 Pt 2, e159–e174. [Google Scholar] [CrossRef]
  16. Conroy, T.; Desseigne, F.; Ychou, M.; Bouche, O.; Guimbaud, R.; Becouarn, Y.; Adenis, A.; Raoul, J.L.; Gourgou-Bourgade, S.; de la Fouchardiere, C.; et al. FOLFIRINOX versus gemcitabine for metastatic pancreatic cancer. N. Engl. J. Med. 2011, 364, 1817–1825. [Google Scholar] [CrossRef]
  17. Von Hoff, D.D.; Ervin, T.; Arena, F.P.; Chiorean, E.G.; Infante, J.; Moore, M.; Seay, T.; Tjulandin, S.A.; Ma, W.W.; Saleh, M.N.; et al. Increased survival in pancreatic cancer with nab-paclitaxel plus gemcitabine. N. Engl. J. Med. 2013, 369, 1691–1703. [Google Scholar] [CrossRef]
  18. Neoptolemos, J.P.; Palmer, D.H.; Ghaneh, P.; Psarelli, E.E.; Valle, J.W.; Halloran, C.M.; Faluyi, O.; O’Reilly, D.A.; Cunningham, D.; Wadsley, J.; et al. Comparison of adjuvant gemcitabine and capecitabine with gemcitabine monotherapy in patients with resected pancreatic cancer (ESPAC-4): A multicentre, open-label, randomised, phase 3 trial. Lancet 2017, 389, 1011–1024. [Google Scholar] [CrossRef]
  19. Conroy, T.; Hammel, P.; Hebbar, M.; Ben Abdelghani, M.; Wei, A.C.; Raoul, J.L.; Chone, L.; Francois, E.; Artru, P.; Biagi, J.J.; et al. FOLFIRINOX or Gemcitabine as Adjuvant Therapy for Pancreatic Cancer. N. Engl. J. Med. 2018, 379, 2395–2406. [Google Scholar] [CrossRef]
  20. Kang, J.; Hwang, I.; Yoo, C.; Kim, K.P.; Jeong, J.H.; Chang, H.M.; Lee, S.S.; Park, D.H.; Song, T.J.; Seo, D.W.; et al. Nab-paclitaxel plus gemcitabine versus FOLFIRINOX as the first-line chemotherapy for patients with metastatic pancreatic cancer: Retrospective analysis. Investig. New Drugs 2018, 36, 732–741. [Google Scholar] [CrossRef]
  21. McGuigan, A.; Kelly, P.; Turkington, R.C.; Jones, C.; Coleman, H.G.; McCain, R.S. Pancreatic cancer: A review of clinical diagnosis, epidemiology, treatment and outcomes. World J. Gastroenterol. 2018, 24, 4846–4861. [Google Scholar] [CrossRef]
  22. Tempero, M.A.; Malafa, M.P.; Al-Hawary, M.; Behrman, S.W.; Benson, A.B.; Cardin, D.B.; Chiorean, E.G.; Chung, V.; Czito, B.; Del Chiaro, M.; et al. Pancreatic Adenocarcinoma, Version 2.2021, NCCN Clinical Practice Guidelines in Oncology. J. Natl. Compr. Cancer Netw. 2021, 19, 439–457. [Google Scholar] [CrossRef] [PubMed]
  23. Loehrer, P.J., Sr.; Feng, Y.; Cardenes, H.; Wagner, L.; Brell, J.M.; Cella, D.; Flynn, P.; Ramanathan, R.K.; Crane, C.H.; Alberts, S.R.; et al. Gemcitabine alone versus gemcitabine plus radiotherapy in patients with locally advanced pancreatic cancer: An Eastern Cooperative Oncology Group trial. J. Clin. Oncol. 2011, 29, 4105–4112. [Google Scholar] [CrossRef]
  24. Shelemey, P.T.; Amaro, C.P.; Ng, D.; Falck, V.; Tam, V.C. Metastatic pancreatic cancer with complete response to FOLFIRINOX treatment. BMJ Case Rep. 2021, 14, e238395. [Google Scholar] [CrossRef]
  25. Pekarek, L.; Fraile-Martinez, O.; Garcia-Montero, C.; Alvarez-Mon, M.A.; Acero, J.; Ruiz-Llorente, L.; Garcia-Honduvilla, N.; Albillos, A.; Bujan, J.; Alvarez-Mon, M.; et al. Towards an updated view on the clinical management of pancreatic adenocarcinoma: Current and future perspectives. Oncol. Lett. 2021, 22, 809. [Google Scholar] [CrossRef] [PubMed]
  26. Shah, A.P.; Strauss, J.B.; Abrams, R.A. Review and commentary on the role of radiation therapy in the adjuvant management of pancreatic cancer. Am. J. Clin. Oncol. 2010, 33, 101–106. [Google Scholar] [CrossRef] [PubMed]
  27. Liermann, J.; Shinoto, M.; Syed, M.; Debus, J.; Herfarth, K.; Naumann, P. Carbon ion radiotherapy in pancreatic cancer: A review of clinical data. Radiother. Oncol. 2020, 147, 145–150. [Google Scholar] [CrossRef]
  28. Hong, T.S.; Ryan, D.P.; Borger, D.R.; Blaszkowsky, L.S.; Yeap, B.Y.; Ancukiewicz, M.; Deshpande, V.; Shinagare, S.; Wo, J.Y.; Boucher, Y.; et al. A phase 1/2 and biomarker study of preoperative short course chemoradiation with proton beam therapy and capecitabine followed by early surgery for resectable pancreatic ductal adenocarcinoma. Int. J. Radiat. Oncol. Biol. Phys. 2014, 89, 830–838. [Google Scholar] [CrossRef]
  29. Hong, T.S.; Ryan, D.P.; Blaszkowsky, L.S.; Mamon, H.J.; Kwak, E.L.; Mino-Kenudson, M.; Adams, J.; Yeap, B.; Winrich, B.; DeLaney, T.F.; et al. Phase I study of preoperative short-course chemoradiation with proton beam therapy and capecitabine for resectable pancreatic ductal adenocarcinoma of the head. Int. J. Radiat. Oncol. Biol. Phys. 2011, 79, 151–157. [Google Scholar] [CrossRef]
  30. Neoptolemos, J.P.; Kleeff, J.; Michl, P.; Costello, E.; Greenhalf, W.; Palmer, D.H. Therapeutic developments in pancreatic cancer: Current and future perspectives. Nat. Rev. Gastroenterol. Hepatol. 2018, 15, 333–348. [Google Scholar] [CrossRef]
  31. Michl, P.; Gress, T.M. Current concepts and novel targets in advanced pancreatic cancer. Gut 2013, 62, 317–326. [Google Scholar] [CrossRef] [PubMed]
  32. Leroux, C.; Konstantinidou, G. Targeted Therapies for Pancreatic Cancer: Overview of Current Treatments and New Opportunities for Personalized Oncology. Cancers 2021, 13, 799. [Google Scholar] [CrossRef] [PubMed]
  33. Kindler, H.L.; Niedzwiecki, D.; Hollis, D.; Sutherland, S.; Schrag, D.; Hurwitz, H.; Innocenti, F.; Mulcahy, M.F.; O’Reilly, E.; Wozniak, T.F.; et al. Gemcitabine plus bevacizumab compared with gemcitabine plus placebo in patients with advanced pancreatic cancer: Phase III trial of the Cancer and Leukemia Group B (CALGB 80303). J. Clin. Oncol. 2010, 28, 3617–3622. [Google Scholar] [CrossRef]
  34. Van Cutsem, E.; Vervenne, W.L.; Bennouna, J.; Humblet, Y.; Gill, S.; Van Laethem, J.L.; Verslype, C.; Scheithauer, W.; Shang, A.; Cosaert, J.; et al. Phase III trial of bevacizumab in combination with gemcitabine and erlotinib in patients with metastatic pancreatic cancer. J. Clin. Oncol. 2009, 27, 2231–2237. [Google Scholar] [CrossRef] [PubMed]
  35. Kindler, H.L.; Ioka, T.; Richel, D.J.; Bennouna, J.; Letourneau, R.; Okusaka, T.; Funakoshi, A.; Furuse, J.; Park, Y.S.; Ohkawa, S.; et al. Axitinib plus gemcitabine versus placebo plus gemcitabine in patients with advanced pancreatic adenocarcinoma: A double-blind randomised phase 3 study. Lancet Oncol. 2011, 12, 256–262. [Google Scholar] [CrossRef] [PubMed]
  36. Goncalves, A.; Gilabert, M.; Francois, E.; Dahan, L.; Perrier, H.; Lamy, R.; Re, D.; Largillier, R.; Gasmi, M.; Tchiknavorian, X.; et al. BAYPAN study: A double-blind phase III randomized trial comparing gemcitabine plus sorafenib and gemcitabine plus placebo in patients with advanced pancreatic cancer. Ann. Oncol. Off. J. Eur. Soc. Med. Oncol. 2012, 23, 2799–2805. [Google Scholar] [CrossRef]
  37. Moore, M.J.; Goldstein, D.; Hamm, J.; Figer, A.; Hecht, J.R.; Gallinger, S.; Au, H.J.; Murawa, P.; Walde, D.; Wolff, R.A.; et al. Erlotinib plus gemcitabine compared with gemcitabine alone in patients with advanced pancreatic cancer: A phase III trial of the National Cancer Institute of Canada Clinical Trials Group. J. Clin. Oncol. 2007, 25, 1960–1966. [Google Scholar] [CrossRef]
  38. Marzoq, A.J.; Mustafa, S.A.; Heidrich, L.; Hoheisel, J.D.; Alhamdani, M.S.S. Impact of the secretome of activated pancreatic stellate cells on growth and differentiation of pancreatic tumour cells. Sci. Rep. 2019, 9, 5303. [Google Scholar] [CrossRef]
  39. Schiarea, S.; Solinas, G.; Allavena, P.; Scigliuolo, G.M.; Bagnati, R.; Fanelli, R.; Chiabrando, C. Secretome analysis of multiple pancreatic cancer cell lines reveals perturbations of key functional networks. J. Proteome Res. 2010, 9, 4376–4392. [Google Scholar] [CrossRef]
  40. de Oliveira, G.; Freire, P.P.; Cury, S.S.; de Moraes, D.; Oliveira, J.S.; Dal-Pai-Silva, M.; Reis, P.P.; Carvalho, R.F. An Integrated Meta-Analysis of Secretome and Proteome Identify Potential Biomarkers of Pancreatic Ductal Adenocarcinoma. Cancers 2020, 12, 716. [Google Scholar] [CrossRef]
  41. Whatcott, C.J.; Diep, C.H.; Jiang, P.; Watanabe, A.; LoBello, J.; Sima, C.; Hostetter, G.; Shepard, H.M.; Von Hoff, D.D.; Han, H. Desmoplasia in Primary Tumors and Metastatic Lesions of Pancreatic Cancer. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2015, 21, 3561–3568. [Google Scholar] [CrossRef] [PubMed]
  42. Ohlund, D.; Lundin, C.; Ardnor, B.; Oman, M.; Naredi, P.; Sund, M. Type IV collagen is a tumour stroma-derived biomarker for pancreas cancer. Br. J. Cancer 2009, 101, 91–97. [Google Scholar] [CrossRef] [PubMed]
  43. Jain, R.K.; Martin, J.D.; Stylianopoulos, T. The role of mechanical forces in tumor growth and therapy. Annu. Rev. Biomed. Eng. 2014, 16, 321–346. [Google Scholar] [CrossRef]
  44. Chen, I.X.; Chauhan, V.P.; Posada, J.; Ng, M.R.; Wu, M.W.; Adstamongkonkul, P.; Huang, P.; Lindeman, N.; Langer, R.; Jain, R.K. Blocking CXCR4 alleviates desmoplasia, increases T-lymphocyte infiltration, and improves immunotherapy in metastatic breast cancer. Proc. Natl. Acad. Sci. USA 2019, 116, 4558–4566. [Google Scholar] [CrossRef]
  45. Ferrara, B.; Pignatelli, C.; Cossutta, M.; Citro, A.; Courty, J.; Piemonti, L. The Extracellular Matrix in Pancreatic Cancer: Description of a Complex Network and Promising Therapeutic Options. Cancers 2021, 13, 4442. [Google Scholar] [CrossRef] [PubMed]
  46. Hosein, A.N.; Brekken, R.A.; Maitra, A. Pancreatic cancer stroma: An update on therapeutic targeting strategies. Nat. Rev. Gastroenterol. Hepatol. 2020, 17, 487–505. [Google Scholar] [CrossRef] [PubMed]
  47. Whatcott, C.J.; Posner, R.G.; Von Hoff, D.D.; Han, H. Desmoplasia and chemoresistance in pancreatic cancer. In Pancreatic Cancer and Tumor Microenvironment; Grippo, P.J., Munshi, H.G., Eds.; Transworld Research Network: Trivandrum, India, 2012. [Google Scholar]
  48. Martin, J.D.; Seano, G.; Jain, R.K. Normalizing Function of Tumor Vessels: Progress, Opportunities, and Challenges. Annu. Rev. Physiol. 2019, 81, 505–534. [Google Scholar] [CrossRef]
  49. Diop-Frimpong, B.; Chauhan, V.P.; Krane, S.; Boucher, Y.; Jain, R.K. Losartan inhibits collagen I synthesis and improves the distribution and efficacy of nanotherapeutics in tumors. Proc. Natl. Acad. Sci. USA 2011, 108, 2909–2914. [Google Scholar] [CrossRef]
  50. Chauhan, V.P.; Martin, J.D.; Liu, H.; Lacorre, D.A.; Jain, S.R.; Kozin, S.V.; Stylianopoulos, T.; Mousa, A.S.; Han, X.; Adstamongkonkul, P.; et al. Angiotensin inhibition enhances drug delivery and potentiates chemotherapy by decompressing tumour blood vessels. Nat. Commun. 2013, 4, 2516. [Google Scholar] [CrossRef]
  51. Kasi, A.; Allen, J.; Mehta, K.; Dandawate, P.; Saha, S.; Bossmann, S.; Anant, S.; Sun, W. Association of losartan with outcomes in metastatic pancreatic cancer patients treated with chemotherapy. J. Clin. Transl. Res. 2021, 7, 257–262. [Google Scholar]
  52. Ohlund, D.; Handly-Santana, A.; Biffi, G.; Elyada, E.; Almeida, A.S.; Ponz-Sarvise, M.; Corbo, V.; Oni, T.E.; Hearn, S.A.; Lee, E.J.; et al. Distinct populations of inflammatory fibroblasts and myofibroblasts in pancreatic cancer. J. Exp. Med. 2017, 214, 579–596. [Google Scholar] [CrossRef]
  53. Zhang, T.; Ren, Y.; Yang, P.; Wang, J.; Zhou, H. Cancer-associated fibroblasts in pancreatic ductal adenocarcinoma. Cell Death Dis. 2022, 13, 897. [Google Scholar] [CrossRef] [PubMed]
  54. Gorchs, L.; Kaipe, H. Interactions between Cancer-Associated Fibroblasts and T Cells in the Pancreatic Tumor Microenvironment and the Role of Chemokines. Cancers 2021, 13, 2995. [Google Scholar] [CrossRef] [PubMed]
  55. Masamune, A.; Shimosegawa, T. Signal transduction in pancreatic stellate cells. J. Gastroenterol. 2009, 44, 249–260. [Google Scholar] [CrossRef] [PubMed]
  56. Thomas, D.; Radhakrishnan, P. Tumor-stromal crosstalk in pancreatic cancer and tissue fibrosis. Mol. Cancer 2019, 18, 14. [Google Scholar] [CrossRef]
  57. Norton, J.; Foster, D.; Chinta, M.; Titan, A.; Longaker, M. Pancreatic Cancer Associated Fibroblasts (CAF): Under-Explored Target for Pancreatic Cancer Treatment. Cancers 2020, 12, 1347. [Google Scholar] [CrossRef]
  58. Ren, B.; Cui, M.; Yang, G.; Wang, H.; Feng, M.; You, L.; Zhao, Y. Tumor microenvironment participates in metastasis of pancreatic cancer. Mol. Cancer 2018, 17, 108. [Google Scholar] [CrossRef]
  59. Hwang, R.F.; Moore, T.; Arumugam, T.; Ramachandran, V.; Amos, K.D.; Rivera, A.; Ji, B.; Evans, D.B.; Logsdon, C.D. Cancer-associated stromal fibroblasts promote pancreatic tumor progression. Cancer Res. 2008, 68, 918–926. [Google Scholar] [CrossRef]
  60. Ene-Obong, A.; Clear, A.J.; Watt, J.; Wang, J.; Fatah, R.; Riches, J.C.; Marshall, J.F.; Chin-Aleong, J.; Chelala, C.; Gribben, J.G.; et al. Activated pancreatic stellate cells sequester CD8+ T cells to reduce their infiltration of the juxtatumoral compartment of pancreatic ductal adenocarcinoma. Gastroenterology 2013, 145, 1121–1132. [Google Scholar] [CrossRef]
  61. Provenzano, P.P.; Cuevas, C.; Chang, A.E.; Goel, V.K.; Von Hoff, D.D.; Hingorani, S.R. Enzymatic targeting of the stroma ablates physical barriers to treatment of pancreatic ductal adenocarcinoma. Cancer Cell 2012, 21, 418–429. [Google Scholar] [CrossRef]
  62. Feig, C.; Gopinathan, A.; Neesse, A.; Chan, D.S.; Cook, N.; Tuveson, D.A. The pancreas cancer microenvironment. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2012, 18, 4266–4276. [Google Scholar] [CrossRef]
  63. Ozdemir, B.C.; Pentcheva-Hoang, T.; Carstens, J.L.; Zheng, X.; Wu, C.C.; Simpson, T.R.; Laklai, H.; Sugimoto, H.; Kahlert, C.; Novitskiy, S.V.; et al. Depletion of carcinoma-associated fibroblasts and fibrosis induces immunosuppression and accelerates pancreas cancer with reduced survival. Cancer Cell 2014, 25, 719–734. [Google Scholar] [CrossRef]
  64. Rhim, A.D.; Oberstein, P.E.; Thomas, D.H.; Mirek, E.T.; Palermo, C.F.; Sastra, S.A.; Dekleva, E.N.; Saunders, T.; Becerra, C.P.; Tattersall, I.W.; et al. Stromal elements act to restrain, rather than support, pancreatic ductal adenocarcinoma. Cancer Cell 2014, 25, 735–747. [Google Scholar] [CrossRef] [PubMed]
  65. Roshani, R.; McCarthy, F.; Hagemann, T. Inflammatory cytokines in human pancreatic cancer. Cancer Lett. 2014, 345, 157–163. [Google Scholar] [CrossRef] [PubMed]
  66. Yao, W.; Maitra, A.; Ying, H. Recent insights into the biology of pancreatic cancer. EBioMedicine 2020, 53, 102655. [Google Scholar] [CrossRef] [PubMed]
  67. Saka, D.; Gokalp, M.; Piyade, B.; Cevik, N.C.; Arik Sever, E.; Unutmaz, D.; Ceyhan, G.O.; Demir, I.E.; Asimgil, H. Mechanisms of T-Cell Exhaustion in Pancreatic Cancer. Cancers 2020, 12, 2274. [Google Scholar] [CrossRef]
  68. Kennel, K.B.; Bozlar, M.; De Valk, A.F.; Greten, F.R. Cancer-Associated Fibroblasts in Inflammation and Antitumor Immunity. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2023, 29, 1009–1016. [Google Scholar] [CrossRef]
  69. Wu, J.; Liu, Y.; Fu, Q.; Cao, Z.; Ma, X.; Li, X. Characterization of tumor-associated endothelial cells and the development of a prognostic model in pancreatic ductal adenocarcinoma. Biochim. Et Biophys. Acta General. Subj. 2024, 1868, 130545. [Google Scholar] [CrossRef]
  70. Buchler, P.; Reber, H.A.; Buchler, M.; Shrinkante, S.; Buchler, M.W.; Friess, H.; Semenza, G.L.; Hines, O.J. Hypoxia-inducible factor 1 regulates vascular endothelial growth factor expression in human pancreatic cancer. Pancreas 2003, 26, 56–64. [Google Scholar] [CrossRef]
  71. Itakura, J.; Ishiwata, T.; Friess, H.; Fujii, H.; Matsumoto, Y.; Buchler, M.W.; Korc, M. Enhanced expression of vascular endothelial growth factor in human pancreatic cancer correlates with local disease progression. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 1997, 3, 1309–1316. [Google Scholar]
  72. Wei, D.; Le, X.; Zheng, L.; Wang, L.; Frey, J.A.; Gao, A.C.; Peng, Z.; Huang, S.; Xiong, H.Q.; Abbruzzese, J.L.; et al. Stat3 activation regulates the expression of vascular endothelial growth factor and human pancreatic cancer angiogenesis and metastasis. Oncogene 2003, 22, 319–329. [Google Scholar] [CrossRef] [PubMed]
  73. Clark, C.E.; Hingorani, S.R.; Mick, R.; Combs, C.; Tuveson, D.A.; Vonderheide, R.H. Dynamics of the immune reaction to pancreatic cancer from inception to invasion. Cancer Res. 2007, 67, 9518–9527. [Google Scholar] [CrossRef] [PubMed]
  74. Esposito, I.; Menicagli, M.; Funel, N.; Bergmann, F.; Boggi, U.; Mosca, F.; Bevilacqua, G.; Campani, D. Inflammatory cells contribute to the generation of an angiogenic phenotype in pancreatic ductal adenocarcinoma. J. Clin. Pathol. 2004, 57, 630–636. [Google Scholar] [CrossRef] [PubMed]
  75. Santoni, M.; Bracarda, S.; Nabissi, M.; Massari, F.; Conti, A.; Bria, E.; Tortora, G.; Santoni, G.; Cascinu, S. CXC and CC chemokines as angiogenic modulators in nonhaematological tumors. Biomed. Res. Int. 2014, 2014, 768758. [Google Scholar] [CrossRef]
  76. Gordon, S.; Taylor, P.R. Monocyte and macrophage heterogeneity. Nat. Rev. Immunol. 2005, 5, 953–964. [Google Scholar] [CrossRef]
  77. Conway, E.M.; Pikor, L.A.; Kung, S.H.; Hamilton, M.J.; Lam, S.; Lam, W.L.; Bennewith, K.L. Macrophages, Inflammation, and Lung Cancer. Am. J. Respir. Crit. Care Med. 2016, 193, 116–130. [Google Scholar] [CrossRef]
  78. Pittet, M.J.; Michielin, O.; Migliorini, D. Clinical relevance of tumour-associated macrophages. Nat. Rev. Clin. Oncol. 2022, 19, 402–421. [Google Scholar] [CrossRef]
  79. Penny, H.L.; Sieow, J.L.; Adriani, G.; Yeap, W.H.; See Chi Ee, P.; San Luis, B.; Lee, B.; Lee, T.; Mak, S.Y.; Ho, Y.S.; et al. Warburg metabolism in tumor-conditioned macrophages promotes metastasis in human pancreatic ductal adenocarcinoma. Oncoimmunology 2016, 5, e1191731. [Google Scholar] [CrossRef]
  80. Ino, Y.; Yamazaki-Itoh, R.; Shimada, K.; Iwasaki, M.; Kosuge, T.; Kanai, Y.; Hiraoka, N. Immune cell infiltration as an indicator of the immune microenvironment of pancreatic cancer. Br. J. Cancer 2013, 108, 914–923. [Google Scholar] [CrossRef]
  81. Sanford, D.E.; Belt, B.A.; Panni, R.Z.; Mayer, A.; Deshpande, A.D.; Carpenter, D.; Mitchem, J.B.; Plambeck-Suess, S.M.; Worley, L.A.; Goetz, B.D.; et al. Inflammatory monocyte mobilization decreases patient survival in pancreatic cancer: A role for targeting the CCL2/CCR2 axis. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2013, 19, 3404–3415. [Google Scholar] [CrossRef]
  82. Yin, Z.; Ma, T.; Huang, B.; Lin, L.; Zhou, Y.; Yan, J.; Zou, Y.; Chen, S. Macrophage-derived exosomal microRNA-501-3p promotes progression of pancreatic ductal adenocarcinoma through the TGFBR3-mediated TGF-beta signaling pathway. J. Exp. Clin. Cancer Res. 2019, 38, 310. [Google Scholar] [CrossRef] [PubMed]
  83. Chen, Y.; Lei, Y.; Li, J.; Wang, X.; Li, G. Macrophage-derived exosomal microRNAs promote metastasis in pancreatic ductal adenocarcinoma. Int. Immunopharmacol. 2024, 129, 111590. [Google Scholar] [CrossRef] [PubMed]
  84. Weizman, N.; Krelin, Y.; Shabtay-Orbach, A.; Amit, M.; Binenbaum, Y.; Wong, R.J.; Gil, Z. Macrophages mediate gemcitabine resistance of pancreatic adenocarcinoma by upregulating cytidine deaminase. Oncogene 2014, 33, 3812–3819. [Google Scholar] [CrossRef]
  85. Stromnes, I.M.; Brockenbrough, J.S.; Izeradjene, K.; Carlson, M.A.; Cuevas, C.; Simmons, R.M.; Greenberg, P.D.; Hingorani, S.R. Targeted depletion of an MDSC subset unmasks pancreatic ductal adenocarcinoma to adaptive immunity. Gut 2014, 63, 1769–1781. [Google Scholar] [CrossRef] [PubMed]
  86. Bayne, L.J.; Beatty, G.L.; Jhala, N.; Clark, C.E.; Rhim, A.D.; Stanger, B.Z.; Vonderheide, R.H. Tumor-derived granulocyte-macrophage colony-stimulating factor regulates myeloid inflammation and T cell immunity in pancreatic cancer. Cancer Cell 2012, 21, 822–835. [Google Scholar] [CrossRef]
  87. Pylayeva-Gupta, Y.; Das, S.; Handler, J.S.; Hajdu, C.H.; Coffre, M.; Koralov, S.B.; Bar-Sagi, D. IL35-Producing B Cells Promote the Development of Pancreatic Neoplasia. Cancer Discov. 2016, 6, 247–255. [Google Scholar] [CrossRef]
  88. Mace, T.A.; Bloomston, M.; Lesinski, G.B. Pancreatic cancer-associated stellate cells: A viable target for reducing immunosuppression in the tumor microenvironment. Oncoimmunology 2013, 2, e24891. [Google Scholar] [CrossRef]
  89. Yang, Y.; Li, C.; Liu, T.; Dai, X.; Bazhin, A.V. Myeloid-Derived Suppressor Cells in Tumors: From Mechanisms to Antigen Specificity and Microenvironmental Regulation. Front. Immunol. 2020, 11, 1371. [Google Scholar] [CrossRef]
  90. Velasco, R.M.; Garcia, A.G.; Sanchez, P.J.; Sellart, I.M.; Sanchez-Arevalo Lobo, V.J. Tumour microenvironment and heterotypic interactions in pancreatic cancer. J. Physiol. Biochem. 2022, 79, 179–192. [Google Scholar] [CrossRef]
  91. Diaz-Montero, C.M.; Salem, M.L.; Nishimura, M.I.; Garrett-Mayer, E.; Cole, D.J.; Montero, A.J. Increased circulating myeloid-derived suppressor cells correlate with clinical cancer stage, metastatic tumor burden, and doxorubicin-cyclophosphamide chemotherapy. Cancer Immunol. Immunother. 2009, 58, 49–59. [Google Scholar] [CrossRef]
  92. Hanson, E.M.; Clements, V.K.; Sinha, P.; Ilkovitch, D.; Ostrand-Rosenberg, S. Myeloid-derived suppressor cells down-regulate L-selectin expression on CD4+ and CD8+ T cells. J. Immunol. 2009, 183, 937–944. [Google Scholar] [CrossRef]
  93. Lanier, L.L.; Testi, R.; Bindl, J.; Phillips, J.H. Identity of Leu-19 (CD56) leukocyte differentiation antigen and neural cell adhesion molecule. J. Exp. Med. 1989, 169, 2233–2238. [Google Scholar] [CrossRef]
  94. Chiossone, L.; Dumas, P.Y.; Vienne, M.; Vivier, E. Natural killer cells and other innate lymphoid cells in cancer. Nat. Rev. Immunol. 2018, 18, 671–688. [Google Scholar] [CrossRef] [PubMed]
  95. Paul, S.; Lal, G. The Molecular Mechanism of Natural Killer Cells Function and Its Importance in Cancer Immunotherapy. Front. Immunol. 2017, 8, 1124. [Google Scholar] [CrossRef] [PubMed]
  96. Davis, M.; Conlon, K.; Bohac, G.C.; Barcenas, J.; Leslie, W.; Watkins, L.; Lamzabi, I.; Deng, Y.; Li, Y.; Plate, J.M. Effect of pemetrexed on innate immune killer cells and adaptive immune T cells in subjects with adenocarcinoma of the pancreas. J. Immunother. 2012, 35, 629–640. [Google Scholar] [CrossRef] [PubMed]
  97. Liu, S.; Dhar, P.; Wu, J.D. NK Cell Plasticity in Cancer. J. Clin. Med. 2019, 8, 1492. [Google Scholar] [CrossRef] [PubMed]
  98. Funa, K.; Nilsson, B.; Jacobsson, G.; Alm, G.V. Decreased natural killer cell activity and interferon production by leucocytes in patients with adenocarcinoma of the pancreas. Br. J. Cancer 1984, 50, 231–233. [Google Scholar] [CrossRef]
  99. Ohta, T.; Elnemr, A.; Kitagawa, H.; Kayahara, M.; Takamura, H.; Fujimura, T.; Nishimura, G.; Shimizu, K.; Yi, S.Q.; Miwa, K. Fas ligand expression in human pancreatic cancer. Oncol. Rep. 2004, 12, 749–754. [Google Scholar] [CrossRef]
  100. Peng, Y.P.; Zhang, J.J.; Liang, W.B.; Tu, M.; Lu, Z.P.; Wei, J.S.; Jiang, K.R.; Gao, W.T.; Wu, J.L.; Xu, Z.K.; et al. Elevation of MMP-9 and IDO induced by pancreatic cancer cells mediates natural killer cell dysfunction. BMC Cancer 2014, 14, 738. [Google Scholar] [CrossRef]
  101. Peng, Y.P.; Zhu, Y.; Zhang, J.J.; Xu, Z.K.; Qian, Z.Y.; Dai, C.C.; Jiang, K.R.; Wu, J.L.; Gao, W.T.; Li, Q.; et al. Comprehensive analysis of the percentage of surface receptors and cytotoxic granules positive natural killer cells in patients with pancreatic cancer, gastric cancer, and colorectal cancer. J. Transl. Med. 2013, 11, 262. [Google Scholar] [CrossRef]
  102. Nywening, T.M.; Belt, B.A.; Cullinan, D.R.; Panni, R.Z.; Han, B.J.; Sanford, D.E.; Jacobs, R.C.; Ye, J.; Patel, A.A.; Gillanders, W.E.; et al. Targeting both tumour-associated CXCR2(+) neutrophils and CCR2(+) macrophages disrupts myeloid recruitment and improves chemotherapeutic responses in pancreatic ductal adenocarcinoma. Gut 2018, 67, 1112–1123. [Google Scholar] [CrossRef] [PubMed]
  103. Mollinedo, F. Neutrophil Degranulation, Plasticity, and Cancer Metastasis. Trends Immunol. 2019, 40, 228–242. [Google Scholar] [CrossRef] [PubMed]
  104. Qin, F.; Liu, X.; Chen, J.; Huang, S.; Wei, W.; Zou, Y.; Liu, X.; Deng, K.; Mo, S.; Chen, J.; et al. Anti-TGF-beta attenuates tumor growth via polarization of tumor associated neutrophils towards an anti-tumor phenotype in colorectal cancer. J. Cancer 2020, 11, 2580–2592. [Google Scholar] [CrossRef]
  105. Felix, K.; Gaida, M.M. Neutrophil-Derived Proteases in the Microenvironment of Pancreatic Cancer -Active Players in Tumor Progression. Int. J. Biol. Sci. 2016, 12, 302–313. [Google Scholar] [CrossRef] [PubMed]
  106. Nielsen, S.R.; Strobech, J.E.; Horton, E.R.; Jackstadt, R.; Laitala, A.; Bravo, M.C.; Maltese, G.; Jensen, A.R.D.; Reuten, R.; Rafaeva, M.; et al. Suppression of tumor-associated neutrophils by lorlatinib attenuates pancreatic cancer growth and improves treatment with immune checkpoint blockade. Nat. Commun. 2021, 12, 3414. [Google Scholar] [CrossRef] [PubMed]
  107. Guo, J.; Wu, M.; Guo, L.; Zuo, Q. Pretreatment blood neutrophil/lymphocyte ratio is associated with metastasis and predicts survival in patients with pancreatic cancer. Bull. Cancer 2018, 105, 146–154. [Google Scholar] [CrossRef]
  108. Takesue, S.; Ohuchida, K.; Shinkawa, T.; Otsubo, Y.; Matsumoto, S.; Sagara, A.; Yonenaga, A.; Ando, Y.; Kibe, S.; Nakayama, H.; et al. Neutrophil extracellular traps promote liver micrometastasis in pancreatic ductal adenocarcinoma via the activation of cancerassociated fibroblasts. Int. J. Oncol. 2020, 56, 596–605. [Google Scholar]
  109. Villadangos, J.A.; Schnorrer, P. Intrinsic and cooperative antigen-presenting functions of dendritic-cell subsets in vivo. Nat. Rev. Immunol. 2007, 7, 543–555. [Google Scholar] [CrossRef]
  110. Lucarini, V.; Melaiu, O.; Tempora, P.; D’Amico, S.; Locatelli, F.; Fruci, D. Dendritic Cells: Behind the Scenes of T-Cell Infiltration into the Tumor Microenvironment. Cancers 2021, 13, 433. [Google Scholar] [CrossRef]
  111. Jang, J.E.; Hajdu, C.H.; Liot, C.; Miller, G.; Dustin, M.L.; Bar-Sagi, D. Crosstalk between Regulatory T Cells and Tumor-Associated Dendritic Cells Negates Anti-tumor Immunity in Pancreatic Cancer. Cell Rep. 2017, 20, 558–571. [Google Scholar] [CrossRef]
  112. Bellone, G.; Carbone, A.; Smirne, C.; Scirelli, T.; Buffolino, A.; Novarino, A.; Stacchini, A.; Bertetto, O.; Palestro, G.; Sorio, C.; et al. Cooperative induction of a tolerogenic dendritic cell phenotype by cytokines secreted by pancreatic carcinoma cells. J. Immunol. 2006, 177, 3448–3460. [Google Scholar] [CrossRef]
  113. Bharadwaj, U.; Li, M.; Zhang, R.; Chen, C.; Yao, Q. Elevated interleukin-6 and G-CSF in human pancreatic cancer cell conditioned medium suppress dendritic cell differentiation and activation. Cancer Res. 2007, 67, 5479–5488. [Google Scholar] [CrossRef] [PubMed]
  114. Wang, T.; Niu, G.; Kortylewski, M.; Burdelya, L.; Shain, K.; Zhang, S.; Bhattacharya, R.; Gabrilovich, D.; Heller, R.; Coppola, D.; et al. Regulation of the innate and adaptive immune responses by Stat-3 signaling in tumor cells. Nat. Med. 2004, 10, 48–54. [Google Scholar] [CrossRef] [PubMed]
  115. Kobayashi, M.; Shimodaira, S.; Nagai, K.; Ogasawara, M.; Takahashi, H.; Abe, H.; Tanii, M.; Okamoto, M.; Tsujitani, S.; Yusa, S.; et al. Prognostic factors related to add-on dendritic cell vaccines on patients with inoperable pancreatic cancer receiving chemotherapy: A multicenter analysis. Cancer Immunol. Immunother. 2014, 63, 797–806. [Google Scholar] [CrossRef]
  116. Goulart, M.R.; Stasinos, K.; Fincham, R.E.A.; Delvecchio, F.R.; Kocher, H.M. T cells in pancreatic cancer stroma. World J. Gastroenterol. 2021, 27, 7956–7968. [Google Scholar] [CrossRef]
  117. Ajina, R.; Weiner, L.M. T-Cell Immunity in Pancreatic Cancer. Pancreas 2020, 49, 1014–1023. [Google Scholar] [CrossRef]
  118. Gorchs, L.; Fernandez Moro, C.; Bankhead, P.; Kern, K.P.; Sadeak, I.; Meng, Q.; Rangelova, E.; Kaipe, H. Human Pancreatic Carcinoma-Associated Fibroblasts Promote Expression of Co-inhibitory Markers on CD4(+) and CD8(+) T-Cells. Front. Immunol. 2019, 10, 847. [Google Scholar] [CrossRef] [PubMed]
  119. Chang, J.H.; Jiang, Y.; Pillarisetty, V.G. Role of immune cells in pancreatic cancer from bench to clinical application: An updated review. Medicine 2016, 95, e5541. [Google Scholar] [CrossRef] [PubMed]
  120. Suzuki, D.; Furukawa, K.; Kimura, F.; Shimizu, H.; Yoshidome, H.; Ohtsuka, M.; Kato, A.; Yoshitomi, H.; Miyazaki, M. Effects of perioperative immunonutrition on cell-mediated immunity, T helper type 1 (Th1)/Th2 differentiation, and Th17 response after pancreaticoduodenectomy. Surgery 2010, 148, 573–581. [Google Scholar] [CrossRef]
  121. Garcia, A.R.; Mendes, A.; Custodia, C.; Faria, C.C.; Barata, J.T.; Malho, R.; Figueira, I.; Brito, M.A. Abrogating Metastatic Properties of Triple-Negative Breast Cancer Cells by EGFR and PI3K Dual Inhibitors. Cancers 2023, 15, 3973. [Google Scholar] [CrossRef]
  122. Zhang, Y.; Lazarus, J.; Steele, N.G.; Yan, W.; Lee, H.J.; Nwosu, Z.C.; Halbrook, C.J.; Menjivar, R.E.; Kemp, S.B.; Sirihorachai, V.R.; et al. Regulatory T-cell Depletion Alters the Tumor Microenvironment and Accelerates Pancreatic Carcinogenesis. Cancer Discov. 2020, 10, 422–439. [Google Scholar] [CrossRef]
  123. Gao, Z.Z.Q.; Zhang, X.; Song, Y. Advance of T regulatory cells in tumor microenvironment remodeling and immunotherapy in pancreatic cancer. Eur. J. Inflamm. 2022, 20, 1721727X221092900. [Google Scholar] [CrossRef]
  124. Engelhard, V.; Conejo-Garcia, J.R.; Ahmed, R.; Nelson, B.H.; Willard-Gallo, K.; Bruno, T.C.; Fridman, W.H. B cells and cancer. Cancer Cell 2021, 39, 1293–1296. [Google Scholar] [CrossRef] [PubMed]
  125. Delvecchio, F.R.; Goulart, M.R.; Fincham, R.E.A.; Bombadieri, M.; Kocher, H.M. B cells in pancreatic cancer stroma. World J. Gastroenterol. 2022, 28, 1088–1101. [Google Scholar] [CrossRef] [PubMed]
  126. Castino, G.F.; Cortese, N.; Capretti, G.; Serio, S.; Di Caro, G.; Mineri, R.; Magrini, E.; Grizzi, F.; Cappello, P.; Novelli, F.; et al. Spatial distribution of B cells predicts prognosis in human pancreatic adenocarcinoma. Oncoimmunology 2016, 5, e1085147. [Google Scholar] [CrossRef] [PubMed]
  127. Shalapour, S.; Font-Burgada, J.; Di Caro, G.; Zhong, Z.; Sanchez-Lopez, E.; Dhar, D.; Willimsky, G.; Ammirante, M.; Strasner, A.; Hansel, D.E.; et al. Immunosuppressive plasma cells impede T-cell-dependent immunogenic chemotherapy. Nature 2015, 521, 94–98. [Google Scholar] [CrossRef]
  128. Catalan, D.; Mansilla, M.A.; Ferrier, A.; Soto, L.; Oleinika, K.; Aguillon, J.C.; Aravena, O. Immunosuppressive Mechanisms of Regulatory B Cells. Front. Immunol. 2021, 12, 611795. [Google Scholar] [CrossRef]
  129. Sindhava, V.J.; Bondada, S. Multiple regulatory mechanisms control B-1 B cell activation. Front. Immunol. 2012, 3, 372. [Google Scholar] [CrossRef]
  130. Lee, K.E.; Spata, M.; Bayne, L.J.; Buza, E.L.; Durham, A.C.; Allman, D.; Vonderheide, R.H.; Simon, M.C. Hif1a Deletion Reveals Pro-Neoplastic Function of B Cells in Pancreatic Neoplasia. Cancer Discov. 2016, 6, 256–269. [Google Scholar] [CrossRef]
  131. Gunderson, A.J.; Kaneda, M.M.; Tsujikawa, T.; Nguyen, A.V.; Affara, N.I.; Ruffell, B.; Gorjestani, S.; Liudahl, S.M.; Truitt, M.; Olson, P.; et al. Bruton Tyrosine Kinase-Dependent Immune Cell Cross-talk Drives Pancreas Cancer. Cancer Discov. 2016, 6, 270–285. [Google Scholar] [CrossRef]
  132. Huber, M.; Brehm, C.U.; Gress, T.M.; Buchholz, M.; Alashkar Alhamwe, B.; von Strandmann, E.P.; Slater, E.P.; Bartsch, J.W.; Bauer, C.; Lauth, M. The Immune Microenvironment in Pancreatic Cancer. Int. J. Mol. Sci. 2020, 21, 7307. [Google Scholar] [CrossRef]
  133. Dougan, S.K. The Pancreatic Cancer Microenvironment. Cancer J. 2017, 23, 321–325. [Google Scholar] [CrossRef] [PubMed]
  134. Pihlak, R.; Weaver, J.M.J.; Valle, J.W.; McNamara, M.G. Advances in Molecular Profiling and Categorisation of Pancreatic Adenocarcinoma and the Implications for Therapy. Cancers 2018, 10, 17. [Google Scholar] [CrossRef] [PubMed]
  135. Riquelme, E.; Maitra, A.; McAllister, F. Immunotherapy for Pancreatic Cancer: More Than Just a Gut Feeling. Cancer Discov. 2018, 8, 386–388. [Google Scholar] [CrossRef]
  136. Murakami, T.; Hiroshima, Y.; Matsuyama, R.; Homma, Y.; Hoffman, R.M.; Endo, I. Role of the tumor microenvironment in pancreatic cancer. Ann. Gastroenterol. Surg. 2019, 3, 130–137. [Google Scholar] [CrossRef] [PubMed]
  137. Herting, C.J.; Karpovsky, I.; Lesinski, G.B. The tumor microenvironment in pancreatic ductal adenocarcinoma: Current perspectives and future directions. Cancer Metastasis Rev. 2021, 40, 675–689. [Google Scholar] [CrossRef] [PubMed]
  138. Dong, H.; Strome, S.E.; Salomao, D.R.; Tamura, H.; Hirano, F.; Flies, D.B.; Roche, P.C.; Lu, J.; Zhu, G.; Tamada, K.; et al. Tumor-associated B7-H1 promotes T-cell apoptosis: A potential mechanism of immune evasion. Nat. Med. 2002, 8, 793–800. [Google Scholar] [CrossRef]
  139. Li, X.; Gulati, M.; Larson, A.C.; Solheim, J.C.; Jain, M.; Kumar, S.; Batra, S.K. Immune checkpoint blockade in pancreatic cancer: Trudging through the immune desert. Semin. Cancer Biol. 2022, 86 Pt 2, 14–27. [Google Scholar] [CrossRef]
  140. Peggs, K.S.; Quezada, S.A.; Chambers, C.A.; Korman, A.J.; Allison, J.P. Blockade of CTLA-4 on both effector and regulatory T cell compartments contributes to the antitumor activity of anti-CTLA-4 antibodies. J. Exp. Med. 2009, 206, 1717–1725. [Google Scholar] [CrossRef]
  141. Sunshine, J.; Taube, J.M. PD-1/PD-L1 inhibitors. Curr. Opin. Pharmacol. 2015, 23, 32–38. [Google Scholar] [CrossRef]
  142. Leach, D.R.; Krummel, M.F.; Allison, J.P. Enhancement of antitumor immunity by CTLA-4 blockade. Science 1996, 271, 1734–1736. [Google Scholar] [CrossRef]
  143. Royal, R.E.; Levy, C.; Turner, K.; Mathur, A.; Hughes, M.; Kammula, U.S.; Sherry, R.M.; Topalian, S.L.; Yang, J.C.; Lowy, I.; et al. Phase 2 trial of single agent Ipilimumab (anti-CTLA-4) for locally advanced or metastatic pancreatic adenocarcinoma. J. Immunother. 2010, 33, 828–833. [Google Scholar] [CrossRef] [PubMed]
  144. Brahmer, J.R.; Tykodi, S.S.; Chow, L.Q.; Hwu, W.J.; Topalian, S.L.; Hwu, P.; Drake, C.G.; Camacho, L.H.; Kauh, J.; Odunsi, K.; et al. Safety and activity of anti-PD-L1 antibody in patients with advanced cancer. N. Engl. J. Med. 2012, 366, 2455–2465. [Google Scholar] [CrossRef]
  145. Stromnes, I.M.; Hulbert, A.; Pierce, R.H.; Greenberg, P.D.; Hingorani, S.R. T-cell Localization, Activation, and Clonal Expansion in Human Pancreatic Ductal Adenocarcinoma. Cancer Immunol. Res. 2017, 5, 978–991. [Google Scholar] [CrossRef]
  146. Ho, W.J.; Jaffee, E.M.; Zheng, L. The tumour microenvironment in pancreatic cancer—Clinical challenges and opportunities. Nat. Rev. Clin. Oncol. 2020, 17, 527–540. [Google Scholar] [CrossRef]
  147. Le, D.T.; Durham, J.N.; Smith, K.N.; Wang, H.; Bartlett, B.R.; Aulakh, L.K.; Lu, S.; Kemberling, H.; Wilt, C.; Luber, B.S.; et al. Mismatch repair deficiency predicts response of solid tumors to PD-1 blockade. Science 2017, 357, 409–413. [Google Scholar] [CrossRef]
  148. Le, D.T.; Uram, J.N.; Wang, H.; Bartlett, B.R.; Kemberling, H.; Eyring, A.D.; Skora, A.D.; Luber, B.S.; Azad, N.S.; Laheru, D.; et al. PD-1 Blockade in Tumors with Mismatch-Repair Deficiency. N. Engl. J. Med. 2015, 372, 2509–2520. [Google Scholar] [CrossRef] [PubMed]
  149. Humphris, J.L.; Patch, A.M.; Nones, K.; Bailey, P.J.; Johns, A.L.; McKay, S.; Chang, D.K.; Miller, D.K.; Pajic, M.; Kassahn, K.S.; et al. Hypermutation In Pancreatic Cancer. Gastroenterology 2017, 152, 68–74.e2. [Google Scholar] [CrossRef]
  150. Bian, J.; Almhanna, K. Pancreatic cancer and immune checkpoint inhibitors-still a long way to go. Transl. Gastroenterol. Hepatol. 2021, 6, 6. [Google Scholar] [CrossRef] [PubMed]
  151. Marabelle, A.; Le, D.T.; Ascierto, P.A.; Di Giacomo, A.M.; De Jesus-Acosta, A.; Delord, J.P.; Geva, R.; Gottfried, M.; Penel, N.; Hansen, A.R.; et al. Efficacy of Pembrolizumab in Patients with Noncolorectal High Microsatellite Instability/Mismatch Repair-Deficient Cancer: Results From the Phase II KEYNOTE-158 Study. J. Clin. Oncol. 2020, 38, 1–10. [Google Scholar] [CrossRef] [PubMed]
  152. Weiss, G.J.; Blaydorn, L.; Beck, J.; Bornemann-Kolatzki, K.; Urnovitz, H.; Schutz, E.; Khemka, V. Phase Ib/II study of gemcitabine, nab-paclitaxel, and pembrolizumab in metastatic pancreatic adenocarcinoma. Investig. New Drugs 2018, 36, 96–102. [Google Scholar] [CrossRef]
  153. Ahmed, A.; Tait, S.W.G. Targeting immunogenic cell death in cancer. Mol. Oncol. 2020, 14, 2994–3006. [Google Scholar] [CrossRef] [PubMed]
  154. Kamath, S.D.; Kalyan, A.; Kircher, S.; Nimeiri, H.; Fought, A.J.; Benson, A., 3rd; Mulcahy, M. Ipilimumab and Gemcitabine for Advanced Pancreatic Cancer: A Phase Ib Study. Oncologist 2020, 25, e808–e815. [Google Scholar] [CrossRef]
  155. O’Reilly, E.M.; Oh, D.Y.; Dhani, N.; Renouf, D.J.; Lee, M.A.; Sun, W.; Fisher, G.; Hezel, A.; Chang, S.C.; Vlahovic, G.; et al. Durvalumab with or without Tremelimumab for Patients with Metastatic Pancreatic Ductal Adenocarcinoma: A Phase 2 Randomized Clinical Trial. JAMA Oncol. 2019, 5, 1431–1438. [Google Scholar] [CrossRef] [PubMed]
  156. Kumar, V.; Boucher, Y.; Liu, H.; Ferreira, D.; Hooker, J.; Catana, C.; Hoover, A.J.; Ritter, T.; Jain, R.K.; Guimaraes, A.R. Noninvasive Assessment of Losartan-Induced Increase in Functional Microvasculature and Drug Delivery in Pancreatic Ductal Adenocarcinoma. Transl. Oncol. 2016, 9, 431–437. [Google Scholar] [CrossRef] [PubMed]
  157. Haen, S.P.; Loffler, M.W.; Rammensee, H.G.; Brossart, P. Towards new horizons: Characterization, classification and implications of the tumour antigenic repertoire. Nat. Rev. Clin. Oncol. 2020, 17, 595–610. [Google Scholar] [CrossRef] [PubMed]
  158. Yarchoan, M.; Albacker, L.A.; Hopkins, A.C.; Montesion, M.; Murugesan, K.; Vithayathil, T.T.; Zaidi, N.; Azad, N.S.; Laheru, D.A.; Frampton, G.M.; et al. PD-L1 expression and tumor mutational burden are independent biomarkers in most cancers. JCI Insight 2019, 4, e126908. [Google Scholar] [CrossRef]
  159. Mayanagi, S.; Kitago, M.; Sakurai, T.; Matsuda, T.; Fujita, T.; Higuchi, H.; Taguchi, J.; Takeuchi, H.; Itano, O.; Aiura, K.; et al. Phase I pilot study of Wilms tumor gene 1 peptide-pulsed dendritic cell vaccination combined with gemcitabine in pancreatic cancer. Cancer Sci. 2015, 106, 397–406. [Google Scholar] [CrossRef]
  160. Le, D.T.; Wang-Gillam, A.; Picozzi, V.; Greten, T.F.; Crocenzi, T.; Springett, G.; Morse, M.; Zeh, H.; Cohen, D.; Fine, R.L.; et al. Safety and survival with GVAX pancreas prime and Listeria Monocytogenes-expressing mesothelin (CRS-207) boost vaccines for metastatic pancreatic cancer. J. Clin. Oncol. 2015, 33, 1325–1333. [Google Scholar] [CrossRef]
  161. Bernhardt, S.L.; Gjertsen, M.K.; Trachsel, S.; Moller, M.; Eriksen, J.A.; Meo, M.; Buanes, T.; Gaudernack, G. Telomerase peptide vaccination of patients with non-resectable pancreatic cancer: A dose escalating phase I/II study. Br. J. Cancer 2006, 95, 1474–1482. [Google Scholar] [CrossRef]
  162. Amedei, A.; Niccolai, E.; Prisco, D. Pancreatic cancer: Role of the immune system in cancer progression and vaccine-based immunotherapy. Hum. Vaccines Immunother. 2014, 10, 3354–3368. [Google Scholar] [CrossRef]
  163. Luo, W.; Yang, G.; Luo, W.; Cao, Z.; Liu, Y.; Qiu, J.; Chen, G.; You, L.; Zhao, F.; Zheng, L.; et al. Novel therapeutic strategies and perspectives for metastatic pancreatic cancer: Vaccine therapy is more than just a theory. Cancer Cell Int. 2020, 20, 66. [Google Scholar] [CrossRef]
  164. Zhang, Y.; Velez-Delgado, A.; Mathew, E.; Li, D.; Mendez, F.M.; Flannagan, K.; Rhim, A.D.; Simeone, D.M.; Beatty, G.L.; Pasca di Magliano, M. Myeloid cells are required for PD-1/PD-L1 checkpoint activation and the establishment of an immunosuppressive environment in pancreatic cancer. Gut 2017, 66, 124–136. [Google Scholar] [CrossRef] [PubMed]
  165. Beatty, G.L.; Chiorean, E.G.; Fishman, M.P.; Saboury, B.; Teitelbaum, U.R.; Sun, W.; Huhn, R.D.; Song, W.; Li, D.; Sharp, L.L.; et al. CD40 agonists alter tumor stroma and show efficacy against pancreatic carcinoma in mice and humans. Science 2011, 331, 1612–1616. [Google Scholar] [CrossRef] [PubMed]
  166. Mitchem, J.B.; Brennan, D.J.; Knolhoff, B.L.; Belt, B.A.; Zhu, Y.; Sanford, D.E.; Belaygorod, L.; Carpenter, D.; Collins, L.; Piwnica-Worms, D.; et al. Targeting tumor-infiltrating macrophages decreases tumor-initiating cells, relieves immunosuppression, and improves chemotherapeutic responses. Cancer Res. 2013, 73, 1128–1141. [Google Scholar] [CrossRef]
  167. Nywening, T.M.; Wang-Gillam, A.; Sanford, D.E.; Belt, B.A.; Panni, R.Z.; Cusworth, B.M.; Toriola, A.T.; Nieman, R.K.; Worley, L.A.; Yano, M.; et al. Targeting tumour-associated macrophages with CCR2 inhibition in combination with FOLFIRINOX in patients with borderline resectable and locally advanced pancreatic cancer: A single-centre, open-label, dose-finding, non-randomised, phase 1b trial. Lancet Oncol. 2016, 17, 651–662. [Google Scholar] [CrossRef]
  168. Zhu, Y.; Knolhoff, B.L.; Meyer, M.A.; Nywening, T.M.; West, B.L.; Luo, J.; Wang-Gillam, A.; Goedegebuure, S.P.; Linehan, D.C.; DeNardo, D.G. CSF1/CSF1R blockade reprograms tumor-infiltrating macrophages and improves response to T-cell checkpoint immunotherapy in pancreatic cancer models. Cancer Res. 2014, 74, 5057–5069. [Google Scholar] [CrossRef]
  169. Kumar, V.; Donthireddy, L.; Marvel, D.; Condamine, T.; Wang, F.; Lavilla-Alonso, S.; Hashimoto, A.; Vonteddu, P.; Behera, R.; Goins, M.A.; et al. Cancer-Associated Fibroblasts Neutralize the Anti-tumor Effect of CSF1 Receptor Blockade by Inducing PMN-MDSC Infiltration of Tumors. Cancer Cell 2017, 32, 654–668.e5. [Google Scholar] [CrossRef]
  170. Steele, C.W.; Karim, S.A.; Leach, J.D.G.; Bailey, P.; Upstill-Goddard, R.; Rishi, L.; Foth, M.; Bryson, S.; McDaid, K.; Wilson, Z.; et al. CXCR2 Inhibition Profoundly Suppresses Metastases and Augments Immunotherapy in Pancreatic Ductal Adenocarcinoma. Cancer Cell 2016, 29, 832–845. [Google Scholar] [CrossRef] [PubMed]
  171. Principe, D.R.; Xiong, R.; Li, Y.; Pham, T.N.D.; Kamath, S.D.; Dubrovskyi, O.; Ratia, K.; Huang, F.; Zhao, J.; Shen, Z.; et al. XP-524 is a dual-BET/EP300 inhibitor that represses oncogenic KRAS and potentiates immune checkpoint inhibition in pancreatic cancer. Proc. Natl. Acad. Sci. USA 2022, 119, e2116764119. [Google Scholar] [CrossRef]
  172. Gulhati, P.; Schalck, A.; Jiang, S.; Shang, X.; Wu, C.J.; Hou, P.; Ruiz, S.H.; Soto, L.S.; Parra, E.; Ying, H.; et al. Targeting T cell checkpoints 41BB and LAG3 and myeloid cell CXCR1/CXCR2 results in antitumor immunity and durable response in pancreatic cancer. Nat. Cancer 2023, 4, 62–80. [Google Scholar] [CrossRef] [PubMed]
  173. Seifert, L.; Plesca, I.; Muller, L.; Sommer, U.; Heiduk, M.; von Renesse, J.; Digomann, D.; Gluck, J.; Klimova, A.; Weitz, J.; et al. LAG-3-Expressing Tumor-Infiltrating T Cells Are Associated with Reduced Disease-Free Survival in Pancreatic Cancer. Cancers 2021, 13, 1297. [Google Scholar] [CrossRef] [PubMed]
  174. Meng, Q.; Liu, Z.; Rangelova, E.; Poiret, T.; Ambati, A.; Rane, L.; Xie, S.; Verbeke, C.; Dodoo, E.; Del Chiaro, M.; et al. Expansion of Tumor-reactive T Cells From Patients with Pancreatic Cancer. J. Immunother. 2016, 39, 81–89. [Google Scholar] [CrossRef]
  175. Alrifai, D.; Sarker, D.; Maher, J. Prospects for adoptive immunotherapy of pancreatic cancer using chimeric antigen receptor-engineered T-cells. Immunopharmacol. Immunotoxicol. 2016, 38, 50–60. [Google Scholar] [CrossRef] [PubMed]
  176. Yeo, D.; Giardina, C.; Saxena, P.; Rasko, J.E.J. The next wave of cellular immunotherapies in pancreatic cancer. Mol. Ther. Oncolytics 2022, 24, 561–576. [Google Scholar] [CrossRef]
  177. Katz, S.C.; Moody, A.E.; Guha, P.; Hardaway, J.C.; Prince, E.; LaPorte, J.; Stancu, M.; Slansky, J.E.; Jordan, K.R.; Schulick, R.D.; et al. HITM-SURE: Hepatic immunotherapy for metastases phase Ib anti-CEA CAR-T study utilizing pressure enabled drug delivery. J. Immunother. Cancer 2020, 8, e001097. [Google Scholar] [CrossRef]
  178. Watanabe, K.; Luo, Y.; Da, T.; Guedan, S.; Ruella, M.; Scholler, J.; Keith, B.; Young, R.M.; Engels, B.; Sorsa, S.; et al. Pancreatic cancer therapy with combined mesothelin-redirected chimeric antigen receptor T cells and cytokine-armed oncolytic adenoviruses. JCI Insight 2018, 3, e99573. [Google Scholar] [CrossRef]
  179. Penack, O.; Koenecke, C. Complications after CD19+ CAR T-Cell Therapy. Cancers 2020, 12, 3445. [Google Scholar] [CrossRef]
  180. Le, D.T.; Lutz, E.; Uram, J.N.; Sugar, E.A.; Onners, B.; Solt, S.; Zheng, L.; Diaz, L.A., Jr.; Donehower, R.C.; Jaffee, E.M.; et al. Evaluation of ipilimumab in combination with allogeneic pancreatic tumor cells transfected with a GM-CSF gene in previously treated pancreatic cancer. J. Immunother. 2013, 36, 382–389. [Google Scholar] [CrossRef]
  181. Gjertsen, M.K.; Buanes, T.; Rosseland, A.R.; Bakka, A.; Gladhaug, I.; Soreide, O.; Eriksen, J.A.; Moller, M.; Baksaas, I.; Lothe, R.A.; et al. Intradermal ras peptide vaccination with granulocyte-macrophage colony-stimulating factor as adjuvant: Clinical and immunological responses in patients with pancreatic adenocarcinoma. Int. J. Cancer 2001, 92, 441–450. [Google Scholar] [CrossRef]
Figure 1. Pancreatic tumor microenvironment remodeling strategies to improve therapeutic efficacy. Uncontrolled proliferation of pancreatic cancer cells within desmoplastic stroma, established by excessive deposition of ECM components such as collagen, hyaluronan, and fibronectin, results in the accumulation of mechanical forces and collapsed blood vessels. As a result, abnormal vascularization and ECM stiffness impair vessel perfusion, tumor oxygenation, and drug delivery. In addition, this hinders the infiltration of cytotoxic immune cells, such as CD8+ T and NK cells, and creates a highly immunosuppressive TME along with the secretion of inflammatory cytokines and chemokines by iCAFs, Tregs cells, MDSCs, Bregs, and M2-type TAMs. ECM remodeling strategies using anti-fibrotic drugs and reprogramming of the immunosuppressive TME using immunomodulatory agents could be used synergistically to reverse this phenomenon. This combinatorial approach could normalize tumor vasculature and enhance vessel perfusion and oxygenation, followed by increased infiltration of CD8+ T and NK cells and a decrease in immunosuppressive Tregs, MDCSs, and iCAFs, as well as polarization of TAMs towards an anti-tumor M1 phenotype. Collectively, these strategies aim to enhance pancreatic anti-tumor immunity and the efficacy of ICBs in combination with chemo- and/or nanotherapy to significantly improve pancreatic cancer immunotherapy and patient outcomes. Created with BioRender.com. TME: tumor microenvironment; ICBs: immune checkpoint blockers; M1 TAMs: M1 type tumor-associated macrophages; M2 TAMs: M2 type tumor-associated macrophages DCs: dendritic cell; MDSCs: myeloid-derived suppressor cells; NK cells: natural killer cells; Treg cells: regulatory T cells; PDAC cells: pancreatic ductal adenocarcinoma cells; myCAFs: myofibroblastic cancer-associated fibroblasts; iCAFs: inflammatory cancer-associated fibroblasts; CD4+ T cells: CD4+ cytotoxic T cells, CD8+ T cells: CD8+ cytotoxic T cells; Breg cells: regulatory B cells.
Figure 1. Pancreatic tumor microenvironment remodeling strategies to improve therapeutic efficacy. Uncontrolled proliferation of pancreatic cancer cells within desmoplastic stroma, established by excessive deposition of ECM components such as collagen, hyaluronan, and fibronectin, results in the accumulation of mechanical forces and collapsed blood vessels. As a result, abnormal vascularization and ECM stiffness impair vessel perfusion, tumor oxygenation, and drug delivery. In addition, this hinders the infiltration of cytotoxic immune cells, such as CD8+ T and NK cells, and creates a highly immunosuppressive TME along with the secretion of inflammatory cytokines and chemokines by iCAFs, Tregs cells, MDSCs, Bregs, and M2-type TAMs. ECM remodeling strategies using anti-fibrotic drugs and reprogramming of the immunosuppressive TME using immunomodulatory agents could be used synergistically to reverse this phenomenon. This combinatorial approach could normalize tumor vasculature and enhance vessel perfusion and oxygenation, followed by increased infiltration of CD8+ T and NK cells and a decrease in immunosuppressive Tregs, MDCSs, and iCAFs, as well as polarization of TAMs towards an anti-tumor M1 phenotype. Collectively, these strategies aim to enhance pancreatic anti-tumor immunity and the efficacy of ICBs in combination with chemo- and/or nanotherapy to significantly improve pancreatic cancer immunotherapy and patient outcomes. Created with BioRender.com. TME: tumor microenvironment; ICBs: immune checkpoint blockers; M1 TAMs: M1 type tumor-associated macrophages; M2 TAMs: M2 type tumor-associated macrophages DCs: dendritic cell; MDSCs: myeloid-derived suppressor cells; NK cells: natural killer cells; Treg cells: regulatory T cells; PDAC cells: pancreatic ductal adenocarcinoma cells; myCAFs: myofibroblastic cancer-associated fibroblasts; iCAFs: inflammatory cancer-associated fibroblasts; CD4+ T cells: CD4+ cytotoxic T cells, CD8+ T cells: CD8+ cytotoxic T cells; Breg cells: regulatory B cells.
Ijms 25 09555 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Poyia, F.; Neophytou, C.M.; Christodoulou, M.-I.; Papageorgis, P. The Role of Tumor Microenvironment in Pancreatic Cancer Immunotherapy: Current Status and Future Perspectives. Int. J. Mol. Sci. 2024, 25, 9555. https://doi.org/10.3390/ijms25179555

AMA Style

Poyia F, Neophytou CM, Christodoulou M-I, Papageorgis P. The Role of Tumor Microenvironment in Pancreatic Cancer Immunotherapy: Current Status and Future Perspectives. International Journal of Molecular Sciences. 2024; 25(17):9555. https://doi.org/10.3390/ijms25179555

Chicago/Turabian Style

Poyia, Fotini, Christiana M. Neophytou, Maria-Ioanna Christodoulou, and Panagiotis Papageorgis. 2024. "The Role of Tumor Microenvironment in Pancreatic Cancer Immunotherapy: Current Status and Future Perspectives" International Journal of Molecular Sciences 25, no. 17: 9555. https://doi.org/10.3390/ijms25179555

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop