Next Article in Journal
Nano-Hydroxyapatite Composite Scaffolds Loaded with Bioactive Factors and Drugs for Bone Tissue Engineering
Next Article in Special Issue
Phenotype Switching and the Melanoma Microenvironment; Impact on Immunotherapy and Drug Resistance
Previous Article in Journal
Carboxyl Functionalization of N-MWCNTs with Stone–Wales Defects and Possibility of HIF-1α Wave-Diffusive Delivery
Previous Article in Special Issue
Tissue Biomarkers Predicting Lymph Node Status in Cutaneous Melanoma
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Shaping the Future of Immunotherapy Targets and Biomarkers in Melanoma and Non-Melanoma Cutaneous Cancers

1
Princess Margaret Cancer Centre, University Health Network, Toronto, ON M5G 2C1, Canada
2
Mount Sinai Hospital, University Health Network, Toronto, ON M5G 1X5, Canada
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(2), 1294; https://doi.org/10.3390/ijms24021294
Submission received: 30 November 2022 / Revised: 4 January 2023 / Accepted: 5 January 2023 / Published: 9 January 2023

Abstract

:
Recent advances in treating cutaneous melanoma have resulted in impressive patient survival gains. Refinement of disease staging and accurate patient risk classification have significantly improved our prognostic knowledge and ability to accurately stratify treatment. Undoubtedly, the most important step towards optimizing patient outcomes has been the advent of cancer immunotherapy, in the form of immune checkpoint inhibition (ICI). Immunotherapy has established its cardinal role in the management of both early and late-stage melanoma. Through leveraging outcomes in melanoma, immunotherapy has also extended its benefit to other types of skin cancers. In this review, we endeavor to summarize the current role of immunotherapy in melanoma and non-melanoma skin cancers, highlight the most pertinent immunotherapy-related molecular biomarkers, and lastly, shed light on future research directions.

1. Introduction

The unprecedented survival gains observed with immunotherapy in cutaneous melanoma, over the course of just one decade, are demonstrated by the improvement of 1-year survival rate of 25% to more than 50% for patients with metastatic disease [1,2]. Immunotherapy aims at circumventing the immune evasion mechanisms employed by cancer cells [3], which exploit immune checkpoint receptor pathways to decrease immune activation [4,5,6,7]. Cytotoxic T lymphocyte antigen-4 (CTLA-4) and programmed cell death protein 1 (PD-1) are two critical receptors expressed on T lymphocytes that, upon ligand binding, trigger a signaling cascade that inhibits T-cell activation and limits immune stimulation [8,9]. Antibodies against these receptors (such as ipilimumab, nivolumab, pembrolizumab) prevent receptor-ligand interaction, “releasing the brake” on the immune response [10,11]. Immune checkpoint inhibitors (ICIs) that abrogate CTLA-4 and PD-1 are now the standards of care for advanced melanoma. The observed efficacy in the metastatic/advanced setting has naturally led the way to exploring the efficacy of immunotherapy in earlier stages of melanoma.
In this review, we summarize the current state of play of immunotherapy treatments in melanoma and non-melanoma skin cancers. We also present evidence underpinning the therapeutic inhibition of novel immune markers. Lastly, we also touch upon preclinical and clinical molecular signatures that can guide decision making during immunotherapy treatment.

2. State of Play of Immunotherapy in Melanoma and Non-Melanoma Skin Cancers (NMSC)

2.1. Melanoma

Current standards in adjuvant/metastatic setting. In advanced/metastatic melanoma, which mainly encompasses unresectable stage III and stage IV disease (excluding patients with brain metastases), the most recently mature data show that treatment with a combination of anti-CTLA-4/anti-PD-1 (ipilimumab/nivolumab) confers a median overall survival of 72 months, which is almost double that of nivolumab monotherapy (36.9 months) and certainly incomparably superior to that of ipilimumab monotherapy (19.9 months) [12]. Importantly, most patients treated with ipilimumab/nivolumab (77%) or nivolumab monotherapy (69%) remain treatment-free at 6.5 years of follow-up after treatment, demonstrating significant durability of disease response. Moreover, the benefits of combined anti-CTLA-4/anti-PD-1 inhibition were similar for patients, irrespective of their BRAFV600 mutational and PD-L1 expression status. The success of ICI treatment in advanced melanoma, has raised the unavoidable question of treatment duration of maintenance anti-PD-1 inhibition [13], especially for those patients who experience complete and/or partial radiological response. There is certainly compelling data that in patients with complete responses to anti-PD-1, radiological disease response persists even after treatment discontinuation, especially in those who received at least 6 months of treatment [14,15]. However, the rate of disease relapses and subsequent responses to treatment re-challenge for those who decide to discontinue treatment varies between published reports [13,15,16]; Therefore there is still controversy on the subject of elective discontinuation after achievement of radiological “no evidence of disease” and studies examining the question of continuous versus intermittent anti-PD-1 treatment are on-going [17].
Despite the practice-changing outcomes of the pivotal CheckMate 067 study, the question on ICI-activity in patients with brain metastases [stage IVd by American Joint Committee on Cancer (AJCC) 8th edition] remained; these patients were excluded from study participation at the time of protocol inception [18]. Since then, a series of studies have attempted to answer this question. Collective results suggest that treatment with anti-CTLA-4, anti-PD-1, or the combination thereof, is active in patients with intracranial metastases, offering the greatest benefit to those who are asymptomatic of the malignant infiltration at the time of presentation [19,20,21,22,23]. Intracranial responses to ipilimumab/nivolumab were found to be as high as 53.5% by Tawbi et al. in patients whose metastases has not been previously treated with brain radiotherapy [19]. Two randomized studies in patients with brain metastases treated with immunotherapy have been reported so far. Long et al. found that ipilimumab/nivolumab was significantly superior to nivolumab with a 5-year survival rate of 51% compared to 34%, in patients with asymptomatic disease [20], whereas in a study by Di Giacomo et al., ipilimumab/nivolumab conferred better survival at 4 years, compared to chemoimmunotherapy (41% vs. 10%, p = 0.015) [21]. Outcomes appear to be inferior for patients with established neurological symptoms secondary to their metastases and hence, these patients frequently require upfront stereotactic radiosurgery, which can be safely combined with immunotherapy [24,25].
Lastly, another critical question surrounding the combination of anti-CTLA-4/anti-PD-1 treatment continues to be the optimal ipilimumab dosing. The pivotal CheckMate-067 study that led to the approval of ipilimumab/nivolumab as first line treatment in advanced/metastatic melanoma employed the dosing of 3 mg/kg of ipilimumab and 1 mg/kg for nivolumab (IPI3 + NIVO1), during the first 4 cycles of co-administration [26]. CheckMate-511 study was the first, large-scale randomized study exploring a lower dose of ipilimumab (1 mg/kg), alongside a higher dose of nivolumab (3 mg/kg), IPI1 + NIVO3, in an attempt to compare this combination to the approved IPI3 + NIVO1 dosing [27]. Although not powered enough to prove efficacy, the study demonstrated that, descriptively, efficacy measures such as overall response rate (ORR) response, progression-free survival (PFS), and overall survival (OS) were similar between groups, but importantly, toxicity was significantly lower with the IPI1 + NIVO3 regimen [27,28]. Longer follow-up data may be required to prove the non-inferiority of the lower ipilimumab dose regimen and potentially a change in clinical practice, to spare patients from the high rate of immune-related adverse events.
Attempts to augment ICI efficacy in advanced/metastatic melanoma have been made with the addition of targeted treatment in patients who have actionable mutations [29,30,31,32]. Approximately 50% of patients with metastatic melanoma harbor somatic mutations in the B-Raf proto-oncogene (BRAF) [33], which leads to constitutive activation of the BRAF oncoprotein [34,35]. Currently, targeted treatment with combined BRAF/MEK inhibition can be used in clinical practice for those patients with aberrantly activated signaling pathways downstream of the BRAF protein [36,37]. Concomitant treatment of BRAF/MEK inhibitors (vemurafenib/cobimetinib) with the anti-PD-L1 antibody, atezolizumab, gained approval by the US Food and Drug Administration (FDA) agency in 2020, based on the results of IMspire150 study [31]. In this study, the addition of atezolizumab to targeted therapy led to a significantly increased progression-free survival (PFS) in patients with somatic BRAFV600 mutation-positive advanced melanoma, when compared to treatment with targeted treatment alone. Further maturing of the data from IMspire150 showed a continuous PFS benefit, however, median survival gains have not reached statistical significance and hence, the combination of targeted and ICI treatment has not been met with excitement, so far.
Additional studies combining targeted treatment with anti-PD-1 based immunotherapy (either pembrolizumab or spartalizumab), only yielded modest benefits from the combination, at the expense of heightened toxicity [29,30]. Overall, given the limited number of available systemic treatments for metastatic melanoma, the hypothesis of combining all active treatments into one line has now been replaced with questions on the optimal sequencing of these available therapies [38,39,40]. Indeed, recently published data suggest that for patients with actionable BRAF mutations, prioritizing treatment with ICI as first line of treatment, followed by targeting BRAF/MEK only upon disease progression leads to survival gains for this subset of melanoma patients [38,39].
With the rapid pace of developments in the therapeutic field of melanoma, prior immunostimulatory treatments such as inteleukin-2 (IL-2) were superseded by immune checkpoint inhibitors. Although overall response rates with systemic IL-2 are modest (<20%), patients achieving complete response with IL-2 tend to experience remarkably durable benefits and high rates of long-term survival, sometimes exceeding the 10-year mark [41,42,43,44]. Interestingly, a retrospective comparative study found that response rates to high-dose IL-2 are greater among patients treated with prior ipilimumab compared to patients with no prior exposure to ICI therapy (ORR 21% vs. 12%) [45].
Adjuvant/neo-adjuvant setting. For high-risk, early-stage melanoma, the benefit of adjuvant immunotherapy was underpinned by a seminal study by Eggermont et al. in which the CTLA-4 inhibitor ipilimumab was found to increase the 3-year recurrence free rate by more than 10% ((46·5% (95% CI 41.5–51.3) vs. 34.8% (30.1–39.5)) compared to placebo [46]. This study included stage IIIa/b/c without in-transit metastases, based on the former AJCC staging version (7th edition), whereby the IIIa subgroup included patients with thicker melanoma and of higher risk prognosis compared to the newer AJCC 8th edition. Importantly, this study also highlighted the need to further define the risk-benefit ratio of adjuvant immunotherapy in the face of unacceptably high toxicity rate with adjuvant anti-CTLA-4 treatment, which was administered at 10 mg/kg, a higher than usual dose. Drug-related deaths due to colitis and myocarditis were observed at this dose of ipilimumab and since then, subsequent studies focused on examining single-agent immunotherapy with anti-PD-1 inhibitors following definitive surgery. Instead of adopting anti-CTLA-4 as adjuvant treatment, anti-PD-1 treatment with nivolumab [47] and later on with pembrolizumab [48], were established as effective strategies against disease recurrence, with improved toxicity profiles compared to ipilimumab. It is worth noting that on the one hand, nivolumab was tested against high dose ipilimumab in stage IIIb/IIIc/IV (AJCC 7th edition) population of patients who underwent complete tumor resection. On the other hand, pembrolizumab was tested against a placebo and in the stage III patient population that also included stage IIIa melanoma (with a lymph node metastasis of more than 1 mm in dimension), a sub-cohort of patients not included in the study of adjuvant nivolumab versus ipilimumab [47].
An exciting new step forward for early-stage melanoma was the recent FDA approval of pembrolizumab as adjuvant therapy for resected stage IIb and IIc melanoma (in accordance with 8th AJCC edition).This is melanoma without lymphatic dissemination which is either between 2–4 mm thick and ulcerated, or thicker than 4 mm irrespective of ulceration. The approval was based on an improvement in recurrence-free survival in KEYNOTE-716 study with a 35% reduction in risk of recurrence with pembrolizumab, compared to placebo (hazard ratio (HR) 0.65 [95% CI 0.46–0.92]; p = 0.0066) [49].
Aside from the adjuvant/metastatic settings, current research in the sphere of neo-adjuvant immunotherapy is expected to soon change the treatment paradigm for early-stage high-risk melanoma. Based on the hypothesis that the presence of tumor (and its antigenicity), as well as the presence of peritumoral lymphocytes, will support neoadjuvant ICI mediate a robust anti-tumor response before tumor resection, neo-adjuvant ICI is being examined in a series of studies [50,51,52]. From a clinical standpoint, ICI administration before surgery can (a) reduce tumor burden and hence surgical morbidity, (b) determine therapy efficacy in an individual patient basis for possible additional adjuvant therapy, and (c) use the pathological response as a surrogate biomarker of ICI efficacy and event-free survival. Initial reports from the OpACIN study showed that only two cycles of neo-adjuvant ICI (with a combination of ipilimumab/nivolumab) can augment the expansion of T cell clones and upregulate favorable immune gene signatures that lead to pathological complete responses in the tumor microenvironment [50]. At the ICI doses used in this study; however, the toxicity rate was considerably high and superiority over adjuvant approach was not established. Subsequently, in the OpACIN-neo study which examined a lower dose of ipilimumab and a higher dose of nivolumab, pathological responses were found to translate into durable clinical benefit with 97% of patients with complete responses being recurrence-free at 2 years following diagnosis [52]. In the most recently reported SWOG S1801 randomized study of neo-adjuvant anti-PD-1 blockade, followed by surgery and further adjuvant anti-PD-1, versus adjuvant anti-PD-1 alone in stage III melanoma, recurrence-free survival was 72% at 2 years with the neo-adjuvant approach compared to 49% with adjuvant only treatment [53]. This recent data is thought compelling enough to change the treatment paradigm of early-stage melanoma and crystalize the role of immunotherapy in the neo-adjuvant space.
Another interesting approach into maximizing the anti-cancer immune response in the neo-adjuvant setting, by capitalizing on the presence of the tumor and its antigenicity, is the addition of intralesional administration of oncolytic virus directly into the immune microenvironment. Talimogene laherparepvec (T-VEC) is a genetically modified herpes simplex type 1 virus that selectively infects and replicates tumor cells, whilst expressing granulocyte macrophage colony-stimulating factor (GM-CSF) which, in turn, stimulates antigen-presenting cells to initiate and propagate lymphocyte activation. T-VEC is currently used as monotherapy in Europe, US, and Australia for the treatment of unresectable IIIb, IIIc, and IV M1a stage melanoma (no visceral metastases), based on the results of the OPTiM study, which showed a 19% durable response rate (≥6 months) and a median OS of 23.3 months, at its final analysis [54]. Despite the fact that the addition of T-VEC to standard of care ICI with Pembrolizumab was not found to elicit a synergistic effect in ICI-naïve patients with stage IIIb-IVM1c melanoma [55], there still remains the hypothesis that T-VEC can synergize with systemic ICI in the neo-adjuvant setting to achieve local clearance and increase recurrence-free survival following surgical resection. The results of the recently registered NIVEC [56] trial, whereby nivolumab will be combined with T-VEC pre-operatively, will be greatly anticipated.
New promising treatments/future standards. Unsurprisingly, given the success of immune checkpoint inhibitors in melanoma, current research directions are focusing on the discovery of novel immune checkpoint molecules that could augment the efficacy of ICI in melanoma even further, or reverse the emergence of resistance to the available immune checkpoint blockers.
Lymphocyte-activation gene 3 (LAG-3, also known as CD223), is a cell membrane protein that is upregulated in activated T cells and binds major histocompatibility complex (MHC) class II molecules with high affinity [57,58]. LAG-3 is overexpressed on the membrane of tumor infiltrating lymphocytes and inhibition of LAG-3 results in a more efficient immune-mediated tumor clearance [59,60,61]. The randomized RELATIVITY-047 study which compared the co-formulation of anti-PD-1/anti-LAG-3 antibody, opdualag, versus anti-PD-1 treatment only in patients with untreated advanced melanoma, confirmed that abrogation of these two immune checkpoints is superior to nivolumab alone, with a PFS that is comparable to the combination treatment of anti-CTLA-4/anti-PD-1; although these two combinations have not been directly compared as yet in the same study [62]. Given that the toxicity profile of anti-PD-1/anti-LAG-3 dual inhibition appears more favorable than that of anti-CTLA-4/anti-PD-1, there is a high expectation for this new ICI combination to be soon compared with the standard-of-care anti-CTLA-4/anti-PD-1, in a randomized fashion in advanced melanoma. Anti-LAG-3 inhibition is also being examined in the neo-adjuvant space, almost in parallel with its development in advanced disease, and has already demonstrated exciting results. In a multi-arm study examining different neo-adjuvant regimens (NCT02519322), in patients with resectable clinical stage III or oligometastatic stage IV melanoma, neoadjuvant relatlimab with nivolumab resulted in high pathological complete response (pCR) rate (57%; 95% CI, 37–75%) and improvement in the 2-year recurrence-free survival (RFS) rate in patients who achieved any degree of pathologic response compared to those without a pathologic response (p = 0.005). Neoadjuvant anti-LAG-3/anti-PD-1 treatment mediated tumor infiltration by memory CD4+ and effector CD8+ T cells in the post-treatment tumor specimens of patients with favorable treatment response. Overall, these findings indicate that LAG-3 is now established as the third targetable immune checkpoint in melanoma.
Other immune checkpoint molecules exhibiting signals of activity in melanoma are (i) the T cell immunoreceptor with immunoglobulin and ITIM domain (TIGIT) and (ii) the glucocorticoid-induced TNF receptor (GITR). TIGIT down-regulate T cell and NK cell functions upon binding to its two ligands, CD155 and CD112, which are expressed by both tumor cells and antigen presenting cells [63]. When TIGIT inhibition was combined with anti-PD-1 inhibition in vitro, it was found to enhance tumor antigen specific CD8+ T cell expansion and activity, as compared with anti-PD-1 inhibition alone [64]. Double inhibition also increases proliferation and function of tumor infiltrating lymphocytes (TILs) isolated from patients with melanoma. Studies currently investigating TIGIT inhibition in patients with advanced melanoma are depicted in Table 1. Another target of interest is the glucocorticoid-induced TNF receptor (GITR), a member of the tumor necrosis factor (TNF) receptor superfamily. GITR mediates immunosuppression through constitutive expression on T regulatory cells (Tregs), whereas its expression is upregulated on CD4+, CD8+ T cells, and natural killer (NK) cells, only upon T-cell activation [65,66,67]. Hence, targeting the GITR pathway promotes antitumor activity of T cells through T cell proliferation and enhanced effector function, abrogation of Treg function, as well as by upregulating IL-2 and IFNγ pathways [68,69,70]. There are several GITR inhibitors in clinical development and MK-4166, a humanized IgG1 antibody against GITR showed promising response rates when combined with anti-PD-1 in ICI naïve patients; it did not however exhibit activity in patients with prior exposure to ICI [71].
The critical question needing urgent answer in the field of immune-oncology (IO) in advanced melanoma is undoubtedly that of reversing the emergence of ICI-resistance and discovering novel therapeutic options for patients harboring ICI-resistant disease. Ample pre-clinical research is currently focusing on answering this question and we endeavor to address some of it later in this review. From a clinical standpoint, promising results were recently reported by the LEAP-004 study [72]. In this single-arm phase 2 study, patients who were exposed and whose disease had progressed on anti-PD-1/PD-L1-based immunotherapy, were treated with the combination of pembrolizumab plus lenvatinib. Lenvatinib, a tyrosine kinase inhibitor that selectively inhibits VEGFRs 1–3, FGFRs 1–4, platelet-derived growth factor receptor α, RET, and KIT [73,74], was found to enhance the anti-tumor response of anti-PD-1 blockade in preclinical models [75,76]. In LEAP-004 study, up to 21% of patients experienced disease response, amongst them complete and partial responses [72], supporting the hypothesis that abrogation of pivotal intracellular pathways can reverse tumor resistance to anti-PD-1/PD-L1-based treatments. This hypothesis has generated several combinatorial clinical studies in patients with ICI-resistant melanoma (Table 1).
Immense excitement in the sphere of ICI-resistant melanoma has also been sparked by the efficacy of adoptive cell therapy, in the form of treatment of surgically harvested tumor-infiltrating leucocytes (TILs). Lifileucel (LN-144) is an autologous TIL product that utilizes harvested tumor-tissue T cells, using a centralized manufacturing process. Given that melanoma is a disease characterized by high mutational burden, a unique cellular product that harbors polyclonal cells with diverse antigen specificity is capable of offering a tailored immune response for each patient. In a study reported by Sarnaik et al., lifileucel was administered to patients with unresectable or metastatic melanoma, previously progressed through anti-PD-1 blockade (and targeted treatment if they had BRAF V600 mutation-positive disease) [77]. With an overall response rate of 36% and disease control rate of 80%, these results indicate that once the practices of administering cellular therapies can be streamlined, adoptive cellular therapy will undoubtedly become a standard of care for melanoma and a robust option, at least for patients with ICI-resistant disease and available resectable tumor present.
Non-cutaneous melanoma, current, and future therapeutic pathways. Non-cutaneous melanoma including uveal and mucosal melanoma are rare tumors, characterized by specific epidemiology, biological behavior, and molecular profile. Uveal melanoma represents only 5% of all cases of melanoma; however, it is the most common primary intra-ocular cancer [78]. While cutaneous melanoma is often characterized by BRAF, NRAS and KIT mutations, uveal melanoma typically harbors GNAQ/GNA11, EIF1AX, SF3B1, BAP1 mutations and chromosomal abnormalities [79]. In more than 90% of the cases uveal melanoma is diagnosed at an early stage, when still amenable to local treatment [80]. Unfortunately, the proportion of patients who develop metastatic disease ranges from 20% for stage I to 70% for stage III [81,82]. The liver is the most frequent site of metastasis and when feasible, liver-directed treatments including surgery, radio-ablation and chemoembolization represent the treatment of choice for metastatic disease [83]. For patients with metastatic disease that is not amenable to local treatment, immunotherapy currently represent the mainstay of therapeutic options. Immune checkpoint inhibitors targeting CTLA-4 and the PD-1/PDL-1 axis have demonstrated antitumor activity in uveal melanoma. However, when compared to cutaneous melanoma, these agents have limited efficacy, resulting in an ORR between 10 and 20% when used in combination [84,85,86,87] and even lower when used as single agents [88,89,90]. Anatomical, genomic, and immunological factors are known to contribute to ICI resistance in uveal melanoma. The eye is an immune privileged organ characterized by multiple local and systemic mechanisms limiting and preventing inflammation to protect visual function. The immunotolerant microenvironment of the eye, enriched on Tregs, may impair the efficacy of ICI on the primary tumor [91]. As compared with its cutaneous counterpart uveal melanoma is characterized by lower molecular burden, with a mutation rate of respectively 0.5 mutations per megabase (Mb) for uveal melanoma, as opposed to 49.2 mutations per Mb for cutaneous melanoma [92,93,94]. Another molecular feature responsible for the lack of responsiveness to immunotherapy is the low PD-L1 expression [95,96,97]. Epigenetic changes downregulating PD-L1 and other key immune checkpoints, including LAG-3 and CTLA-4, have been observed in uveal melanoma with BAP1 disruption [98]. These observations supported the development of novel agents exploiting alternative targets and the investigation of combinatorial strategies to overcome the resistance to anti-PD-1 and anti-CTLA-4 agents.
Tebentafusp, a bispecific gp100 peptide-HLA-directed CD3+ T lymphocyte engager has been recently approved by FDA for the treatment of patients with metastatic uveal melanoma, who harbor the HLA-A*02:01 allele. This approval followed the results of a randomized phase III trial in which tebentafusp resulted in prolonged overall and progression-free survival in comparison to investigators’ therapeutic choice (including dacarbazine, anti-CTLA-4, or anti-PD-1 treatment) [99]. At 1 year, the overall survival of patients receiving tebentafusp was 73% vs. 59% in the control group (HR for death 0.51; p < 0.001), while progression-free survival at 6 months was 31% vs. 19% (HR for disease progression 0.73; p = 0.01).
Different therapeutic strategies are under investigation to enhance the effect of immune checkpoint inhibitors. The histone deacetylase inhibitor entinostat was used in combination with pembrolizumab in twenty-nine uveal melanoma patients resulting in an ORR of 14%, PFS of 2.1 months, and OS of 13.4 months [100]. The clinical benefit correlated with BAP1 status with a higher chance of response and prolonged survival in patients with BAP1 wild type. Clinical trials are ongoing testing the efficacy of other combinatorial regimens, such as with anti-angiogenic agents (NCT05282901, NCT05308901), anti-LAG-3 (NCT04552223), TLR9 agonists (NCT04935229), PARP inhibitors (NCT05524935), and liver-directed radio-embolization (NCT02913417) (Table 2).
Mucosal melanoma, another rare subset of melanoma, accounts for 1 to 3% of all melanoma diagnoses. It can originate from the gastrointestinal, genitourinary, and respiratory tracts; however, 80% of mucosal melanomas primaries arise from the head and neck area [101,102]. Mucosal melanoma has a peculiar genomic profile characterized by higher rate of SF3B1 and KIT and lower incidence of BRAF and NRAS mutations [103]. Similarly, to uveal melanoma, mucosal melanoma is characterized by low PD-L1 expression and tumor mutational burden that blunt the efficacy of ICIs [104,105]. Anti-PD-1 antibodies alone or in combination with anti-CTLA-4 represents the main treatment option; however, clinical data indicate that these agents’ efficacy is modest when compared to cutaneous melanoma [106,107]. Similarly to the other melanoma subtypes, the combination of anti-PD-1 and anti-CTLA-4 seems to improve patients’ outcomes over monotherapy. Seventy-nine patients with mucosal melanoma were included in the CheckMate-067 study and randomized to the combination of ipilimumab/nivolumab, nivolumab or ipilimumab monotherapies. In this population the combination resulted in a superior ORR (43% vs. 30% with nivolumab and 7% with ipilimumab monotherapy) and improved 5-year PFS and OS rates (5-year PFS rate: 29% vs. 14% vs. 0% and 5-year OS rate 36% vs. 17% vs. 7%) [108]. Vascular endothelial growth factor (VEGF) upregulation is a known biomarker of poor prognosis in mucosal melanoma [109] and clinical data support the combination of antiangiogenic and immunotherapeutic agents in this population. A phase 1 study testing anti-PD-1 blockade in combination with the VEGF inhibitor axitinib in 33 mucosal melanoma patients resulted in an ORR of 48% and a median PFS and OS of 7.5 and 20.7 months, respectively [110,111]. Interestingly, the biomarker analysis revealed no correlation between PD-L1 expression or tumor mutational burden and survival outcomes while an association was observed between improved PFS and a gene-expression profile signature including angiogenesis and immune-related genes [111]. The potential role of anti-VEGF inhibition and ICI in mucosa melanoma is supported also by the results of a phase II study testing atezolizumab and bevacizumab [112]. Forty-three mucosal melanoma patients were treated with this combination, achieving an ORR of 43%, with a median PFS of 8.2 months. Further agents currently under evaluation in clinical trials testing combination strategies with ICIs (Table 2).

2.2. Non-Melanoma Skin Cancers Overview

Non-melanoma skin cancers (NMSCs) consist of a heterogeneous group of tumors. It was not until recently that immunotherapy treatment was incorporated into the therapeutic landscape of NMSCs. In this review, we mostly discuss the most common types such as cutaneous squamous cell carcinoma (cSCC), Merkel cell carcinoma (MCC), and basal cell carcinoma (BCC). High curability with surgery and/or radiation results in a proportionally low death rate, however, due to very high incidence, accounting for 30% of all diagnosed cancer types, the absolute mortality in advanced stages is comparable to melanoma [113]. NMSCs harbor features predictive of response to immunotherapy, including high tumor mutational burden (TMB) associated with chronic ultraviolet radiation exposure, and other common risk factors like immunosuppression, as well as viral etiology (Merkel cell polyomavirus in MCC) and advanced patients age [114,115,116]. In the past decade, multiple clinical trials confirmed the efficacy of ICIs in NMSCs, mainly in SCC and MCC [117,118,119,120,121]. Moreover, more recently, immunotherapy exhibited clinically meaningful antitumor activity in advanced BCC following progression on first-line hedgehog inhibitor therapy (HHI) [122].
Cutaneous SCC (cSCC): Use of single-agent anti-PD-1 inhibition with either pembrolizumab or cemiplimab as an upfront first-line therapy for advanced or metastatic cSCC, that is not amenable to surgery or radiation, showed ORR of 42% and 52% in recurrent/metastatic disease and locally advanced disease, respectively [117,118]. Both drugs demonstrated durable responses [median duration of response (DOR) not reached] and extended OS (median not reached), along with tolerable toxicity profile. Favorable toxicity profile is very pertinent to this patient population, given the advanced age and high comorbidity prevalence associated with it. These practice-changing results led to the introduction of immunotherapy as a new 1st line standard of care treatment in patients with advanced SCC [123]. Following encouraging results in advanced disease, multiple studies are currently investigating the role of ICI in adjuvant and neoadjuvant/perioperative settings (Table 3). Indeed, Gross et al. recently reported results of neoadjuvant cemiplimab in stage II to IV cSCC. The study demonstrated a 68% objective response rate, as well as a 51% pathological complete response rate, and a 13% pathological major response rate [124].
Merkle Cell Carcinoma (MCC): MCC is a rare and aggressive malignancy with historical 5-year OS rate of 14% for metastatic disease and 27% for clinically detected locally advanced disease (nodal involvement) [125]. Similarly to cSCC, MCC is a highly immunogenic disease [114,116], representing a very good target for ICI-based immunotherapy. JAVELIN Merkel 200 trial demonstrated ORR of 33.0% and a median duration of response of 40.5 months with avelumab for pre-treated patients with advanced MCC. Reported median OS was 12.6 months [126]. Subsequently, pembrolizumab and avelumab were separately investigated in first line treatment settings, with ORR of 56–62% and complete response rate of up to 24% [120,121]. Median duration of response and median OS are not reported yet, whilst the Kaplan Meier estimate of median OS for 1st line pembrolizumab at 3 years was 59.4% for all patients, and 89.5% for responders [127]. Multiple trials are currently on-going to investigate efficacy of ICI as monotherapy or in combination with targeted agents in neoadjuvant and adjuvant settings, along with an attempt to overcome primary and acquired resistance to ICI in advanced MCC (Table 3).
Basal cell carcinoma (BCC): BCCs are at least twice more common than cSCC. Compared to cSCC or MCC, BCCs are much less likely to metastasize, with a reported metastatic rate of less than 0.1% [128]. Locally advanced disease is more common, usually causing significant destruction and deformation of the underlying tissues, including soft tissue, bone, and cartilage [129]. Patients with recurrent and destructive infiltration of the surrounding tissues most commonly require systemic treatment. Among published studies, Stratigos et al. recently reported that cemiplimab can induce a significant antitumor response in patients with locally advanced BCC, after progression on HHI therapy. The study reports 31% ORR with the survival data still being immature. Surprisingly, the rate of grade 3–4 adverse events was quite significant, reaching 48%, which might become a serious challenge in this vulnerable patient population [122]. Similarly to MCC and cSCC, multiple trials are ongoing for the evaluation of the ICI role in other treatment settings (Table 3).
The approval of immunotherapy-based treatment for both melanoma and NMSCs in a timeline fashion is depicted in Figure 1.

3. Biomarkers of Immunotherapy Response/Resistance in Melanoma

Melanoma has immensely contributed to our understanding of the mechanisms of response or resistance to ICI. However, unlike other tumor types such as lung or head & neck cancer, where PD-L1 expression by immunohistochemistry (IHC) is used to guide patient stratification to ICI treatment, PD-L1 expression by IHC has not been proven to be a reliable predictive biomarker in melanoma, as evidenced by exploratory analyses of large scale studies [130,131,132]. Tumor mutational burden (TMB), on the other hand, is a universally accepted surrogate marker of ICI response in a tumor-agnostic fashion. Melanoma has inherently a high level of TMB, showing superior responses to ICI depending on the magnitude of the TMB [133,134] but notwithstanding this, it is not routinely utilized in the clinic as a predictor of ICI response. Given the extremely favorable response of melanoma to ICI, a biomarker of ICI-resistance is perhaps more critical in the clinical management of melanoma, rather than a biomarker of response. As we gain more insight into multi-dimensional data on intratumoral, immunological, and systemic factors modulating the anti-cancer immune response, more and more host intrinsic as well as extrinsic players are starting to emerge (Figure 2).
Genomic markers. Loss of PTEN function has been found to have a negative correlation with response to immunotherapy. Both in preclinical work but also in patient samples interrogated through the Cancer Genome Atlas Program (TCGA), PTEN loss correlates with decreased T-cell infiltration at tumor sites and increased expression of immunosuppressive cytokines [135,136,137]. Whole exome analysis of the resistant tumor clones identified loss-of-function mutations in the genes encoding interferon-receptor-associated Janus kinase 1 (JAK1) or Janus kinase 2 (JAK2), which leads to insensitivity to INFα, β and γ [138,139]. Activating mutations in the Wnt/b-catenin pathway can also induce resistance to ICI through altering the expression of PD-L1 and PD-L2 in a broad group of tumors [140,141] and mechanistic studies specifically in melanoma have revealed that Wnt-induced decrease in the expression of the chemokine CCL4 hindered the recruitment of CD103+ DCs and T cells to the tumor microenvironment [142].
DNA mismatch repair deficiencies (dMMR) and consequently microsatellite instability (MSI) predispose tumor cells to the accumulation of somatic mutations and increased TMB [143]. The connection between novel somatic mutations and the generation of neo-antigens is extremely intriguing and nuanced, at the same time. In melanoma specifically, it has been demonstrated that the generation and landscape of clonal neoantigens can determine T cell infiltration and durable clinical benefit for patient [144]. Moreover, large genomic aberrations, such as somatic copy number alterations (SCNAs), have also been found to mediate immune evasion [145]. Loss of 9p21.3 locus in a pan-cancer cohort, including patients with melanoma is linked to reduced ICI responsiveness and poorer clinical outcomes [146,147]. Specifically in melanoma, high dosage of arm- or chromosome-level SCNAs were associated with poorer response to anti-CTLA-4 in a retrospective analysis [148], one such example being gains in chromosome 7 that are accompanied by poor lymphocyte infiltrate and aberrant neutrophil activation [149].
Transcriptomic markers. Transcriptomic readouts have also been used in an attempt to predict responsiveness to immunotherapeutic modification with ICI treatment. These mainly pertain to immune gene signatures and characterize the immune tumor microenvironment more accurately. Melanoma can be therefore stratified into two different groups: a T-cell specific, antigen presentation-related, and IFNγ signaling–related signature helps identify the immune “hot” tumors [150,151,152]. On the contrary, signatures encompassing markers related to higher prevalence of myeloid-derived suppressive cells (MDSC), tumor-associated macrophages (TAMs), or cancer-associated fibroblasts (CAFs) associated genes are more specific to immune “cold” tumors, and correlate with reduced cytotoxic lymphocyte activity [139,153,154].
Circulating markers of inflammation have also been interrogated to discover links between biomarkers and response to ICI in melanoma. Recently, Rossi et al. demonstrated that the levels of interleukin-6 (IL-6), hepatocyte growth factor (HGF), and monocyte chemotactic protein 2 (MCP-2), found in the serum of patients with melanoma whose disease did not respond to ICI, are higher compared to ICI-responders [155]. The three chemokines could be clustered in a signature with an overall negative effect on ICI response. Transforming growth factor-β (TGFβ) is a cytokine that plays important roles in angiogenesis and immunosuppression by stimulating Tregs, promoting tumor immune escape and immunotherapy resistance [156]. High expression of TGFβ1 in tumor frequently correlates with inferior survival, even when cytotoxic T lymphocyte infiltration is rich [153,154].
Recent studies also shed light on the role of extracellular vesicles, such as exosomes and microvesicles, in their ability to carry molecules that can modulate the immune response systemically [157,158,159]. Chen et al. showed that metastatic melanomas release exosomes, that carry PD-L1 on their surface and suppress CD8+ T cell effector function and consequently facilitate tumor growth [160]. In patients with metastatic melanoma, treated with pembrolizumab, the pre-treatment level of circulating exosomal PD-L1 by enzyme-linked immunosorbent assay (ELISA) was significantly higher in patients whose disease failed to respond to treatment.
Immunological biomarkers in melanoma. The tumor microenvironment plays a critical role in the anti-cancer immune response following IO treatment [161]. In metastatic melanoma, the co-occurrence of tumor-associated CD8+ T lymphocytes and CD20+ B cells has predictive value in the treatment outcome of patients receiving ICIs [162,163,164]. On the other hand, tumor infiltration by Tregs, TAMs, and MDSCs, as well as accumulation of cancer-associated fibroblasts is commonly associated with immune “cold” tumors and adoptive resistance to CPI [165,166,167]. Organized aggregates of T, B, and dendritic cells form tertiary lymphoid structures (TLS) that participate in the adaptive antitumor immune response. B cell infiltrate and the presence of TLS in tumor samples in melanoma patients were found to be predictive of response to treatment and associated with improved survival [168,169,170,171,172].
During interrogation of immune gene signatures in RNA sequencing (RNA-seq) data of baseline and on-treatment tumor samples, enrichment of a B cell signature in responders vs. non-responders at baseline and early on-treatment was observed [168]. Furthermore, tumors with low infiltration of B cells had a significantly increased risk of death in comparison to the B-cell-high signature cohort of patients [169]. The density of CD20+ B cells and TLSs and the ratio of TLSs to tumor area were higher in responders than in non-responders, in early on-treatment samples of patients with melanoma treated with neoadjuvant ICI [169]. These findings were also supported by recent analyses of a TCGA cohort that demonstrate an association between a plasmablast-like B cell signature and improved survival, as well as with an increased expression of CD8a and infiltration of CD8+ T cells [170].
Recently reported meta-analysis on whole-exome and transcriptomic data of ICI-sensitive tumors, identified CCR5 and CXCL13 as T-cell-intrinsic markers of ICI sensitivity. CXC chemokine ligand 13 (CXCL13) exclusively binds CXC chemokine receptor type 5 (CXCR5), which plays a critical role in immune cell recruitment and activation and regulation of the adaptive immune response. CXCL13 is a key molecular determinant of the formation of TLSs [173]. Selective expression of CCR5 and CXCL13 in neoantigen-specific T lymphocytes suggests that a key feature of ICI-responsiveness is the ability to sustain ongoing priming and recruitment of tumor reactive T cells supported by CXCR5+ lymphocytes, including both T and B cells, and form TLSs [169].
With regards to genomic biomarkers, signatures encompassing T cell-specific gene expression profiles (GEP) have attracted a lot of interest. A 28-gene set including genes enriched for cytolytic activity (e.g., granzyme A/B/K, PRF1), cytokines/chemokines secretion (CXCR6, CXCL9, CCL5, and CCR5), T cell markers (CD3D, CD3E, CD2, IL2RG [encoding IL-2Rγ]), NK cell activity (NKG7, HLA-E), antigen presentation (CIITA, HLA-DRA), and additional immunomodulatory factors (LAG-3, IDO1, SLAMF6). A signature including all the above was shown to predict response to anti–PD-1–directed therapy in melanoma patients, and it appears to be a promising multi-dimensional biomarker, currently undergoing validation in larger clinical trials [152].
Dysfunction of antigen-presenting system was confirmed as one of the mechanisms of resistance to CPI, mostly related to HLA class I loss [174]. Β-microglobulin 2 gene (B2M) plays an important role in an assembly of all HLA class I complexes, as well as antigen transport [175]. Mutations in B2M are commonly associated with complete or partial loss of HLA class I function. Alterations in this gene are commonly identified in non-responders, and much less are seen in responders [176]. Accordingly, normal expression of B2M correlates with better overall survival of melanoma patients, treated with immunotherapy [177].
Clonality and diversity of intratumoral and peripheral TCR repertoire have also recently emerged as potential markers of responsiveness to immune checkpoint inhibitors in multiple cancer settings including melanoma [178,179,180,181,182,183,184,185]. Overall, these studies suggest that patients who exhibit a higher diversity of TCR repertoire before treatment and early evolution of the T cell populations after treatment start have a greater chance to benefit from immunotherapy. An association between high pre-treatment clonality of the T cell population and the probability of response to anti-PD-1 antibodies has also been observed [178]. Interestingly the same parameter showed an association with a lower chance of response to anti-CTLA-4 agents in melanoma [178]. This suggests that in the presence of an already expanded and ineffective clone of tumor-specific T cells, the use of anti-CTLA-4 antibodies will not be able to overcome cancer immune escape.
Asides from the intrinsic immune pathways directly connected to shaping the immune response, accumulating evidence underlines the presence of extrinsic factors also playing a pivotal role. Extracellular matrix (ECM) is emerging as another key element that can influence response to anticancer drugs, including immunotherapeutic agents [186]. ECM has indeed been demonstrated to regulate immune cell trafficking and activation, as well as cancer-associated antigen release and presentation [186]. High collagen levels have been associated with CD8+ suppression and exhaustion and with a lower probability of deriving benefit from anti-PD-1/PD-L1 blockade in melanoma and lung cancer [179,187]. Collagen is known to exert an inhibitory effect on T cells through the SHP1 pathway activation, via the leukocyte-specific collagen receptor LAIR1 [188]. These observations indicate ECM remodeling as a potential strategy to overcome resistance to immunotherapy. As an example, Peng et al. demonstrated the possibility to restore sensitivity to anti-PD-1 antibodies by inhibiting intratumoral collagen deposition or blocking the LAIR1-SHP1 pathway [187].
The gut microbiome as a biomarker in melanoma. The gut microbiota influence metabolism, antitumor activity, and toxicity profile of a broad variety of anticancer treatments, including immunotherapy [189,190,191,192,193]. An association between higher microbiota biodiversity and a greater probability to benefit from anti-PD-1/PD-L1 antibodies has been reported for a variety of cancer types, including melanoma [194]. The abundance of specific bacteria populations such as Bifidobacterium and Ruminococcaceae has been described in the gut microbiota of patients benefitting from immunotherapy [194,195]. On the other hand, the microbiota of non-responders appears to be enriched with different strains such as Bacteroidaes and Clostridium species [194,196]. Several studies indicate a correlation between antibiotic-induced gut microbiota dysbiosis and poor clinical outcomes in cancer patients receiving immunotherapy [196,197,198,199,200].
The mechanisms underlying the relationship between microbiota composition and response to immune checkpoint inhibitors are not fully understood. Preclinical evidence indicates that microbiota can induce and activate multiple immune effectors, including NK, dendritic, CD4+, and CD8+ T cells, as well as Tregs and monocytes, through the production of different metabolites such as short-chain fatty acids, inosine, and bile-acid [201,202,203,204,205,206,207,208]. Interestingly, gut microbiota not only can influence the immune cell population but may also have a role in the remodeling of the extracellular matrix [209]. For example, adhesin A secreted by Fusobacterium species, Bacteroides fragilis toxin, and gelatinase E produced by E faecalis can modulate cell-to-cell interaction and mediate tumor progression and metastatic spread [210]. On the other hand, cross-reactivity between cancer cells and microbiota antigens can augment cancer cell immunogenicity and hence, promote ICI efficacy [211,212].
Following these findings, significant efforts have been made to develop effective strategies to correct gut dysbiosis with an aim to prevent or revert resistance to immunotherapy, with the strongest evidence supporting for fecal microbiota transplantation (FMT) so far [202]. The potential role of FMT in overcoming immunotherapy resistance is supported by preclinical and early clinical data. FMT from patients with cancer who responded to immunotherapy demonstrated an increased probability of benefit from anti-PD-1 blockade in xenograft mice [213]. In these models A. muciniphila and E. hirae shown the ability to induce secretion of IL-12 from dendritic cells, supporting immune-surveillance. Oral administration of A. muciniphila demonstrated the capability to revert resistance to immune checkpoint inhibitors in xenografts receiving FMT from non-responders [213]. Two proof-of-concept studies tested the efficacy of FMT administered by colonoscopy as a potential strategy to revert anti-PD-1 resistance in melanoma patients with disease progression following anti-PD-1 treatment [214,215]. FMT resulted in prolonged modification of the microbiota of the recipients and an objective response or prolonged disease stability lasting for more than 12 months was observed in 36% of the patients enrolled in the two studies. In patients who responded to ICI, FMT induced changes in the cytokine profile, with down-regulation of CCL2, IL-8, IL-18, IL12p70, and IFNγ and upregulation of IL-21, CXCL13, IL-5, IL-13, IL-10, TNF, and TRAIL [215]. Unsupervised single cell analysis of blood samples collected at baseline ad at serial time points during treatment was performed in one of the two studies [215]. This analysis demonstrated a higher prevalence of activated CD8+ cells in post-treatment samples from responding patients, characterized by overexpression of TIGIT, Tbet, and LAG-3 and suppression of CD27, on their cell surface. Moreover, responders had a higher percentage of peripheral T memory cells. Single-cell RNA sequencing of tissue samples collected before and after treatment was also extremely informative. This evaluation showed a higher prevalence of myeloid cells and Tregs in post-treatment biopsies from non-responders, whilst an increased expression of MHC class II genes was observed in samples obtained by patients who responded to treatment. Notwithstanding these encouraging results, FMT is an invasive procedure and therefore alternative approaches to modulate microbiome in cancer patients are currently under investigation including the use of probiotics, dietary intervention, and engineered microbiome (NCT04753359, NCT03637803, NCT03686202) [202].

4. Biomarkers of Disease Activity/Treatment Monitoring in Melanoma

ctDNA as a tool. Circulating tumor DNA (ctDNA) is the component of fragmented cell-free DNA (cfDNA) derived from tumor cells. Genomic profiles based on ctDNA are highly concordant with those of primary tumors, and ctDNA levels correlate directly with tumor burden. The detection of ctDNA following radical surgery may identify patients at the highest risk of clinical relapse and provide a “real-time” disease status at the molecular level. Therefore, the role of ctDNA as a non-invasive biomarker of molecular residual disease is widely evaluated in multiple tumor types. Liquid biopsy techniques measuring ctDNA can identify cancer DNA traces at the microscopic level [216,217]. This approach offers the possibility to monitor disease longitudinally and enables timely detection of disease progression or relapse [218].
Quantitative and digital polymerase chain reactions (like digital droplet PCR, ddPCR) were the first ctDNA detection approaches for selective gene targets [219,220]. Although highly validated, ddPCR is limited in its breadth to cover a large number of mutations in a tumor-specific and personalized way for each patient. Application of next-generation sequencing-based techniques is increasingly growing and allows for the detection of multiple mutations per patient and high sensitivity ctDNA tracking [221,222]. Next-generation sequencing (NGS) methods applied in a “tumor-informed” manner (i.e., tissue analysis is conducted to verify tumor specificity of aberrations observed in plasma) have the advantage of providing a signature of putative cancer-derived aberrations through analysis of concordant mutations [223,224]. Highly sensitive assays have permitted novel tracking of dynamic changes during therapy in a more sensitive way compared to ddPCR. For example, in early lung, colorectal, and breast cancers, postoperative detection of ctDNA prognosticates disease recurrence with high accuracy [225,226,227,228,229,230].
ctDNA as disease predictor in melanoma. Independently of standard AJCC staging, pre-operative detection of ctDNA in patients with stage III melanoma is associated with a 3-fold increased likelihood of disease relapse and half the time-to-distant metastatic relapse (6.2 months vs. 13.9 months (HR 1.59; 95% CI 1.0–2.52; p = 0.027)) [231,232]. This translates into a significantly shorter median melanoma-specific survival of 17.6 months compared with 49.4 months in patients with undetectable ctDNA levels (HR 2.11; 95% CI 1.20–3.71, p < 0.01), confirming the validity of ctDNA as a prognostic biomarker in early-stage melanoma. With serial assessments of ctDNA, a retrospective study of patients with stage III melanoma and no radiological evidence of disease, demonstrated that detection of BRAF/NRAS mutant ctDNA (by digital droplet PCR, ddPCR) at a single timepoint within 12 weeks of surgery was associated with worse disease-free interval and OS [233].
The detectability of ctDNA after surgical resection in melanoma patients is low, with detection rates below 25% even when using highly sensitive NGS-based assays [218,233,234,235,236]. Significant correlation has been observed between the presence of detectable ctDNA after surgery and shorter survival outcomes [232,233,234,237]. Despite the limited amount of data some studies seem to indicate that the administration of adjuvant immunotherapy might prolong the progression free survival of patients with positive ctDNA after surgery, nullifying the difference with the ctDNA negative population [232,234].
Perhaps the most robust research on the predictive value of ctDNA in melanoma, both in terms of patient sample size and use of most advanced ctDNA sequencing methodology so far, was conducted and reported as part of the CheckMate 915 study translational analysis. Long et al. confirmed in a cohort of 1127 patients with either resected stage IIIB-D and IV melanoma, that ctDNA detection at baseline can be an efficient predictor of recurrence-free survival for patients undergoing ICI treatment [237]. The predictive value was even more robust when ctDNA detection was combined with tumor TMB level and a tumor-derived IFNγ signature including HLA-DRA, CXCL9, GZMA, PRF1, CCR5, IFNG, CXCL10, IDO1, STAT1, and CXCL11 [152]. Overall, current evidence supports the use of postoperative ctDNA measurement to enable risk stratification of disease recurrence.
ctDNA and immunotherapy response in melanoma. Several studies have also investigated the potential use of ctDNA to predict anticancer therapy efficacy in melanoma patients [238,239]. These studies might be heterogeneous in terms of patient population and methodology used for ctDNA detection; however, they offer some important insights. Baseline levels and dynamic changes of ctDNA quantified with digital droplet PCR or with NGS-based assays were demonstrated to predict survival outcomes and response to immunotherapy in patients with metastatic disease [134,240,241,242]. Interestingly Lee et al. observed that ctDNA quantification might also be a valuable tool to distinguish patients who are having pseudo-progression on ICI from those whose disease is truly progressing [243].
Aside from the quantification of tumor burden and minimal residual disease, ctDNA evaluation can provide important information regarding the molecular features of cancer cells. The use of blood-based assays to evaluate known predictors of immunotherapy sensitivity, such as tumor mutational burden and microsatellite instability is rapidly expanding [244,245,246,247,248,249]. This approach not only overtakes the need for tissue samples but might also overcome the issue of tumor heterogeneity. Furthermore, ctDNA analysis is a promising methodology to study clonal evolution and the mechanisms underlying drug resistance [218,250,251]. Li et al. performed ctDNA analysis in 12 patients with NSCLC treated with pembrolizumab [252]. Serial samples were collected before and during treatment and sequenced against a 329 pan-cancer gene panel. Increasing levels of a PTCH1 mutation were observed in a patient who was developing a new metastatic lesion more than 3 months before disease progression in CT scans. The authors also observed the sequential emergence of two different acquired mutations of B2M in another patient whose disease progressed after achieving an initial response to pembrolizumab. Similarly, Jin et al. used a 425-genes next-generation sequencing panel to profile tumor tissue and blood samples collected before and after treatment in 46 gastric cancer patients treated with anti-PD-1 antibodies [253]. Baseline TGFBR2, RHOA, and PREX2 mutations were identified as predictors of shorter PFS. To investigate potential markers of acquired resistance, blood samples obtained at the time of ICI-resistance emergence from 16 patients were analyzed. FOXL2 gene mutations and copy number variations of FGFR2 were identified as new alterations in two patients and a new RHOA mutation was observed in another patient suggesting possible implication of these genes in driving the newly acquired immune treatment resistance.
Evidence supporting the use of ctDNA to study clonal dynamics exists in melanoma too. Takai et al. reported the results of whole exome sequencing of sequential ctDNA samples collected from 14 patients with metastatic melanoma receiving ICI [254]. Newly emerged ctDNA mutations—such as in ARID1B—were observed at the timepoint of disease progression. Overall, this data showcase ctDNA as a powerful tool to guide treatment decision and inform the design of clinical trials testing novel therapeutic strategies to overcome drug resistance. In a disease like melanoma, where ICI-resistance therapeutic strategies are critically needed, ctDNA can be proven immensely instrumental.

5. Molecular Biomarkers in Non-Melanoma Skin Cancers

Cutaneous squamous cell carcinoma (cSCC) is affecting more than 1 million people in the United States, every year. Incidence is rising due to increasing sun exposure and an aging population. Approximately 2–5% of the patients present with locoregionally advanced cSCC [255]. Ultraviolet exposure, with subsequent DNA damage that leads to a high tumor mutational burden (TMB), as well as a state of immunosuppression, is the main risk factors for cSCC development and interestingly, both factors correlate with superior response to immunotherapy [116,256]. In the recent decade, multiple trials confirmed the efficacy of ICI in various stages of the disease [117,118,124,257,258,259]. In most published studies, tumor response and survival subgroup analyses were mainly based on clinical characteristics, along with well-established biomarkers, such as TMB and PD-L1 status. PD-L1 positivity (>1%) is reported to correlate with better response to treatment, although up to 20% response rate was observed in PD-L1 negative (less than 1%) tumors [124,260,261]. Tumor mutations burden is considered to be generally high in cSCC, but nevertheless, this can vary significantly. In a recently published study with neo-adjuvant cemiplimab [124], patients’ median tumor mutational burden of 61.1 mut/Mb was found to correlate with the achievement of major or complete pathologic response [124]. From an immune microenvironment perspective, as expected, infiltration of T cells, including CD8+ T cells and CD4+ Th1 cells was associated with superior treatment response. This was accompanied by a significantly higher expression of IFNγ-related immune genes. On the other hand, abundant Tregs and CD68+ myeloid cells expressing the inhibitory checkpoint V-domain immunoglobulin suppressor of T cell activation (VISTA), were reported as biomarkers of resistance [261].
Merkel cell carcinoma (MCC) remains a very rare cancer type, usually affecting elderly patients, with 5-year overall survival ranging from 63% in stage I, to 13% in stage IV [125]. Two main factors are implicated in MCC tumorigenesis: Merkel cell polyomavirus (MCPyV) or chronic UV exposure. MCC is highly immunogenic due to the presence of either MCPyV-derived viral antigens or UV exposure-associated neoantigens, thus representing a disease amenable to ICI treatment [262,263,264,265]. Mutational profile varies between these two subgroups. For MCPyV-negative patients, next-generation sequencing (NGS) and whole exome sequencing (WES) have revealed a high tumor mutational burden (TMB), with recurrent mutations in TP53 and RB1. On the other hand, MCPyV-positive patients demonstrated 100-fold lower TMB than MCPyV-negative tumors, without a pattern of recurrent mutations [263,264]. Notwithstanding these differences, both subtypes remain highly immunogenic and responsive to ICI, independent of MCPyV or TMB status. As a result, to this point, TMB cannot be utilized as a predictive biomarker for response to ICI [114,266,267]. In the same vein, PD-L1 does not represent a robust biomarker either for this type of malignancy [266,267].
On the other hand, several studies have reported clinical biomarkers as surrogates of improved response to ICI. These include prior lines of treatment, where the use of ICI as a first-line therapy was associated with two-fold response rates as compared to administering it in the second-line setting [120,121,126,266]. Moreover, other clinical features associated with prolonged survival are less advanced disease at diagnosis, significant tumor shrinkage after exposure to ICI, limited number of tumor-affected organs, shorter disease-free interval between completion of initial treatment and recurrence and lack of immunosuppression [114,127,268]. Evaluation of peripheral blood indices showed longitudinal correlation of lower neutrophil to leukocyte ratio with improved patient outcomes. However, the correlation was not confirmed when the neutrophil/leucocyte ratio was assessed at baseline or individual timepoints only during treatment with ICI [127]. From immunological and molecular biomarker perspective, predominance of CD8+ T cells among tumor infiltrating lymphocytes (TIL) at baseline and presence of ARID2 and NTRK1 mutations were associated with a favorable response to PD-1/PD-L1 ICI therapy in advanced MCC [114,120,268].
Lastly, basal cell carcinoma (BCC) is the most common type of malignancy, although it rarely evolves into advanced disease, due to very latent behavior [255]. Along with cSCC, BCC has one of the highest mutational burdens among cancer types [269]. Due to the high efficacy of hedgehog inhibitors (HHI) in advanced BCC, immunotherapy was not evaluated in this disease only until recently [270]. After initial positive results of ICI following progression on HHI [122], multiple trials were opened to evaluate the role of ICI in the BCC treatment algorithm (Table 3). From a biomarker perspective, very limited data is currently available, although exploratory biomarker analysis from a recently reported trial [122] showed no clinically meaningful associations between objective response and any of the biomarkers, established as predictive in other tumor types, including PD-L1 expression, TMB or MHC-I expression.

6. Conclusions

The unparalleled survival improvement brought about by immune checkpoint treatment in patients with melanoma has undoubtedly changed the therapeutic landscape of the entire field of oncology, as well as the current focus of cancer research. One could draw similarities to the discovery of intracellular pathways and the advent of targeted treatments which was yet another important milestone for cancer clinical practice at the end of the previous century. In contrast to intracellular pathway blockade though, immune checkpoint inhibition offers longevity of disease response and the so-elusive prospect of a cure in the metastatic setting.
Notwithstanding the success of immunotherapy, there is still a significant number of patients with melanoma who have either primary or acquired resistance to immune checkpoint blockade. This particular cohort of patients is severely disadvantaged by the paucity of clinically validated biomarkers of immunotherapy resistance and in the worst-case scenario, they also have to endure the immune-related adverse events that come with treatment. There is an imperative need to discover the biological mechanisms underpinning this resistance and to determine high-sensitivity molecular markers to describe these mechanisms.
The focus now is turning to unraveling immune pathways that work synergistically with the known immune targets, with an aim to push the boundaries of treatment resistance even further. These discoveries will not be fruitful though unless accompanied by biomarkers of response and resistance that will guide target selection and treatment sequencing. The moment when “personalized immunotherapy” becomes a reality in clinical practice will hopefully come in the not-too-distant future.

Author Contributions

P.S., O.V., and S.G.: writing—original draft preparation. A.S.: writing—review and editing, supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

AS: Consultant for (Advisory Board, compensated): Merck, Bristol-Myers Squibb, Oncorus, Janssen. Grant/Research support from (Clinical Trials, paid to the Institution): Novartis, Bristol-Myers Squibb, Symphogen AstraZeneca/Medimmune, Merck, Bayer, Surface Oncology, Northern Biologics, Janssen Oncology/Johnson & Johnson, Roche, Regeneron, Alkermes, Array Biopharma/Pfizer, GSK, Oncorus, Treadwell, Amgen. PS/OV/SG have no conflict of interest to declare.

References

  1. Korn, E.L.; Liu, P.Y.; Lee, S.J.; Chapman, J.A.W.; Niedzwiecki, D.; Suman, V.J.; Moon, J.; Sondak, V.K.; Atkins, M.B.; Eisenhauer, E.A.; et al. Meta-analysis of phase II cooperative group trials in metastatic stage IV melanoma to determine progression-free and overall survival benchmarks for future phase II trials. J. Clin. Oncol. 2008, 26, 527–534. [Google Scholar] [CrossRef] [PubMed]
  2. Ugurel, S.; Röhmel, J.; Ascierto, P.A.; Flaherty, K.T.; Grob, J.J.; Hauschild, A.; Larkin, J.; Long, G.V.; Lorigan, P.; McArthur, G.A.; et al. Survival of patients with advanced metastatic melanoma: The impact of novel therapies-update 2017. Eur. J. Cancer 2017, 83, 247–257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Bhatia, A.; Kumar, Y. Cellular and molecular mechanisms in cancer immune escape: A comprehensive review. Expert Rev. Clin. Immunol. 2014, 10, 41–62. [Google Scholar] [CrossRef] [PubMed]
  4. Teft, W.A.; Kirchhof, M.G.; Madrenas, J. A molecular perspective of CTLA-4 function. Annu. Rev. Immunol. 2006, 24, 65–97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Tivol, E.A.; Borriello, F.; Schweitzer, A.N.; Lynch, W.P.; Bluestone, J.A.; Sharpe, A.H. Loss of CTLA-4 leads to massive lymphoproliferation and fatal multiorgan tissue destruction, revealing a critical negative regulatory role of CTLA-4. Immunity 1995, 3, 541–547. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Chen, L. Co-inhibitory molecules of the B7-CD28 family in the control of T-cell immunity. Nat. Rev. Immunol. 2004, 4, 336–347. [Google Scholar] [CrossRef]
  7. Greenwald, R.J.; Freeman, G.J.; Sharpe, A.H. The B7 family revisited. Annu. Rev. Immunol. 2005, 23, 515–548. [Google Scholar] [CrossRef] [PubMed]
  8. Krummel, M.F.; Allison, J.P. CD28 and CTLA-4 have opposing effects on the response of T cells to stimulation. J. Exp. Med. 1995, 182, 459–465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Pardoll, D.M. The blockade of immune checkpoints in cancer immunotherapy. Nat. Rev. Cancer 2012, 12, 252–264. [Google Scholar] [CrossRef] [Green Version]
  10. Wang, C.; Thudium, K.B.; Han, M.; Wang, X.T.; Huang, H.; Feingersh, D.; Garcia, C.; Wu, Y.; Kuhne, M.; Srinivasan, M.; et al. In vitro characterization of the anti-PD-1 antibody nivolumab, BMS-936558, and in vivo toxicology in non-human primates. Cancer Immunol. Res. 2014, 2, 846–856. [Google Scholar] [CrossRef]
  11. Abbosh, C.; Birkbak, N.J.; Wilson, G.A.; Jamal-Hanjani, M.; Constantin, T.; Salari, R.; Le Quesne, J.; Moore, D.A.; Veeriah, S.; Rosenthal, R.; et al. Phylogenetic ctDNA analysis depicts early-stage lung cancer evolution. Nature 2017, 545, 446–451. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Wolchok, J.D.; Chiarion-Sileni, V.; Gonzalez, R.; Grob, J.J.; Rutkowski, P.; Lao, C.D.; Cowey, C.L.; Schadendorf, D.; Wagstaff, J.; Dummer, R.; et al. Long-Term Outcomes with Nivolumab Plus Ipilimumab or Nivolumab Alone Versus Ipilimumab in Patients with Advanced Melanoma. J. Clin. Oncol. 2022, 40, 127–137. [Google Scholar] [CrossRef] [PubMed]
  13. Davies, M.A. Is It Safe to Stop Anti–PD-1 Immunotherapy in Patients with Metastatic Melanoma Who Achieve a Complete Response? J. Clin. Oncol. 2020, 38, 1645–1647. [Google Scholar] [CrossRef]
  14. Robert, C.; Ribas, A.; Hamid, O.; Daud, A.; Wolchok, J.D.; Joshua, A.M.; Hwu, W.J.; Weber, J.S.; Gangadhar, T.C.; Joseph, R.W.; et al. Durable Complete Response After Discontinuation of Pembrolizumab in Patients with Metastatic Melanoma. J. Clin. Oncol. 2018, 36, 1668–1674. [Google Scholar] [CrossRef]
  15. Jansen, Y.J.L.; Rozeman, E.A.; Mason, R.; Goldinger, S.M.; Geukes Foppen, M.H.; Hoejberg, L.; Schmidt, H.; van Thienen, J.V.; Haanen, J.; Tiainen, L.; et al. Discontinuation of anti-PD-1 antibody therapy in the absence of disease progression or treatment limiting toxicity: Clinical outcomes in advanced melanoma. Ann. Oncol. 2019, 30, 1154–1161. [Google Scholar] [CrossRef] [Green Version]
  16. Betof Warner, A.; Palmer, J.S.; Shoushtari, A.N.; Goldman, D.A.; Panageas, K.S.; Hayes, S.A.; Bajwa, R.; Momtaz, P.; Callahan, M.K.; Wolchok, J.D.; et al. Long-Term Outcomes and Responses to Retreatment in Patients with Melanoma Treated with PD-1 Blockade. J. Clin. Oncol. 2020, 38, 1655–1663. [Google Scholar] [CrossRef]
  17. ClinicalTrials.gov. Duration of Anti-PD-1 Therapy in Metastatic Melanoma (STOP-GAP) 2022. Available online: https://clinicaltrials.gov/ct2/show/NCT02821013 (accessed on 13 November 2022).
  18. Larkin, J.; Chiarion-Sileni, V.; Gonzalez, R.; Grob, J.J.; Cowey, C.L.; Lao, C.D.; Schadendorf, D.; Dummer, R.; Smylie, M.; Rutkowski, P.; et al. Combined Nivolumab and Ipilimumab or Monotherapy in Untreated Melanoma. N Engl. J. Med. 2015, 373, 23–34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Tawbi, H.A.; Forsyth, P.A.; Hodi, F.S.; Algazi, A.P.; Hamid, O.; Lao, C.D.; Moschos, S.J.; Atkins, M.B.; Lewis, K.; Postow, M.A.; et al. Long-term outcomes of patients with active melanoma brain metastases treated with combination nivolumab plus ipilimumab (CheckMate 204): Final results of an open-label, multicentre, phase 2 study. Lancet Oncol. 2021, 22, 1692–1704. [Google Scholar] [CrossRef]
  20. Long, G.V.; Atkinson, V.; Lo, S.; Guminski, A.D.; Sandhu, S.K.; Brown, M.P.; Gonzalez, M.; Scolyer, R.A.; Emmett, L.; McArthur, G.A.; et al. Five-year overall survival from the anti-PD1 brain collaboration (ABC Study): Randomized phase 2 study of nivolumab (nivo) or nivo+ipilimumab (ipi) in patients (pts) with melanoma brain metastases (mets). J. Clin. Oncol. 2021, 39, 9508. [Google Scholar] [CrossRef]
  21. Di Giacomo, A.M.; Chiarion-Sileni, V.; Del Vecchio, M.; Ferrucci, P.F.; Guida, M.; Quaglino, P.; Guidoboni, M.; Marchetti, P.; Cutaia, O.; Amato, G.; et al. Primary Analysis and 4-Year Follow-Up of the Phase III NIBIT-M2 Trial in Melanoma Patients with Brain Metastases. Clin. Cancer Res. 2021, 27, 4737–4745. [Google Scholar] [CrossRef] [PubMed]
  22. Margolin, K.; Ernstoff, M.S.; Hamid, O.; Lawrence, D.; McDermott, D.; Puzanov, I.; Wolchok, J.D.; Clark, J.I.; Sznol, M.; Logan, T.F.; et al. Ipilimumab in patients with melanoma and brain metastases: An open-label, phase 2 trial. Lancet Oncol. 2012, 13, 459–465. [Google Scholar] [CrossRef] [PubMed]
  23. Pedersen, S.; Møller, S.; Donia, M.; Persson, G.F.; Svane, I.M.; Ellebaek, E. Real-world data on melanoma brain metastases and survival outcome. Melanoma Res. 2022, 32, 173–182. [Google Scholar] [CrossRef] [PubMed]
  24. Carron, R.; Gaudy-Marqueste, C.; Amatore, F.; Padovani, L.; Malissen, N.; Balossier, A.; Loundou, A.; Bonnet, N.; Muracciole, X.; Régis, J.-M.; et al. Stereotactic radiosurgery combined with anti-PD1 for the management of melanoma brain metastases: A retrospective study of safety and efficacy. Eur. J. Cancer 2020, 135, 52–61. [Google Scholar] [CrossRef] [PubMed]
  25. Gatterbauer, B.; Hirschmann, D.; Eberherr, N.; Untersteiner, H.; Cho, A.; Shaltout, A.; Göbl, P.; Fitschek, F.; Dorfer, C.; Wolfsberger, S.; et al. Toxicity and efficacy of Gamma Knife radiosurgery for brain metastases in melanoma patients treated with immunotherapy or targeted therapy—A retrospective cohort study. Cancer Med. 2020, 9, 4026–4036. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Hodi, F.S.; Chiarion -Sileni, V.; Lewis, K.D.; Grob, J.-J.; Rutkowski, P.; Lao, C.D.; Cowey, C.L.; Schadendorf, D.; Wagstaff, J.; Dummer, R.; et al. Long-term survival in advanced melanoma for patients treated with nivolumab plus ipilimumab in CheckMate 067. J. Clin. Oncol. 2022, 40, 9522. [Google Scholar] [CrossRef]
  27. Lebbé, C.; Meyer, N.; Mortier, L.; Marquez-Rodas, I.; Robert, C.; Rutkowski, P.; Menzies, A.M.; Eigentler, T.; Ascierto, P.A.; Smylie, M.; et al. Evaluation of Two Dosing Regimens for Nivolumab in Combination with Ipilimumab in Patients with Advanced Melanoma: Results from the Phase IIIb/IV CheckMate 511 Trial. J. Clin. Oncol. 2019, 37, 867–875. [Google Scholar] [CrossRef]
  28. Lebbe, C.; Meyer, N.; Mortier, L.; Marquez-Rodas, I.; Robert, C.; Rutkowski, P.; Butler, M.O.; Eigentler, T.; Menzies, A.M.; Smylie, M.; et al. Two dosing regimens of nivolumab (NIVO) plus ipilimumab (IPI) for advanced (adv) melanoma: Three-year results of CheckMate 511. J. Clin. Oncol. 2021, 39, 9516. [Google Scholar] [CrossRef]
  29. Ascierto, P.A.; Ferrucci, P.F.; Fisher, R.; Del Vecchio, M.; Atkinson, V.; Schmidt, H.; Schachter, J.; Queirolo, P.; Long, G.V.; Di Giacomo, A.M.; et al. Dabrafenib, trametinib and pembrolizumab or placebo in BRAF-mutant melanoma. Nat. Med. 2019, 25, 941–946. [Google Scholar] [CrossRef]
  30. Dummer, R.; Long, G.V.; Robert, C.; Tawbi, H.A.; Flaherty, K.T.; Ascierto, P.A.; Nathan, P.D.; Rutkowski, P.; Leonov, O.; Dutriaux, C.; et al. Randomized Phase III Trial Evaluating Spartalizumab Plus Dabrafenib and Trametinib for BRAF V600–Mutant Unresectable or Metastatic Melanoma. J. Clin. Oncol. 2022, 40, 1428–1438. [Google Scholar] [CrossRef]
  31. Gutzmer, R.; Stroyakovskiy, D.; Gogas, H.; Robert, C.; Lewis, K.; Protsenko, S.; Pereira, R.P.; Eigentler, T.; Rutkowski, P.; Demidov, L.; et al. Atezolizumab, vemurafenib, and cobimetinib as first-line treatment for unresectable advanced BRAF V600 mutation-positive melanoma (IMspire150): Primary analysis of the randomised, double-blind, placebo-controlled, phase 3 trial. Lancet 2020, 395, 1835–1844. [Google Scholar] [CrossRef]
  32. Ferrucci, P.F.; Di Giacomo, A.M.; Del Vecchio, M.; Atkinson, V.; Schmidt, H.; Schachter, J.; Queirolo, P.; Long, G.V.; Stephens, R.; Svane, I.M.; et al. KEYNOTE-022 part 3: A randomized, double-blind, phase 2 study of pembrolizumab, dabrafenib, and trametinib in BRAF mutant melanoma. J. Immunother.Immunother. Cancer 2020, 8, e001806. [Google Scholar] [CrossRef] [PubMed]
  33. Long, G.V.; Menzies, A.M.; Nagrial, A.M.; Haydu, L.E.; Hamilton, A.L.; Mann, G.J.; Hughes, T.M.; Thompson, J.F.; Scolyer, R.A.; Kefford, R.F. Prognostic and Clinicopathologic Associations of Oncogenic BRAF in Metastatic Melanoma. J. Clin. Oncol. 2011, 29, 1239–1246. [Google Scholar] [CrossRef] [PubMed]
  34. Davies, H.; Bignell, G.R.; Cox, C.; Stephens, P.; Edkins, S.; Clegg, S.; Teague, J.; Woffendin, H.; Garnett, M.J.; Bottomley, W.; et al. Mutations of the BRAF gene in human cancer. Nature 2002, 417, 949–954. [Google Scholar] [CrossRef] [Green Version]
  35. Cheng, L.; Lopez-Beltran, A.; Massari, F.; MacLennan, G.T.; Montironi, R. Molecular testing for BRAF mutations to inform melanoma treatment decisions: A move toward precision medicine. Mod. Pathol. 2018, 31, 24–38. [Google Scholar] [CrossRef] [Green Version]
  36. Long, G.V.; Stroyakovskiy, D.; Gogas, H.; Levchenko, E.; de Braud, F.; Larkin, J.; Garbe, C.; Jouary, T.; Hauschild, A.; Grob, J.J.; et al. Combined BRAF and MEK Inhibition versus BRAF Inhibition Alone in Melanoma. N. Engl. J. Med. 2014, 371, 1877–1888. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Chapman, P.B.; Hauschild, A.; Robert, C.; Haanen, J.B.; Ascierto, P.; Larkin, J.; Dummer, R.; Garbe, C.; Testori, A.; Maio, M.; et al. Improved Survival with Vemurafenib in Melanoma with BRAF V600E Mutation. N. Engl. J. Med. 2011, 364, 2507–2516. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Atkins, M.B.; Lee, S.J.; Chmielowski, B.; Tarhini, A.A.; Cohen, G.I.; Truong, T.-G.; Moon, H.H.; Davar, D.; O’Rourke, M.; Stephenson, J.J.; et al. Combination Dabrafenib and Trametinib Versus Combination Nivolumab and Ipilimumab for Patients with Advanced BRAF-Mutant Melanoma: The DREAMseq Trial—ECOG-ACRIN EA6134. J. Clin. Oncol. 2022, 41, JCO-22. [Google Scholar] [CrossRef]
  39. Ascierto, P.A.; Mandalà, M.; Ferrucci, P.F.; Guidoboni, M.; Rutkowski, P.; Ferraresi, V.; Arance, A.; Guida, M.; Maiello, E.; Gogas, H.; et al. Sequencing of Ipilimumab Plus Nivolumab and Encorafenib Plus Binimetinib for Untreated BRAF-Mutated Metastatic Melanoma (SECOMBIT): A Randomized, Three-Arm, Open-Label Phase II Trial. J. Clin. Oncol. 2022, 41, Jco2102961. [Google Scholar] [CrossRef]
  40. Schadendorf, D.; Gogas, H.; Kandolf Sekulovic, L.; Meier, F.E.; Eigentler, T.; Simon, J.-C.; Terheyden, P.A.M.; Gesierich, A.H.; Herbst, R.A.; Kähler, K.C.; et al. Efficacy and safety of sequencing with vemurafenib (V) plus cobimetinib (C) followed by atezolizumab (Atezo) in patients (pts) with advanced BRAFV600-positive melanoma: Interim analysis of the ImmunoCobiVem study. J. Clin. Oncol. 2022, 40, 9548. [Google Scholar] [CrossRef]
  41. Rosenberg, S.A.; Yang, J.C.; Topalian, S.L.; Schwartzentruber, D.J.; Weber, J.S.; Parkinson, D.R.; Seipp, C.A.; Einhorn, J.H.; White, D.E. Treatment of 283 consecutive patients with metastatic melanoma or renal cell cancer using high-dose bolus interleukin 2. Jama 1994, 271, 907–913. [Google Scholar] [CrossRef]
  42. Alva, A.; Daniels, G.A.; Wong, M.K.; Kaufman, H.L.; Morse, M.A.; McDermott, D.F.; Clark, J.I.; Agarwala, S.S.; Miletello, G.; Logan, T.F.; et al. Contemporary experience with high-dose interleukin-2 therapy and impact on survival in patients with metastatic melanoma and metastatic renal cell carcinoma. Cancer Immunol. Immunother. 2016, 65, 1533–1544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Davar, D.; Ding, F.; Saul, M.; Sander, C.; Tarhini, A.A.; Kirkwood, J.M.; Tawbi, H.A. High-dose interleukin-2 (HD IL-2) for advanced melanoma: A single center experience from the University of Pittsburgh Cancer Institute. J. Immunother. Cancer 2017, 5, 74. [Google Scholar] [CrossRef] [PubMed]
  44. Atkins, M.B.; Kunkel, L.; Sznol, M.; Rosenberg, S.A. High-dose recombinant interleukin-2 therapy in patients with metastatic melanoma: Long-term survival update. Cancer J. Sci. Am. 2000, 6 (Suppl. S1), S11–S14. [Google Scholar] [PubMed]
  45. Buchbinder, E.I.; Gunturi, A.; Perritt, J.; Dutcher, J.; Aung, S.; Kaufman, H.L.; Ernstoff, M.S.; Miletello, G.P.; Curti, B.D.; Daniels, G.A.; et al. A retrospective analysis of High-Dose Interleukin-2 (HD IL-2) following Ipilimumab in metastatic melanoma. J. Immunother. Cancer 2016, 4, 52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Eggermont, A.M.; Chiarion-Sileni, V.; Grob, J.J.; Dummer, R.; Wolchok, J.D.; Schmidt, H.; Hamid, O.; Robert, C.; Ascierto, P.A.; Richards, J.M.; et al. Adjuvant ipilimumab versus placebo after complete resection of high-risk stage III melanoma (EORTC 18071): A randomised, double-blind, phase 3 trial. Lancet Oncol. 2015, 16, 522–530. [Google Scholar] [CrossRef]
  47. Weber, J.; Mandala, M.; Del Vecchio, M.; Gogas, H.J.; Arance, A.M.; Cowey, C.L.; Dalle, S.; Schenker, M.; Chiarion-Sileni, V.; Marquez-Rodas, I.; et al. Adjuvant Nivolumab versus Ipilimumab in Resected Stage III or IV Melanoma. N. Engl. J. Med. 2017, 377, 1824–1835. [Google Scholar] [CrossRef]
  48. Eggermont, A.M.M.; Blank, C.U.; Mandalà, M.; Long, G.V.; Atkinson, V.G.; Dalle, S.; Haydon, A.M.; Meshcheryakov, A.; Khattak, A.; Carlino, M.S.; et al. Adjuvant pembrolizumab versus placebo in resected stage III melanoma (EORTC 1325-MG/KEYNOTE-054): Distant metastasis-free survival results from a double-blind, randomised, controlled, phase 3 trial. Lancet Oncol. 2021, 22, 643–654. [Google Scholar] [CrossRef]
  49. Luke, J.J.; Rutkowski, P.; Queirolo, P.; Del Vecchio, M.; Mackiewicz, J.; Chiarion-Sileni, V.; de la Cruz Merino, L.; Khattak, M.A.; Schadendorf, D.; Long, G.V.; et al. Pembrolizumab versus placebo as adjuvant therapy in completely resected stage IIB or IIC melanoma (KEYNOTE-716): A randomised, double-blind, phase 3 trial. Lancet 2022, 399, 1718–1729. [Google Scholar] [CrossRef]
  50. Blank, C.U.; Rozeman, E.A.; Fanchi, L.F.; Sikorska, K.; van de Wiel, B.; Kvistborg, P.; Krijgsman, O.; van den Braber, M.; Philips, D.; Broeks, A.; et al. Neoadjuvant versus adjuvant ipilimumab plus nivolumab in macroscopic stage III melanoma. Nat. Med. 2018, 24, 1655–1661. [Google Scholar] [CrossRef]
  51. Blank, C.U.; Reijers, I.L.M.; Saw, R.P.M.; Versluis, J.M.; Pennington, T.; Kapiteijn, E.; Van Der Veldt, A.A.M.; Suijkerbuijk, K.; Hospers, G.; van Houdt, W.J.; et al. Survival data of PRADO: A phase 2 study of personalized response-driven surgery and adjuvant therapy after neoadjuvant ipilimumab (IPI) and nivolumab (NIVO) in resectable stage III melanoma. J. Clin. Oncol. 2022, 40, 9501. [Google Scholar] [CrossRef]
  52. Rozeman, E.A.; Hoefsmit, E.P.; Reijers, I.L.M.; Saw, R.P.M.; Versluis, J.M.; Krijgsman, O.; Dimitriadis, P.; Sikorska, K.; van de Wiel, B.A.; Eriksson, H.; et al. Survival and biomarker analyses from the OpACIN-neo and OpACIN neoadjuvant immunotherapy trials in stage III melanoma. Nat. Med. 2021, 27, 256–263. [Google Scholar] [CrossRef] [PubMed]
  53. Patel, S.; Othus, M.; Prieto, V.; Lowe, M.; Buchbinder, E.; Chen, Y.; Hyngstrom, J.; Lao, C.D.; Truong, T.; Chandra, S.; et al. LBA6—Neoadjvuant versus adjuvant pembrolizumab for resected stage III-IV melanoma (SWOG S1801). Ann. Oncol. 2022, 33 (Suppl. S7), S808–S869. [Google Scholar] [CrossRef]
  54. Andtbacka, R.H.I.; Collichio, F.; Harrington, K.J.; Middleton, M.R.; Downey, G.; Öhrling, K.; Kaufman, H.L. Final analyses of OPTiM: A randomized phase III trial of talimogene laherparepvec versus granulocyte-macrophage colony-stimulating factor in unresectable stage III–IV melanoma. J. Immunother. Cancer 2019, 7, 145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Ribas, A.; Chesney, J.; Long, G.V.; Kirkwood, J.M.; Dummer, R.; Puzanov, I.; Hoeller, C.; Gajewski, T.F.; Gutzmer, R.; Rutkowski, P.; et al. 1037O MASTERKEY-265: A phase III, randomized, placebo (Pbo)-controlled study of talimogene laherparepvec (T) plus pembrolizumab (P) for unresectable stage IIIB–IVM1c melanoma (MEL). Ann. Oncol. 2021, 32, S868–S869. [Google Scholar] [CrossRef]
  56. Rohaan, M.W.; Stahlie, E.H.A.; Franke, V.; Zijlker, L.P.; Wilgenhof, S.; van der Noort, V.; van Akkooi, A.C.J.; Haanen, J.B.A.G. Neoadjuvant nivolumab + T-VEC combination therapy for resectable early stage or metastatic (IIIB-IVM1a) melanoma with injectable disease: Study protocol of the NIVEC trial. BMC Cancer 2022, 22, 851. [Google Scholar] [CrossRef] [PubMed]
  57. Triebel, F.; Jitsukawa, S.; Baixeras, E.; Roman-Roman, S.; Genevee, C.; Viegas-Pequignot, E.; Hercend, T. LAG-3, a novel lymphocyte activation gene closely related to CD4. J. Exp. Med. 1990, 171, 1393–1405. [Google Scholar] [CrossRef] [Green Version]
  58. Workman, C.J.; Cauley, L.S.; Kim, I.-J.; Blackman, M.A.; Woodland, D.L.; Vignali, D.A.A. Lymphocyte Activation Gene-3 (CD223) Regulates the Size of the Expanding T Cell Population Following Antigen Activation In Vivo. J. Immunol. 2004, 172, 5450. [Google Scholar] [CrossRef] [Green Version]
  59. Matsuzaki, J.; Gnjatic, S.; Mhawech-Fauceglia, P.; Beck, A.; Miller, A.; Tsuji, T.; Eppolito, C.; Qian, F.; Lele, S.; Shrikant, P.; et al. Tumor-infiltrating NY-ESO-1-specific CD8+ T cells are negatively regulated by LAG-3 and PD-1 in human ovarian cancer. Natl. Acad. Sci. USA 2010, 107, 7875–7880. [Google Scholar] [CrossRef] [Green Version]
  60. Demeure, C.E.; Wolfers, J.; Martin-Garcia, N.; Gaulard, P.; Triebel, F. T Lymphocytes infiltrating various tumour types express the MHC class II ligand lymphocyte activation gene-3 (LAG-3): Role of LAG-3/MHC class II interactions in cell-cell contacts. Eur. J. Cancer 2001, 37, 1709–1718. [Google Scholar] [CrossRef]
  61. Baitsch, L.; Baumgaertner, P.; Devêvre, E.; Raghav, S.K.; Legat, A.; Barba, L.; Wieckowski, S.; Bouzourene, H.; Deplancke, B.; Romero, P.; et al. Exhaustion of tumor-specific CD8+ T cells in metastases from melanoma patients. J. Clin. Investig. 2011, 121, 2350–2360. [Google Scholar] [CrossRef]
  62. Tawbi, H.A.; Schadendorf, D.; Lipson, E.J.; Ascierto, P.A.; Matamala, L.; Castillo Gutierrez, E.; Rutkowski, P.; Gogas, H.J.; Lao, C.D.; De Menezes, J.J.; et al. Relatlimab and Nivolumab versus Nivolumab in Untreated Advanced Melanoma. N Engl. J. Med. 2022, 386, 24–34. [Google Scholar] [CrossRef] [PubMed]
  63. Chauvin, J.-M.; Zarour, H.M. TIGIT in cancer immunotherapy. J. Immunother. Cancer 2020, 8, e000957. [Google Scholar] [CrossRef] [PubMed]
  64. Chauvin, J.-M.; Pagliano, O.; Fourcade, J.; Sun, Z.; Wang, H.; Sander, C.; Kirkwood, J.M.; Chen, T.-h.T.; Maurer, M.; Korman, A.J.; et al. TIGIT and PD-1 impair tumor antigen–specific CD8+ T cells in melanoma patients. J. Clin. Investig. 2015, 125, 2046–2058. [Google Scholar] [CrossRef] [PubMed]
  65. Liu, B.; Li, Z.; Mahesh, S.P.; Pantanelli, S.; Hwang, F.S.; Siu, W.O.; Nussenblatt, R.B. Glucocorticoid-induced Tumor Necrosis Factor Receptor Negatively Regulates Activation of Human Primary Natural Killer (NK) Cells by Blocking Proliferative Signals and Increasing NK Cell Apoptosis. J. Biol. Chem. 2008, 283, 8202–8210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Schoenhals, J.E.; Cushman, T.R.; Barsoumian, H.B.; Li, A.; Cadena, A.P.; Niknam, S.; Younes, A.I.; Caetano, M.d.S.; Cortez, M.A.; Welsh, J.W. Anti-glucocorticoid-induced Tumor Necrosis Factor–Related Protein (GITR) Therapy Overcomes Radiation-Induced Treg Immunosuppression and Drives Abscopal Effects. Front. Immunol. 2018, 9, 2170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Schaer, D.A.; Murphy, J.T.; Wolchok, J.D. Modulation of GITR for cancer immunotherapy. Curr. Opin. Immunol. 2012, 24, 217–224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Clouthier, D.L.; Watts, T.H. Cell-specific and context-dependent effects of GITR in cancer, autoimmunity, and infection. Cytokine Growth Factor Rev. 2014, 25, 91–106. [Google Scholar] [CrossRef] [PubMed]
  69. Shimizu, J.; Yamazaki, S.; Takahashi, T.; Ishida, Y.; Sakaguchi, S. Stimulation of CD25+CD4+ regulatory T cells through GITR breaks immunological self-tolerance. Nat. Immunol. 2002, 3, 135–142. [Google Scholar] [CrossRef]
  70. Narumi, K.; Miyakawa, R.; Shibasaki, C.; Henmi, M.; Mizoguchi, Y.; Ueda, R.; Hashimoto, H.; Hiraoka, N.; Yoshida, T.; Aoki, K. Local Administration of GITR Agonistic Antibody Induces a Stronger Antitumor Immunity than Systemic Delivery. Sci. Rep. 2019, 9, 5562. [Google Scholar] [CrossRef] [Green Version]
  71. Papadopoulos, K.P.; Autio, K.; Golan, T.; Dobrenkov, K.; Chartash, E.; Chen, Q.; Wnek, R.; Long, G.V. Phase I Study of MK-4166, an Anti-human Glucocorticoid-Induced TNF Receptor Antibody, Alone or with Pembrolizumab in Advanced Solid Tumors. Clin. Cancer Res. 2021, 27, 1904–1911. [Google Scholar] [CrossRef]
  72. Arance, A.; de la Cruz-Merino, L.; Petrella, T.M.; Jamal, R.; Ny, L.; Carneiro, A.; Berrocal, A.; Márquez-Rodas, I.; Spreafico, A.; Atkinson, V.; et al. Phase II LEAP-004 Study of Lenvatinib Plus Pembrolizumab for Melanoma with Confirmed Progression on a Programmed Cell Death Protein-1 or Programmed Death Ligand 1 Inhibitor Given as Monotherapy or in Combination. J. Clin. Oncol. 2022, 41, 75–85. [Google Scholar] [CrossRef]
  73. Tohyama, O.; Matsui, J.; Kodama, K.; Hata-Sugi, N.; Kimura, T.; Okamoto, K.; Minoshima, Y.; Iwata, M.; Funahashi, Y. Antitumor Activity of Lenvatinib (E7080): An Angiogenesis Inhibitor That Targets Multiple Receptor Tyrosine Kinases in Preclinical Human Thyroid Cancer Models. J. Thyroid Res. 2014, 2014, 638747. [Google Scholar] [CrossRef]
  74. Yamamoto, Y.; Matsui, J.; Matsushima, T.; Obaishi, H.; Miyazaki, K.; Nakamura, K.; Tohyama, O.; Semba, T.; Yamaguchi, A.; Hoshi, S.S.; et al. Lenvatinib, an angiogenesis inhibitor targeting VEGFR/FGFR, shows broad antitumor activity in human tumor xenograft models associated with microvessel density and pericyte coverage. Vasc. Cell 2014, 6, 18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Adachi, Y.; Kamiyama, H.; Ichikawa, K.; Fukushima, S.; Ozawa, Y.; Yamaguchi, S.; Goda, S.; Kimura, T.; Kodama, K.; Matsuki, M.; et al. Inhibition of FGFR Reactivates IFNγ Signaling in Tumor Cells to Enhance the Combined Antitumor Activity of Lenvatinib with Anti-PD-1 Antibodies. Cancer Res. 2022, 82, 292–306. [Google Scholar] [CrossRef] [PubMed]
  76. Yi, C.; Chen, L.; Lin, Z.; Liu, L.; Shao, W.; Zhang, R.; Lin, J.; Zhang, J.; Zhu, W.; Jia, H.; et al. Lenvatinib Targets FGF Receptor 4 to Enhance Antitumor Immune Response of Anti-Programmed Cell Death-1 in HCC. Hepatology 2021, 74, 2544–2560. [Google Scholar] [CrossRef] [PubMed]
  77. Sarnaik, A.A.; Hamid, O.; Khushalani, N.I.; Lewis, K.D.; Medina, T.; Kluger, H.M.; Thomas, S.S.; Domingo-Musibay, E.; Pavlick, A.C.; Whitman, E.D.; et al. Lifileucel, a Tumor-Infiltrating Lymphocyte Therapy, in Metastatic Melanoma. J. Clin. Oncol. 2021, 39, 2656–2666. [Google Scholar] [CrossRef] [PubMed]
  78. Kaliki, S.; Shields, C.L. Uveal melanoma: Relatively rare but deadly cancer. Eye 2017, 31, 241–257. [Google Scholar] [CrossRef] [Green Version]
  79. Chattopadhyay, C.; Kim, D.W.; Gombos, D.S.; Oba, J.; Qin, Y.; Williams, M.D.; Esmaeli, B.; Grimm, E.A.; Wargo, J.A.; Woodman, S.E.; et al. Uveal melanoma: From diagnosis to treatment and the science in between. Cancer 2016, 122, 2299–2312. [Google Scholar] [CrossRef] [Green Version]
  80. Mahendraraj, K.; Lau, C.S.; Lee, I.; Chamberlain, R.S. Trends in incidence, survival, and management of uveal melanoma: A population-based study of 7,516 patients from the Surveillance, Epidemiology, and End Results database (1973–2012). Clin. Ophthalmol. 2016, 10, 2113–2119. [Google Scholar] [CrossRef] [Green Version]
  81. Shields, C.L.; Kaliki, S.; Furuta, M.; Fulco, E.; Alarcon, C.; Shields, J.A. American Joint Committee on Cancer classification of posterior uveal melanoma (tumor size category) predicts prognosis in 7731 patients. Ophthalmology 2013, 120, 2066–2071. [Google Scholar] [CrossRef]
  82. Shields, C.L.; Kaliki, S.; Furuta, M.; Fulco, E.; Alarcon, C.; Shields, J.A. American Joint Committee on Cancer Classification of Uveal Melanoma (Anatomic Stage) Predicts Prognosis in 7731 Patients: The 2013 Zimmerman Lecture. Ophthalmology 2015, 122, 1180–1186. [Google Scholar] [CrossRef] [PubMed]
  83. Rowcroft, A.; Loveday, B.P.T.; Thomson, B.N.J.; Banting, S.; Knowles, B. Systematic review of liver directed therapy for uveal melanoma hepatic metastases. HPB 2020, 22, 497–505. [Google Scholar] [CrossRef] [PubMed]
  84. Pelster, M.S.; Gruschkus, S.K.; Bassett, R.; Gombos, D.S.; Shephard, M.; Posada, L.; Glover, M.S.; Simien, R.; Diab, A.; Hwu, P.; et al. Nivolumab and Ipilimumab in Metastatic Uveal Melanoma: Results from a Single-Arm Phase II Study. J. Clin. Oncol. 2021, 39, 599–607. [Google Scholar] [CrossRef]
  85. Piulats, J.M.; Espinosa, E.; de la Cruz Merino, L.; Varela, M.; Alonso Carrion, L.; Martin-Algarra, S.; Lopez Castro, R.; Curiel, T.; Rodriguez-Abreu, D.; Redrado, M.; et al. Nivolumab Plus Ipilimumab for Treatment-Naive Metastatic Uveal Melanoma: An Open-Label, Multicenter, Phase II Trial by the Spanish Multidisciplinary Melanoma Group (GEM-1402). J. Clin. Oncol. 2021, 39, 586–598. [Google Scholar] [CrossRef] [PubMed]
  86. Najjar, Y.G.; Navrazhina, K.; Ding, F.; Bhatia, R.; Tsai, K.; Abbate, K.; Durden, B.; Eroglu, Z.; Bhatia, S.; Park, S.; et al. Ipilimumab plus nivolumab for patients with metastatic uveal melanoma: A multicenter, retrospective study. J. Immunother. Cancer 2020, 8, e000331. [Google Scholar] [CrossRef] [PubMed]
  87. Heppt, M.V.; Amaral, T.; Kahler, K.C.; Heinzerling, L.; Hassel, J.C.; Meissner, M.; Kreuzberg, N.; Loquai, C.; Reinhardt, L.; Utikal, J.; et al. Combined immune checkpoint blockade for metastatic uveal melanoma: A retrospective, multi-center study. J. Immunother. Cancer 2019, 7, 299. [Google Scholar] [CrossRef] [PubMed]
  88. Heppt, M.V.; Heinzerling, L.; Kahler, K.C.; Forschner, A.; Kirchberger, M.C.; Loquai, C.; Meissner, M.; Meier, F.; Terheyden, P.; Schell, B.; et al. Prognostic factors and outcomes in metastatic uveal melanoma treated with programmed cell death-1 or combined PD-1/cytotoxic T-lymphocyte antigen-4 inhibition. Eur. J. Cancer 2017, 82, 56–65. [Google Scholar] [CrossRef] [PubMed]
  89. Namikawa, K.; Takahashi, A.; Mori, T.; Tsutsumida, A.; Suzuki, S.; Motoi, N.; Jinnai, S.; Kage, Y.; Mizuta, H.; Muto, Y.; et al. Nivolumab for patients with metastatic uveal melanoma previously untreated with ipilimumab: A single-institution retrospective study. Melanoma Res. 2020, 30, 76–84. [Google Scholar] [CrossRef]
  90. Klemen, N.D.; Wang, M.; Rubinstein, J.C.; Olino, K.; Clune, J.; Ariyan, S.; Cha, C.; Weiss, S.A.; Kluger, H.M.; Sznol, M. Survival after checkpoint inhibitors for metastatic acral, mucosal and uveal melanoma. J. Immunother. Cancer 2020, 8, e000341b. [Google Scholar] [CrossRef]
  91. Taylor, A.W.; Hsu, S.; Ng, T.F. The Role of Retinal Pigment Epithelial Cells in Regulation of Macrophages/Microglial Cells in Retinal Immunobiology. Front. Immunol. 2021, 12, 724601. [Google Scholar] [CrossRef]
  92. Johansson, P.; Aoude, L.G.; Wadt, K.; Glasson, W.J.; Warrier, S.K.; Hewitt, A.W.; Kiilgaard, J.F.; Heegaard, S.; Isaacs, T.; Franchina, M.; et al. Deep sequencing of uveal melanoma identifies a recurrent mutation in PLCB4. Oncotarget 2016, 7, 4624–4631. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Vergara, I.A.; Wilmott, J.S.; Long, G.V.; Scolyer, R.A. Genetic drivers of non-cutaneous melanomas: Challenges and opportunities in a heterogeneous landscape. Exp. Derm. 2022, 31, 13–30. [Google Scholar] [CrossRef] [PubMed]
  94. Marseglia, M.; Amaro, A.; Solari, N.; Gangemi, R.; Croce, E.; Tanda, E.T.; Spagnolo, F.; Filaci, G.; Pfeffer, U.; Croce, M. How to Make Immunotherapy an Effective Therapeutic Choice for Uveal Melanoma. Cancers 2021, 13, 2043. [Google Scholar] [CrossRef] [PubMed]
  95. Kaunitz, G.J.; Cottrell, T.R.; Lilo, M.; Muthappan, V.; Esandrio, J.; Berry, S.; Xu, H.; Ogurtsova, A.; Anders, R.A.; Fischer, A.H.; et al. Melanoma subtypes demonstrate distinct PD-L1 expression profiles. Lab. Invest. 2017, 97, 1063–1071. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Javed, A.; Arguello, D.; Johnston, C.; Gatalica, Z.; Terai, M.; Weight, R.M.; Orloff, M.; Mastrangelo, M.J.; Sato, T. PD-L1 expression in tumor metastasis is different between uveal melanoma and cutaneous melanoma. Immunotherapy 2017, 9, 1323–1330. [Google Scholar] [CrossRef] [Green Version]
  97. Wessely, A.; Steeb, T.; Erdmann, M.; Heinzerling, L.; Vera, J.; Schlaak, M.; Berking, C.; Heppt, M.V. The Role of Immune Checkpoint Blockade in Uveal Melanoma. Int. J. Mol. Sci. 2020, 21, 879. [Google Scholar] [CrossRef] [Green Version]
  98. de Vos, L.; Carrillo Cano, T.M.; Zarbl, R.; Klumper, N.; Ralser, D.J.; Franzen, A.; Herr, E.; Gabrielpillai, J.; Vogt, T.J.; Dietrich, J.; et al. CTLA4, PD-1, PD-L1, PD-L2, TIM-3, TIGIT, and LAG3 DNA Methylation Is Associated with BAP1 -Aberrancy, Transcriptional Activity, and Overall Survival in Uveal Melanoma. J. Immunother. 2022, 45, 324–334. [Google Scholar] [CrossRef]
  99. Nathan, P.; Hassel, J.C.; Rutkowski, P.; Baurain, J.F.; Butler, M.O.; Schlaak, M.; Sullivan, R.J.; Ochsenreither, S.; Dummer, R.; Kirkwood, J.M.; et al. Overall Survival Benefit with Tebentafusp in Metastatic Uveal Melanoma. N Engl. J. Med. 2021, 385, 1196–1206. [Google Scholar] [CrossRef]
  100. Ny, L.; Jespersen, H.; Karlsson, J.; Alsen, S.; Filges, S.; All-Eriksson, C.; Andersson, B.; Carneiro, A.; Helgadottir, H.; Levin, M.; et al. The PEMDAC phase 2 study of pembrolizumab and entinostat in patients with metastatic uveal melanoma. Nat. Commun. 2021, 12, 5155. [Google Scholar] [CrossRef] [PubMed]
  101. Spencer, K.R.; Mehnert, J.M. Mucosal Melanoma: Epidemiology, Biology and Treatment. Cancer Treat. Res. 2016, 167, 295–320. [Google Scholar] [CrossRef]
  102. Postow, M.A.; Hamid, O.; Carvajal, R.D. Mucosal melanoma: Pathogenesis, clinical behavior, and management. Curr. Oncol. Rep. 2012, 14, 441–448. [Google Scholar] [CrossRef] [PubMed]
  103. Nassar, K.W.; Tan, A.C. The mutational landscape of mucosal melanoma. Semin. Cancer Biol. 2020, 61, 139–148. [Google Scholar] [CrossRef]
  104. Wang, L.; Liu, H. PD-L1 expression in 117 sinonasal mucosal melanomas and its association with clinical outcome. Ann. Diagn Pathol 2022, 60, 151976. [Google Scholar] [CrossRef] [PubMed]
  105. Newell, F.; Kong, Y.; Wilmott, J.S.; Johansson, P.A.; Ferguson, P.M.; Cui, C.; Li, Z.; Kazakoff, S.H.; Burke, H.; Dodds, T.J.; et al. Whole-genome landscape of mucosal melanoma reveals diverse drivers and therapeutic targets. Nat. Commun. 2019, 10, 3163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Mao, L.; Qi, Z.; Zhang, L.; Guo, J.; Si, L. Immunotherapy in Acral and Mucosal Melanoma: Current Status and Future Directions. Front. Immunol. 2021, 12, 680407. [Google Scholar] [CrossRef] [PubMed]
  107. Moya-Plana, A.; Herrera Gomez, R.G.; Rossoni, C.; Dercle, L.; Ammari, S.; Girault, I.; Roy, S.; Scoazec, J.Y.; Vagner, S.; Janot, F.; et al. Evaluation of the efficacy of immunotherapy for non-resectable mucosal melanoma. Cancer Immunol. Immunother. 2019, 68, 1171–1178. [Google Scholar] [CrossRef] [PubMed]
  108. Shoushtari, A.N.; Wagstaff, J.; Ascierto, P.A.; Butler, M.O.; Lao, C.D.; Marquez-Rodas, I.; Chiarion-Sileni, V.; Dummer, R.; Ferrucci, P.F.; Lorigan, P.; et al. CheckMate 067: Long-term outcomes in patients with mucosal melanoma. J. Clin. Oncol. 2020, 38, 10019. [Google Scholar] [CrossRef]
  109. Simonetti, O.; Lucarini, G.; Rubini, C.; Lazzarini, R.; Di Primio, R.; Offidani, A. Clinical and prognostic significance of survivin, AKT and VEGF in primary mucosal oral melanoma. Anticancer. Res. 2015, 35, 2113–2120. [Google Scholar]
  110. Sheng, X.; Yan, X.; Chi, Z.; Si, L.; Cui, C.; Tang, B.; Li, S.; Mao, L.; Lian, B.; Wang, X.; et al. Axitinib in Combination with Toripalimab, a Humanized Immunoglobulin G4 Monoclonal Antibody Against Programmed Cell Death-1, in Patients with Metastatic Mucosal Melanoma: An Open-Label Phase IB Trial. J. Clin. Oncol. 2019, 37, 2987–2999. [Google Scholar] [CrossRef]
  111. Li, S.; Wu, X.; Yan, X.; Zhou, L.; Chi, Z.; Si, L.; Cui, C.; Tang, B.; Mao, L.; Lian, B.; et al. Toripalimab plus axitinib in patients with metastatic mucosal melanoma: 3-year survival update and biomarker analysis. J. Immunother. Cancer 2022, 10, e004036. [Google Scholar] [CrossRef]
  112. Mao, L.; Fang, M.; Chen, Y.; Wei, X.; Cao, J.; Lin, J.; Zhang, P.; Chen, L.; Cao, X.; Chen, Y.; et al. Atezolizumab Plus Bevacizumab in Patients with Unresectable or Metastatic Mucosal Melanoma: A Multicenter, Open-label, Single-arm Phase 2 Study. Clin. Cancer Res. 2022, 28, 4642–4648. [Google Scholar] [CrossRef] [PubMed]
  113. Global Burden of Disease Cancer, C.; Fitzmaurice, C.; Abate, D.; Abbasi, N.; Abbastabar, H.; Abd-Allah, F.; Abdel-Rahman, O.; Abdelalim, A.; Abdoli, A.; Abdollahpour, I.; et al. Global, Regional, and National Cancer Incidence, Mortality, Years of Life Lost, Years Lived with Disability, and Disability-Adjusted Life-Years for 29 Cancer Groups, 1990 to 2017: A Systematic Analysis for the Global Burden of Disease Study. JAMA Oncol. 2019, 5, 1749–1768. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Kacew, A.J.; Dharaneeswaran, H.; Starrett, G.J.; Thakuria, M.; LeBoeuf, N.R.; Silk, A.W.; DeCaprio, J.A.; Hanna, G.J. Predictors of immunotherapy benefit in Merkel cell carcinoma. Oncotarget 2020, 11, 4401–4410. [Google Scholar] [CrossRef] [PubMed]
  115. Pickering, C.R.; Zhou, J.H.; Lee, J.J.; Drummond, J.A.; Peng, S.A.; Saade, R.E.; Tsai, K.Y.; Curry, J.L.; Tetzlaff, M.T.; Lai, S.Y.; et al. Mutational landscape of aggressive cutaneous squamous cell carcinoma. Clin. Cancer Res. 2014, 20, 6582–6592. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Yarchoan, M.; Hopkins, A.; Jaffee, E.M. Tumor Mutational Burden and Response Rate to PD-1 Inhibition. N Engl. J. Med. 2017, 377, 2500–2501. [Google Scholar] [CrossRef]
  117. Rischin, D.; Khushalani, N.I.; Schmults, C.D.; Guminski, A.; Chang, A.L.S.; Lewis, K.D.; Lim, A.M.; Hernandez-Aya, L.; Hughes, B.G.M.; Schadendorf, D.; et al. Integrated analysis of a phase 2 study of cemiplimab in advanced cutaneous squamous cell carcinoma: Extended follow-up of outcomes and quality of life analysis. J. Immunother. Cancer 2021, 9, e002757. [Google Scholar] [CrossRef]
  118. Hughes, B.G.M.; Munoz-Couselo, E.; Mortier, L.; Bratland, A.; Gutzmer, R.; Roshdy, O.; Gonzalez Mendoza, R.; Schachter, J.; Arance, A.; Grange, F.; et al. Pembrolizumab for locally advanced and recurrent/metastatic cutaneous squamous cell carcinoma (KEYNOTE-629 study): An open-label, nonrandomized, multicenter, phase II trial. Ann. Oncol. 2021, 32, 1276–1285. [Google Scholar] [CrossRef]
  119. Ramella Munhoz, R.; De Camargo, V.P.; Nader Marta, G.; Queiroz, M.M.; Cury-Martins, J.; Nardo, M.; Chaul-Barbosa, C.; Ricci, H.; De Mattos, M.R.; De Menezes, T.A.F.; et al. 1064P Final results of CA209-9JC: A phase II study of first-line nivolumab in patients with advanced cutaneous squamous cell carcinoma. Ann. Oncol. 2021, 32, S885. [Google Scholar] [CrossRef]
  120. D’Angelo, S.P.; Lebbe, C.; Mortier, L.; Brohl, A.S.; Fazio, N.; Grob, J.J.; Prinzi, N.; Hanna, G.J.; Hassel, J.C.; Kiecker, F.; et al. First-line avelumab in a cohort of 116 patients with metastatic Merkel cell carcinoma (JAVELIN Merkel 200): Primary and biomarker analyses of a phase II study. J. Immunother. Cancer 2021, 9, e002646. [Google Scholar] [CrossRef]
  121. Nghiem, P.; Bhatia, S.; Lipson, E.J.; Sharfman, W.H.; Kudchadkar, R.R.; Brohl, A.S.; Friedlander, P.A.; Daud, A.; Kluger, H.M.; Reddy, S.A.; et al. Durable Tumor Regression and Overall Survival in Patients with Advanced Merkel Cell Carcinoma Receiving Pembrolizumab as First-Line Therapy. J. Clin. Oncol. 2019, 37, 693–702. [Google Scholar] [CrossRef]
  122. Stratigos, A.J.; Sekulic, A.; Peris, K.; Bechter, O.; Prey, S.; Kaatz, M.; Lewis, K.D.; Basset-Seguin, N.; Chang, A.L.S.; Dalle, S.; et al. Cemiplimab in locally advanced basal cell carcinoma after hedgehog inhibitor therapy: An open-label, multi-centre, single-arm, phase 2 trial. Lancet Oncol. 2021, 22, 848–857. [Google Scholar] [CrossRef] [PubMed]
  123. Schmults, C.D.; Blitzblau, R.; Aasi, S.Z.; Alam, M.; Andersen, J.S.; Baumann, B.C.; Bordeaux, J.; Chen, P.L.; Chin, R.; Contreras, C.M.; et al. NCCN Guidelines(R) Insights: Squamous Cell Skin Cancer, Version 1.2022. J. Natl. Compr. Canc. Netw. 2021, 19, 1382–1394. [Google Scholar] [CrossRef] [PubMed]
  124. Gross, N.D.; Miller, D.M.; Khushalani, N.I.; Divi, V.; Ruiz, E.S.; Lipson, E.J.; Meier, F.; Su, Y.B.; Swiecicki, P.L.; Atlas, J.; et al. Neoadjuvant Cemiplimab for Stage II to IV Cutaneous Squamous-Cell Carcinoma. N. Engl. J. Med. 2022, 387, 1557–1568. [Google Scholar] [CrossRef] [PubMed]
  125. Harms, K.L.; Healy, M.A.; Nghiem, P.; Sober, A.J.; Johnson, T.M.; Bichakjian, C.K.; Wong, S.L. Analysis of Prognostic Factors from 9387 Merkel Cell Carcinoma Cases Forms the Basis for the New 8th Edition AJCC Staging System. Ann. Surg. Oncol. 2016, 23, 3564–3571. [Google Scholar] [CrossRef] [PubMed]
  126. D’Angelo, S.P.; Bhatia, S.; Brohl, A.S.; Hamid, O.; Mehnert, J.M.; Terheyden, P.; Shih, K.C.; Brownell, I.; Lebbe, C.; Lewis, K.D.; et al. Avelumab in patients with previously treated metastatic Merkel cell carcinoma: Long-term data and biomarker analyses from the single-arm phase 2 JAVELIN Merkel 200 trial. J. Immunother. Cancer 2020, 8, e000674. [Google Scholar] [CrossRef] [PubMed]
  127. Nghiem, P.; Bhatia, S.; Lipson, E.J.; Sharfman, W.H.; Kudchadkar, R.R.; Brohl, A.S.; Friedlander, P.A.; Daud, A.; Kluger, H.M.; Reddy, S.A.; et al. Three-year survival, correlates and salvage therapies in patients receiving first-line pembrolizumab for advanced Merkel cell carcinoma. J. Immunother. Cancer 2021, 9, e002478. [Google Scholar] [CrossRef] [PubMed]
  128. Wadhera, A.; Fazio, M.; Bricca, G.; Stanton, O. Metastatic basal cell carcinoma: A case report and literature review. How accurate is our incidence data? Derm. Online J. 2006, 12, 7. [Google Scholar] [CrossRef]
  129. Bichakjian, C.K.; Olencki, T.; Aasi, S.Z.; Alam, M.; Andersen, J.S.; Berg, D.; Bowen, G.M.; Cheney, R.T.; Daniels, G.A.; Glass, L.F.; et al. Basal Cell Skin Cancer, Version 1.2016, NCCN Clinical Practice Guidelines in Oncology. J. Natl. Compr. Canc. Netw. 2016, 14, 574–597. [Google Scholar] [CrossRef]
  130. Larkin, J.; Chiarion-Sileni, V.; Gonzalez, R.; Grob, J.J.; Rutkowski, P.; Lao, C.D.; Cowey, C.L.; Schadendorf, D.; Wagstaff, J.; Dummer, R.; et al. Five-Year Survival with Combined Nivolumab and Ipilimumab in Advanced Melanoma. N. Engl. J. Med. 2019, 381, 1535–1546. [Google Scholar] [CrossRef] [Green Version]
  131. Wolchok, J.D.; Chiarion-Sileni, V.; Gonzalez, R.; Rutkowski, P.; Grob, J.J.; Cowey, C.L.; Lao, C.D.; Wagstaff, J.; Schadendorf, D.; Ferrucci, P.F.; et al. Overall Survival with Combined Nivolumab and Ipilimumab in Advanced Melanoma. N. Engl. J. Med. 2017, 377, 1345–1356. [Google Scholar] [CrossRef]
  132. Hodi, F.S.; Chiarion-Sileni, V.; Gonzalez, R.; Grob, J.J.; Rutkowski, P.; Cowey, C.L.; Lao, C.D.; Schadendorf, D.; Wagstaff, J.; Dummer, R.; et al. Nivolumab plus ipilimumab or nivolumab alone versus ipilimumab alone in advanced melanoma (CheckMate 067): 4-year outcomes of a multicentre, randomised, phase 3 trial. Lancet Oncol. 2018, 19, 1480–1492. [Google Scholar] [CrossRef] [PubMed]
  133. Goodman, A.M.; Kato, S.; Bazhenova, L.; Patel, S.P.; Frampton, G.M.; Miller, V.; Stephens, P.J.; Daniels, G.A.; Kurzrock, R. Tumor Mutational Burden as an Independent Predictor of Response to Immunotherapy in Diverse Cancers. Mol. Cancer 2017, 16, 2598–2608. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Forschner, A.; Battke, F.; Hadaschik, D.; Schulze, M.; Weissgraeber, S.; Han, C.T.; Kopp, M.; Frick, M.; Klumpp, B.; Tietze, N.; et al. Tumor mutation burden and circulating tumor DNA in combined CTLA-4 and PD-1 antibody therapy in metastatic melanoma—Results of a prospective biomarker study. J. Immunother. Cancer 2019, 7, 180. [Google Scholar] [CrossRef] [Green Version]
  135. Peng, W.; Chen, J.Q.; Liu, C.; Malu, S.; Creasy, C.; Tetzlaff, M.T.; Xu, C.; McKenzie, J.A.; Zhang, C.; Liang, X.; et al. Loss of PTEN Promotes Resistance to T Cell-Mediated Immunotherapy. Cancer Discov. 2016, 6, 202–216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Piro, G.; Carbone, C.; Carbognin, L.; Pilotto, S.; Ciccarese, C.; Iacovelli, R.; Milella, M.; Bria, E.; Tortora, G. Revising PTEN in the Era of Immunotherapy: New Perspectives for an Old Story. Cancers 2019, 11, 1525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Dong, Y.; Richards, J.A.; Gupta, R.; Aung, P.P.; Emley, A.; Kluger, Y.; Dogra, S.K.; Mahalingam, M.; Wajapeyee, N. PTEN functions as a melanoma tumor suppressor by promoting host immune response. Oncogene 2014, 33, 4632–4642. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Zaretsky, J.M.; Garcia-Diaz, A.; Shin, D.S.; Escuin-Ordinas, H.; Hugo, W.; Hu-Lieskovan, S.; Torrejon, D.Y.; Abril-Rodriguez, G.; Sandoval, S.; Barthly, L.; et al. Mutations Associated with Acquired Resistance to PD-1 Blockade in Melanoma. N Engl. J. Med. 2016, 375, 819–829. [Google Scholar] [CrossRef]
  139. Olbryt, M.; Rajczykowski, M.; Widlak, W. Biological Factors behind Melanoma Response to Immune Checkpoint Inhibitors. Int. J. Mol. Sci. 2020, 21, 4071. [Google Scholar] [CrossRef]
  140. Castagnoli, L.; Cancila, V.; Cordoba-Romero, S.L.; Faraci, S.; Talarico, G.; Belmonte, B.; Iorio, M.V.; Milani, M.; Volpari, T.; Chiodoni, C.; et al. WNT signaling modulates PD-L1 expression in the stem cell compartment of triple-negative breast cancer. Oncogene 2019, 38, 4047–4060. [Google Scholar] [CrossRef] [Green Version]
  141. Bockmayr, M.; Mohme, M.; Klauschen, F.; Winkler, B.; Budczies, J.; Rutkowski, S.; Schüller, U. Subgroup-specific immune and stromal microenvironment in medulloblastoma. OncoImmunology 2018, 7, e1462430. [Google Scholar] [CrossRef] [Green Version]
  142. Spranger, S.; Bao, R.; Gajewski, T.F. Melanoma-intrinsic β-catenin signalling prevents anti-tumour immunity. Nature 2015, 523, 231–235. [Google Scholar] [CrossRef] [PubMed]
  143. Lengauer, C.; Kinzler, K.W.; Vogelstein, B. Genetic instabilities in human cancers. Nature 1998, 396, 643–649. [Google Scholar] [CrossRef] [PubMed]
  144. McGranahan, N.; Furness, A.J.S.; Rosenthal, R.; Ramskov, S.; Lyngaa, R.; Saini, S.K.; Jamal-Hanjani, M.; Wilson, G.A.; Birkbak, N.J.; Hiley, C.T.; et al. Clonal neoantigens elicit T cell immunoreactivity and sensitivity to immune checkpoint blockade. Science 2016, 351, 1463–1469. [Google Scholar] [CrossRef] [Green Version]
  145. Roh, W.; Chen, P.-L.; Reuben, A.; Spencer Christine, N.; Prieto Peter, A.; Miller John, P.; Gopalakrishnan, V.; Wang, F.; Cooper Zachary, A.; Reddy Sangeetha, M.; et al. Integrated molecular analysis of tumor biopsies on sequential CTLA-4 and PD-1 blockade reveals markers of response and resistance. Sci. Transl. Med. 2017, 9, eaah3560. [Google Scholar] [CrossRef] [Green Version]
  146. Spiliopoulou, P.; Yang, S.Y.C.; Bruce, J.P.; Wang, B.X.; Berman, H.K.; Pugh, T.J.; Siu, L.L. All is not lost: Learning from 9p21 loss in cancer. Trends Immunol. 2022, 43, 379–390. [Google Scholar] [CrossRef] [PubMed]
  147. Han, G.; Yang, G.; Hao, D.; Lu, Y.; Thein, K.; Simpson, B.S.; Chen, J.; Sun, R.; Alhalabi, O.; Wang, R.; et al. 9p21 loss confers a cold tumor immune microenvironment and primary resistance to immune checkpoint therapy. Nat. Commun. 2021, 12, 5606. [Google Scholar] [CrossRef] [PubMed]
  148. Davoli, T.; Uno, H.; Wooten Eric, C.; Elledge Stephen, J. Tumor aneuploidy correlates with markers of immune evasion and with reduced response to immunotherapy. Science 2017, 355, eaaf8399. [Google Scholar] [CrossRef] [Green Version]
  149. Mitra, A.; Andrews, M.C.; Roh, W.; De Macedo, M.P.; Hudgens, C.W.; Carapeto, F.; Singh, S.; Reuben, A.; Wang, F.; Mao, X.; et al. Spatially resolved analyses link genomic and immune diversity and reveal unfavorable neutrophil activation in melanoma. Nat. Commun. 2020, 11, 1839. [Google Scholar] [CrossRef] [Green Version]
  150. Ji, R.R.; Chasalow, S.D.; Wang, L.; Hamid, O.; Schmidt, H.; Cogswell, J.; Alaparthy, S.; Berman, D.; Jure-Kunkel, M.; Siemers, N.O.; et al. An immune-active tumor microenvironment favors clinical response to ipilimumab. Cancer Immunol. Immunother. 2012, 61, 1019–1031. [Google Scholar] [CrossRef]
  151. Ribas, A.; Robert, C.; Hodi, F.S.; Wolchok, J.D.; Joshua, A.M.; Hwu, W.-J.; Weber, J.S.; Zarour, H.M.; Kefford, R.; Loboda, A.; et al. Association of response to programmed death receptor 1 (PD-1) blockade with pembrolizumab (MK-3475) with an interferon-inflammatory immune gene signature. J. Clin. Oncol. 2015, 33, 3001. [Google Scholar] [CrossRef]
  152. Ayers, M.; Lunceford, J.; Nebozhyn, M.; Murphy, E.; Loboda, A.; Kaufman, D.R.; Albright, A.; Cheng, J.D.; Kang, S.P.; Shankaran, V.; et al. IFN-gamma-related mRNA profile predicts clinical response to PD-1 blockade. J. Clin. Invest. 2017, 127, 2930–2940. [Google Scholar] [CrossRef] [PubMed]
  153. Sharma, P.; Hu-Lieskovan, S.; Wargo, J.A.; Ribas, A. Primary, Adaptive, and Acquired Resistance to Cancer Immunotherapy. Cell 2017, 168, 707–723. [Google Scholar] [CrossRef] [Green Version]
  154. Jiang, P.; Gu, S.; Pan, D.; Fu, J.; Sahu, A.; Hu, X.; Li, Z.; Traugh, N.; Bu, X.; Li, B.; et al. Signatures of T cell dysfunction and exclusion predict cancer immunotherapy response. Nat. Med. 2018, 24, 1550–1558. [Google Scholar] [CrossRef] [PubMed]
  155. Rossi, N.; Lee, K.A.; Bermudez, M.V.; Visconti, A.; Thomas, A.M.; Bolte, L.A.; Björk, J.R.; de Ruijter, L.K.; Newton-Bishop, J.; Harland, M.; et al. Circulating inflammatory proteins associate with response to immune checkpoint inhibition therapy in patients with advanced melanoma. eBioMedicine 2022, 83, 104235. [Google Scholar] [CrossRef]
  156. Lebrun, J.J. The Dual Role of TGFbeta in Human Cancer: From Tumor Suppression to Cancer Metastasis. ISRN Mol. Biol. 2012, 2012, 381428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Robbins, P.D.; Morelli, A.E. Regulation of immune responses by extracellular vesicles. Nat. Rev. Immunol. 2014, 14, 195–208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Kalluri, R. The biology and function of exosomes in cancer. J. Clin. Invest. 2016, 126, 1208–1215. [Google Scholar] [CrossRef] [PubMed]
  159. Becker, A.; Thakur, B.K.; Weiss, J.M.; Kim, H.S.; Peinado, H.; Lyden, D. Extracellular Vesicles in Cancer: Cell-to-Cell Mediators of Metastasis. Cancer Cell 2016, 30, 836–848. [Google Scholar] [CrossRef] [Green Version]
  160. Chen, G.; Huang, A.C.; Zhang, W.; Zhang, G.; Wu, M.; Xu, W.; Yu, Z.; Yang, J.; Wang, B.; Sun, H.; et al. Exosomal PD-L1 contributes to immunosuppression and is associated with anti-PD-1 response. Nature 2018, 560, 382–386. [Google Scholar] [CrossRef]
  161. Smyth, M.J.; Ngiow, S.F.; Ribas, A.; Teng, M.W. Combination cancer immunotherapies tailored to the tumour microenvironment. Nat. Rev. Clin. Oncol. 2016, 13, 143–158. [Google Scholar] [CrossRef]
  162. Tumeh, P.C.; Harview, C.L.; Yearley, J.H.; Shintaku, I.P.; Taylor, E.J.; Robert, L.; Chmielowski, B.; Spasic, M.; Henry, G.; Ciobanu, V.; et al. PD-1 blockade induces responses by inhibiting adaptive immune resistance. Nature 2014, 515, 568–571. [Google Scholar] [CrossRef] [PubMed]
  163. Cabrita, R.; Lauss, M.; Sanna, A.; Donia, M.; Skaarup Larsen, M.; Mitra, S.; Johansson, I.; Phung, B.; Harbst, K.; Vallon-Christersson, J.; et al. Tertiary lymphoid structures improve immunotherapy and survival in melanoma. Nature 2020, 577, 561–565. [Google Scholar] [CrossRef] [PubMed]
  164. Nelson, B.H. CD20+ B Cells: The Other Tumor-Infiltrating Lymphocytes. J. Immunol. 2010, 185, 4977–4982. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Ouyang, Z.; Wu, H.; Li, L.; Luo, Y.; Li, X.; Huang, G. Regulatory T cells in the immunotherapy of melanoma. Tumour. Biol. 2016, 37, 77–85. [Google Scholar] [CrossRef] [PubMed]
  166. Tobin, R.P.; Jordan, K.R.; Kapoor, P.; Spongberg, E.; Davis, D.; Vorwald, V.M.; Couts, K.L.; Gao, D.; Smith, D.E.; Borgers, J.S.W.; et al. IL-6 and IL-8 Are Linked with Myeloid-Derived Suppressor Cell Accumulation and Correlate with Poor Clinical Outcomes in Melanoma Patients. Front. Oncol. 2019, 9, 1223. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Noy, R.; Pollard, J.W. Tumor-associated macrophages: From mechanisms to therapy. Immunity 2014, 41, 49–61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Helmink, B.A.; Reddy, S.M.; Gao, J.; Zhang, S.; Basar, R.; Thakur, R.; Yizhak, K.; Sade-Feldman, M.; Blando, J.; Han, G.; et al. B cells and tertiary lymphoid structures promote immunotherapy response. Nature 2020, 577, 549–555. [Google Scholar] [CrossRef]
  169. Litchfield, K.; Reading, J.L.; Puttick, C.; Thakkar, K.; Abbosh, C.; Bentham, R.; Watkins, T.B.K.; Rosenthal, R.; Biswas, D.; Rowan, A.; et al. Meta-analysis of tumor- and T cell-intrinsic mechanisms of sensitization to checkpoint inhibition. Cell 2021, 184, 596–614. [Google Scholar] [CrossRef]
  170. Griss, J.; Bauer, W.; Wagner, C.; Simon, M.; Chen, M.; Grabmeier-Pfistershammer, K.; Maurer-Granofszky, M.; Roka, F.; Penz, T.; Bock, C.; et al. B cells sustain inflammation and predict response to immune checkpoint blockade in human melanoma. Nat. Commun. 2019, 10, 4186. [Google Scholar] [CrossRef] [Green Version]
  171. Vanhersecke, L.; Brunet, M.; Guegan, J.P.; Rey, C.; Bougouin, A.; Cousin, S.; Moulec, S.L.; Besse, B.; Loriot, Y.; Larroquette, M.; et al. Mature tertiary lymphoid structures predict immune checkpoint inhibitor efficacy in solid tumors independently of PD-L1 expression. Nat. Cancer 2021, 2, 794–802. [Google Scholar] [CrossRef]
  172. Lauss, M.; Donia, M.; Svane, I.M.; Jonsson, G. B Cells and Tertiary Lymphoid Structures: Friends or Foes in Cancer Immunotherapy? Clin. Cancer Res. 2022, 28, 1751–1758. [Google Scholar] [CrossRef] [PubMed]
  173. Hsieh, C.H.; Jian, C.Z.; Lin, L.I.; Low, G.S.; Ou, P.Y.; Hsu, C.; Ou, D.L. Potential Role of CXCL13/CXCR5 Signaling in Immune Checkpoint Inhibitor Treatment in Cancer. Cancers 2022, 14, 294. [Google Scholar] [CrossRef] [PubMed]
  174. Sucker, A.; Zhao, F.; Real, B.; Heeke, C.; Bielefeld, N.; Maßen, S.; Horn, S.; Moll, I.; Maltaner, R.; Horn, P.A.; et al. Genetic evolution of T-cell resistance in the course of melanoma progression. Clin. Cancer Res. 2014, 20, 6593–6604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Hicklin, D.J.; Wang, Z.; Arienti, F.; Rivoltini, L.; Parmiani, G.; Ferrone, S. beta2-Microglobulin mutations, HLA class I antigen loss, and tumor progression in melanoma. J. Clin. Investig. 1998, 101, 2720–2729. [Google Scholar] [CrossRef] [PubMed]
  176. Sade-Feldman, M.; Jiao, Y.J.; Chen, J.H.; Rooney, M.S.; Barzily-Rokni, M.; Eliane, J.P.; Bjorgaard, S.L.; Hammond, M.R.; Vitzthum, H.; Blackmon, S.M.; et al. Resistance to checkpoint blockade therapy through inactivation of antigen presentation. Nat. Commun. 2017, 8, 1136. [Google Scholar] [CrossRef] [Green Version]
  177. Patel, S.J.; Sanjana, N.E.; Kishton, R.J.; Eidizadeh, A.; Vodnala, S.K.; Cam, M.; Gartner, J.J.; Jia, L.; Steinberg, S.M.; Yamamoto, T.N.; et al. Identification of essential genes for cancer immunotherapy. Nature 2017, 548, 537–542. [Google Scholar] [CrossRef] [Green Version]
  178. Hogan, S.A.; Courtier, A.; Cheng, P.F.; Jaberg-Bentele, N.F.; Goldinger, S.M.; Manuel, M.; Perez, S.; Plantier, N.; Mouret, J.F.; Nguyen-Kim, T.D.L.; et al. Peripheral Blood TCR Repertoire Profiling May Facilitate Patient Stratification for Immunotherapy against Melanoma. Cancer Immunol. Res. 2019, 7, 77–85. [Google Scholar] [CrossRef] [Green Version]
  179. Riaz, N.; Havel, J.J.; Makarov, V.; Desrichard, A.; Urba, W.J.; Sims, J.S.; Hodi, F.S.; Martin-Algarra, S.; Mandal, R.; Sharfman, W.H.; et al. Tumor and Microenvironment Evolution during Immunotherapy with Nivolumab. Cell 2017, 171, 934–949. [Google Scholar] [CrossRef] [Green Version]
  180. Han, J.; Duan, J.; Bai, H.; Wang, Y.; Wan, R.; Wang, X.; Chen, S.; Tian, Y.; Wang, D.; Fei, K.; et al. TCR Repertoire Diversity of Peripheral PD-1(+)CD8(+) T Cells Predicts Clinical Outcomes after Immunotherapy in Patients with Non-Small Cell Lung Cancer. Cancer Immunol. Res. 2020, 8, 146–154. [Google Scholar] [CrossRef]
  181. Poran, A.; Scherer, J.; Bushway, M.E.; Besada, R.; Balogh, K.N.; Wanamaker, A.; Williams, R.G.; Prabhakara, J.; Ott, P.A.; Hu-Lieskovan, S.; et al. Combined TCR Repertoire Profiles and Blood Cell Phenotypes Predict Melanoma Patient Response to Personalized Neoantigen Therapy plus Anti-PD-1. Cell Rep. Med. 2020, 1, 100141. [Google Scholar] [CrossRef]
  182. Valpione, S.; Mundra, P.A.; Galvani, E.; Campana, L.G.; Lorigan, P.; De Rosa, F.; Gupta, A.; Weightman, J.; Mills, S.; Dhomen, N.; et al. The T cell receptor repertoire of tumor infiltrating T cells is predictive and prognostic for cancer survival. Nat. Commun. 2021, 12, 4098. [Google Scholar] [CrossRef] [PubMed]
  183. Au, L.; Hatipoglu, E.; Robert de Massy, M.; Litchfield, K.; Beattie, G.; Rowan, A.; Schnidrig, D.; Thompson, R.; Byrne, F.; Horswell, S.; et al. Determinants of anti-PD-1 response and resistance in clear cell renal cell carcinoma. Cancer Cell 2021, 39, 1497–1518. [Google Scholar] [CrossRef] [PubMed]
  184. Ji, S.; Li, J.; Chang, L.; Zhao, C.; Jia, R.; Tan, Z.; Liu, R.; Zhang, Y.; Li, Y.; Yin, G.; et al. Peripheral blood T-cell receptor repertoire as a predictor of clinical outcomes in gastrointestinal cancer patients treated with PD-1 inhibitor. Clin. Transl. Oncol. 2021, 23, 1646–1656. [Google Scholar] [CrossRef] [PubMed]
  185. Valpione, S.; Galvani, E.; Tweedy, J.; Mundra, P.A.; Banyard, A.; Middlehurst, P.; Barry, J.; Mills, S.; Salih, Z.; Weightman, J.; et al. Immune-awakening revealed by peripheral T cell dynamics after one cycle of immunotherapy. Nat. Cancer 2020, 1, 210–221. [Google Scholar] [CrossRef]
  186. He, Y.; Liu, T.; Dai, S.; Xu, Z.; Wang, L.; Luo, F. Tumor-Associated Extracellular Matrix: How to Be a Potential Aide to Anti-tumor Immunotherapy? Front. Cell Dev. Biol. 2021, 9, 739161. [Google Scholar] [CrossRef]
  187. Peng, D.H.; Rodriguez, B.L.; Diao, L.; Chen, L.; Wang, J.; Byers, L.A.; Wei, Y.; Chapman, H.A.; Yamauchi, M.; Behrens, C.; et al. Collagen promotes anti-PD-1/PD-L1 resistance in cancer through LAIR1-dependent CD8(+) T cell exhaustion. Nat. Commun. 2020, 11, 4520. [Google Scholar] [CrossRef]
  188. Lebbink, R.J.; Raynal, N.; de Ruiter, T.; Bihan, D.G.; Farndale, R.W.; Meyaard, L. Identification of multiple potent binding sites for human leukocyte associated Ig-like receptor LAIR on collagens II and III. Matrix Biol. 2009, 28, 202–210. [Google Scholar] [CrossRef]
  189. Gopalakrishnan, V.; Helmink, B.A.; Spencer, C.N.; Reuben, A.; Wargo, J.A. The Influence of the Gut Microbiome on Cancer, Immunity, and Cancer Immunotherapy. Cancer Cell 2018, 33, 570–580. [Google Scholar] [CrossRef] [Green Version]
  190. Zitvogel, L.; Ma, Y.; Raoult, D.; Kroemer, G.; Gajewski, T.F. The microbiome in cancer immunotherapy: Diagnostic tools and therapeutic strategies. Science 2018, 359, 1366–1370. [Google Scholar] [CrossRef] [Green Version]
  191. Fessler, J.; Matson, V.; Gajewski, T.F. Exploring the emerging role of the microbiome in cancer immunotherapy. J. Immunother. Cancer 2019, 7, 108. [Google Scholar] [CrossRef]
  192. Xavier, J.B.; Young, V.B.; Skufca, J.; Ginty, F.; Testerman, T.; Pearson, A.T.; Macklin, P.; Mitchell, A.; Shmulevich, I.; Xie, L.; et al. The Cancer Microbiome: Distinguishing Direct and Indirect Effects Requires a Systemic View. Trends Cancer 2020, 6, 192–204. [Google Scholar] [CrossRef] [PubMed]
  193. Andrews, M.C.; Duong, C.P.M.; Gopalakrishnan, V.; Iebba, V.; Chen, W.S.; Derosa, L.; Khan, M.A.W.; Cogdill, A.P.; White, M.G.; Wong, M.C.; et al. Gut microbiota signatures are associated with toxicity to combined CTLA-4 and PD-1 blockade. Nat. Med. 2021, 27, 1432–1441. [Google Scholar] [CrossRef] [PubMed]
  194. Gopalakrishnan, V.; Spencer, C.N.; Nezi, L.; Reuben, A.; Andrews, M.C.; Karpinets, T.V.; Prieto, P.A.; Vicente, D.; Hoffman, K.; Wei, S.C.; et al. Gut microbiome modulates response to anti-PD-1 immunotherapy in melanoma patients. Science 2018, 359, 97–103. [Google Scholar] [CrossRef] [Green Version]
  195. Sivan, A.; Corrales, L.; Hubert, N.; Williams, J.B.; Aquino-Michaels, K.; Earley, Z.M.; Benyamin, F.W.; Lei, Y.M.; Jabri, B.; Alegre, M.L.; et al. Commensal Bifidobacterium promotes antitumor immunity and facilitates anti-PD-L1 efficacy. Science 2015, 350, 1084–1089. [Google Scholar] [CrossRef] [Green Version]
  196. Derosa, L.; Routy, B.; Fidelle, M.; Iebba, V.; Alla, L.; Pasolli, E.; Segata, N.; Desnoyer, A.; Pietrantonio, F.; Ferrere, G.; et al. Gut Bacteria Composition Drives Primary Resistance to Cancer Immunotherapy in Renal Cell Carcinoma Patients. Eur. Urol. 2020, 78, 195–206. [Google Scholar] [CrossRef]
  197. Jiang, S.; Geng, S.; Chen, Q.; Zhang, C.; Cheng, M.; Yu, Y.; Zhang, S.; Shi, N.; Dong, M. Effects of Concomitant Antibiotics Use on Immune Checkpoint Inhibitor Efficacy in Cancer Patients. Front. Oncol. 2022, 12, 823705. [Google Scholar] [CrossRef]
  198. Yang, M.; Wang, Y.; Yuan, M.; Tao, M.; Kong, C.; Li, H.; Tong, J.; Zhu, H.; Yan, X. Antibiotic administration shortly before or after immunotherapy initiation is correlated with poor prognosis in solid cancer patients: An up-to-date systematic review and meta-analysis. Int. Immunopharmacol. 2020, 88, 106876. [Google Scholar] [CrossRef]
  199. Hopkins, A.M.; Kichenadasse, G.; Karapetis, C.S.; Rowland, A.; Sorich, M.J. Concomitant Antibiotic Use and Survival in Urothelial Carcinoma Treated with Atezolizumab. Eur. Urol. 2020, 78, 540–543. [Google Scholar] [CrossRef]
  200. Elkrief, A.; El Raichani, L.; Richard, C.; Messaoudene, M.; Belkaid, W.; Malo, J.; Belanger, K.; Miller, W.; Jamal, R.; Letarte, N.; et al. Antibiotics are associated with decreased progression-free survival of advanced melanoma patients treated with immune checkpoint inhibitors. Oncoimmunology 2019, 8, e1568812. [Google Scholar] [CrossRef]
  201. He, Y.; Fu, L.; Li, Y.; Wang, W.; Gong, M.; Zhang, J.; Dong, X.; Huang, J.; Wang, Q.; Mackay, C.R.; et al. Gut microbial metabolites facilitate anticancer therapy efficacy by modulating cytotoxic CD8(+) T cell immunity. Cell Metab. 2021, 33, 988–1000. [Google Scholar] [CrossRef]
  202. Lu, Y.; Yuan, X.; Wang, M.; He, Z.; Li, H.; Wang, J.; Li, Q. Gut microbiota influence immunotherapy responses: Mechanisms and therapeutic strategies. J. Hematol. Oncol. 2022, 15, 47. [Google Scholar] [CrossRef] [PubMed]
  203. Ma, C.; Han, M.; Heinrich, B.; Fu, Q.; Zhang, Q.; Sandhu, M.; Agdashian, D.; Terabe, M.; Berzofsky, J.A.; Fako, V.; et al. Gut microbiome-mediated bile acid metabolism regulates liver cancer via NKT cells. Science 2018, 360, eaan5931. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Yang, W.; Yu, T.; Huang, X.; Bilotta, A.J.; Xu, L.; Lu, Y.; Sun, J.; Pan, F.; Zhou, J.; Zhang, W.; et al. Intestinal microbiota-derived short-chain fatty acids regulation of immune cell IL-22 production and gut immunity. Nat. Commun. 2020, 11, 4457. [Google Scholar] [CrossRef] [PubMed]
  205. Lam, K.C.; Araya, R.E.; Huang, A.; Chen, Q.; Di Modica, M.; Rodrigues, R.R.; Lopes, A.; Johnson, S.B.; Schwarz, B.; Bohrnsen, E.; et al. Microbiota triggers STING-type I IFN-dependent monocyte reprogramming of the tumor microenvironment. Cell 2021, 184, 5338–5356.e21. [Google Scholar] [CrossRef]
  206. Mager, L.F.; Burkhard, R.; Pett, N.; Cooke, N.C.A.; Brown, K.; Ramay, H.; Paik, S.; Stagg, J.; Groves, R.A.; Gallo, M.; et al. Microbiome-derived inosine modulates response to checkpoint inhibitor immunotherapy. Science 2020, 369, 1481–1489. [Google Scholar] [CrossRef]
  207. Song, X.; Sun, X.; Oh, S.F.; Wu, M.; Zhang, Y.; Zheng, W.; Geva-Zatorsky, N.; Jupp, R.; Mathis, D.; Benoist, C.; et al. Microbial bile acid metabolites modulate gut RORgamma(+) regulatory T cell homeostasis. Nature 2020, 577, 410–415. [Google Scholar] [CrossRef]
  208. Pushalkar, S.; Hundeyin, M.; Daley, D.; Zambirinis, C.P.; Kurz, E.; Mishra, A.; Mohan, N.; Aykut, B.; Usyk, M.; Torres, L.E.; et al. The Pancreatic Cancer Microbiome Promotes Oncogenesis by Induction of Innate and Adaptive Immune Suppression. Cancer Discov. 2018, 8, 403–416. [Google Scholar] [CrossRef] [Green Version]
  209. Mohan, V.; Das, A.; Sagi, I. Emerging roles of ECM remodeling processes in cancer. Semin. Cancer Biol. 2020, 62, 192–200. [Google Scholar] [CrossRef]
  210. Henstra, C.; van Praagh, J.; Olinga, P.; Nagelkerke, A. The gastrointestinal microbiota in colorectal cancer cell migration and invasion. Clin. Exp. Metastasis. 2021, 38, 495–510. [Google Scholar] [CrossRef]
  211. Fluckiger, A.; Daillere, R.; Sassi, M.; Sixt, B.S.; Liu, P.; Loos, F.; Richard, C.; Rabu, C.; Alou, M.T.; Goubet, A.G.; et al. Cross-reactivity between tumor MHC class I-restricted antigens and an enterococcal bacteriophage. Science 2020, 369, 936–942. [Google Scholar] [CrossRef]
  212. Bessell, C.A.; Isser, A.; Havel, J.J.; Lee, S.; Bell, D.R.; Hickey, J.W.; Chaisawangwong, W.; Glick Bieler, J.; Srivastava, R.; Kuo, F.; et al. Commensal bacteria stimulate antitumor responses via T cell cross-reactivity. JCI Insight. 2020, 5, 8. [Google Scholar] [CrossRef]
  213. Routy, B.; Le Chatelier, E.; Derosa, L.; Duong, C.P.M.; Alou, M.T.; Daillere, R.; Fluckiger, A.; Messaoudene, M.; Rauber, C.; Roberti, M.P.; et al. Gut microbiome influences efficacy of PD-1-based immunotherapy against epithelial tumors. Science 2018, 359, 91–97. [Google Scholar] [CrossRef] [Green Version]
  214. Baruch, E.N.; Youngster, I.; Ben-Betzalel, G.; Ortenberg, R.; Lahat, A.; Katz, L.; Adler, K.; Dick-Necula, D.; Raskin, S.; Bloch, N.; et al. Fecal microbiota transplant promotes response in immunotherapy-refractory melanoma patients. Science 2021, 371, 602–609. [Google Scholar] [CrossRef]
  215. Davar, D.; Dzutsev, A.K.; McCulloch, J.A.; Rodrigues, R.R.; Chauvin, J.M.; Morrison, R.M.; Deblasio, R.N.; Menna, C.; Ding, Q.; Pagliano, O.; et al. Fecal microbiota transplant overcomes resistance to anti-PD-1 therapy in melanoma patients. Science 2021, 371, 595–602. [Google Scholar] [CrossRef]
  216. Wan, J.C.M.; Massie, C.; Garcia-Corbacho, J.; Mouliere, F.; Brenton, J.D.; Caldas, C.; Pacey, S.; Baird, R.; Rosenfeld, N. Liquid biopsies come of age: Towards implementation of circulating tumour DNA. Nat. Rev. Cancer 2017, 17, 223–238. [Google Scholar] [CrossRef]
  217. De Mattos-Arruda, L.; Siravegna, G. How to use liquid biopsies to treat patients with cancer. ESMO Open 2021, 6, 100060. [Google Scholar] [CrossRef] [PubMed]
  218. Stadler, J.C.; Belloum, Y.; Deitert, B.; Sementsov, M.; Heidrich, I.; Gebhardt, C.; Keller, L.; Pantel, K. Current and Future Clinical Applications of ctDNA in Immuno-Oncology. Cancer Res. 2022, 82, 349–358. [Google Scholar] [CrossRef]
  219. Kruger, S.; Heinemann, V.; Ross, C.; Diehl, F.; Nagel, D.; Ormanns, S.; Liebmann, S.; Prinz-Bravin, I.; Westphalen, C.B.; Haas, M.; et al. Repeated mutKRAS ctDNA measurements represent a novel and promising tool for early response prediction and therapy monitoring in advanced pancreatic cancer. Ann. Oncol. 2018, 29, 2348–2355. [Google Scholar] [CrossRef]
  220. Elez, E.; Chianese, C.; Sanz-Garcia, E.; Martinelli, E.; Noguerido, A.; Mancuso, F.M.; Caratu, G.; Matito, J.; Grasselli, J.; Cardone, C.; et al. Impact of circulating tumor DNA mutant allele fraction on prognosis in RAS-mutant metastatic colorectal cancer. Mol. Oncol. 2019, 13, 1827–1835. [Google Scholar] [CrossRef] [Green Version]
  221. Cai, Z.; Chen, G.; Zeng, Y.; Dong, X.; Li, Z.; Huang, Y.; Xin, F.; Qiu, L.; Xu, H.; Zhang, W.; et al. Comprehensive Liquid Profiling of Circulating Tumor DNA and Protein Biomarkers in Long-Term Follow-Up Patients with Hepatocellular Carcinoma. Clin. Cancer Res. 2019, 25, 5284–5294. [Google Scholar] [CrossRef]
  222. Nygård, L.; Ahlborn, L.B.; Persson, G.F.; Chandrananda, D.; Langer, J.W.; Fischer, B.M.; Langer, S.W.; Gabrielaite, M.; Kjær, A.; Rosenfeld, N.; et al. Circulating cell free DNA during definitive chemo-radiotherapy in non-small cell lung cancer patients—Initial observations. PLoS ONE 2020, 15, e0231884. [Google Scholar] [CrossRef] [PubMed]
  223. Anagnostou, V.; Forde, P.M.; White, J.R.; Niknafs, N.; Hruban, C.; Naidoo, J.; Marrone, K.; Sivakumar, I.K.A.; Bruhm, D.C.; Rosner, S.; et al. Dynamics of Tumor and Immune Responses during Immune Checkpoint Blockade in Non-Small Cell Lung Cancer. Cancer Res. 2019, 79, 1214–1225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Cescon, D.W.; Bratman, S.V.; Chan, S.M.; Siu, L.L. Circulating tumor DNA and liquid biopsy in oncology. Nat. Cancer 2020, 1, 276–290. [Google Scholar] [CrossRef] [PubMed]
  225. Burgener, J.M.; Zou, J.; Zhao, Z.; Zheng, Y.; Shen, S.Y.; Huang, S.H.; Keshavarzi, S.; Xu, W.; Liu, F.F.; Liu, G.; et al. Tumor-Naive Multimodal Profiling of Circulating Tumor DNA in Head and Neck Squamous Cell Carcinoma. Clin. Cancer Res. 2021, 27, 4230–4244. [Google Scholar] [CrossRef]
  226. García-Pardo, M.; Makarem, M.; Li, J.J.N.; Kelly, D.; Leighl, N.B. Integrating circulating-free DNA (cfDNA) analysis into clinical practice: Opportunities and challenges. Br. J. Cancer 2022, 127, 592–602. [Google Scholar] [CrossRef]
  227. Tie, J.; Wang, Y.; Tomasetti, C.; Li, L.; Springer, S.; Kinde, I.; Silliman, N.; Tacey, M.; Wong, H.L.; Christie, M.; et al. Circulating tumor DNA analysis detects minimal residual disease and predicts recurrence in patients with stage II colon cancer. Sci. Transl. Med. 2016, 8, 346ra392. [Google Scholar] [CrossRef] [Green Version]
  228. Chaudhuri, A.A.; Chabon, J.J.; Lovejoy, A.F.; Newman, A.M.; Stehr, H.; Azad, T.D.; Khodadoust, M.S.; Esfahani, M.S.; Liu, C.L.; Zhou, L.; et al. Early Detection of Molecular Residual Disease in Localized Lung Cancer by Circulating Tumor DNA Profiling. Cancer Discov. 2017, 7, 1394–1403. [Google Scholar] [CrossRef] [Green Version]
  229. Garcia-Murillas, I.; Chopra, N.; Comino-Méndez, I.; Beaney, M.; Tovey, H.; Cutts, R.J.; Swift, C.; Kriplani, D.; Afentakis, M.; Hrebien, S.; et al. Assessment of Molecular Relapse Detection in Early-Stage Breast Cancer. JAMA Oncol. 2019, 5, 1473–1478. [Google Scholar] [CrossRef]
  230. Reinert, T.; Henriksen, T.V.; Christensen, E.; Sharma, S.; Salari, R.; Sethi, H.; Knudsen, M.; Nordentoft, I.; Wu, H.T.; Tin, A.S.; et al. Analysis of Plasma Cell-Free DNA by Ultradeep Sequencing in Patients with Stages I to III Colorectal Cancer. JAMA Oncol. 2019, 5, 1124–1131. [Google Scholar] [CrossRef] [Green Version]
  231. Lee, J.H.; Saw, R.P.; Thompson, J.F.; Lo, S.; Spillane, A.J.; Shannon, K.F.; Stretch, J.R.; Howle, J.; Menzies, A.M.; Carlino, M.S.; et al. Pre-operative ctDNA predicts survival in high-risk stage III cutaneous melanoma patients. Ann. Oncol. 2019, 30, 815–822. [Google Scholar] [CrossRef]
  232. Tan, L.; Sandhu, S.; Lee, R.J.; Li, J.; Callahan, J.; Ftouni, S.; Dhomen, N.; Middlehurst, P.; Wallace, A.; Raleigh, J.; et al. Prediction and monitoring of relapse in stage III melanoma using circulating tumor DNA. Ann. Oncol. J. Eur. Soc. Med. Oncol. 2019, 30, 804–814. [Google Scholar] [CrossRef] [PubMed]
  233. Lee, R.J.; Gremel, G.; Marshall, A.; Myers, K.A.; Fisher, N.; Dunn, J.A.; Dhomen, N.; Corrie, P.G.; Middleton, M.R.; Lorigan, P.; et al. Circulating tumor DNA predicts survival in patients with resected high-risk stage II/III melanoma. Ann. Oncol. 2018, 29, 490–496. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Genta, S.; Araujo, D.V.; Keshavarzi, S.; Muniz, T.P.; Kamil, Z.S.; Howarth, K.; Terrell, S.; Joad, A.; Ventura, R.; Covelli, A.; et al. Leveraging personalized circulating tumor DNA (ctDNA) for detection and monitoring of molecular residual disease in high-risk melanoma. J. Clin. Oncol. 2022, 40, 9579. [Google Scholar] [CrossRef]
  235. McEvoy, A.C.; Pereira, M.R.; Reid, A.; Pearce, R.; Cowell, L.; Al-Ogaili, Z.; Khattak, M.A.; Millward, M.; Meniawy, T.M.; Gray, E.S.; et al. Monitoring melanoma recurrence with circulating tumor DNA: A proof of concept from three case studies. Oncotarget 2019, 10, 113–122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Adam, G.; Rampasek, L.; Safikhani, Z.; Smirnov, P.; Haibe-Kains, B.; Goldenberg, A. Machine learning approaches to drug response prediction: Challenges and recent progress. NPJ Precis. Oncol. 2020, 4, 19. [Google Scholar] [CrossRef] [PubMed]
  237. Long, G.V.; Desai, K.; Tang, T.; Weber, J.S.; Dolfi, S.; Ritchings, C.; Huang, S.P.; Bolisetty, M.; Sausen, M.; Del Vecchio, M.; et al. 788O Association of pre-treatment ctDNA with disease recurrence and clinical and translational factors in patients with stage IIIB-D/IV melanoma treated with adjuvant immunotherapy (CheckMate 915). Ann. Oncol. 2022, 33, S904. [Google Scholar] [CrossRef]
  238. Gracie, L.; Pan, Y.; Atenafu, E.G.; Ward, D.G.; Teng, M.; Pallan, L.; Stevens, N.M.; Khoja, L. Circulating tumour DNA (ctDNA) in metastatic melanoma, a systematic review and meta-analysis. Eur. J. Cancer 2021, 158, 191–207. [Google Scholar] [CrossRef] [PubMed]
  239. Kaminska, P.; Buszka, K.; Zabel, M.; Nowicki, M.; Alix-Panabieres, C.; Budna-Tukan, J. Liquid Biopsy in Melanoma: Significance in Diagnostics, Prediction and Treatment Monitoring. Int. J. Mol. Sci. 2021, 22, 9714. [Google Scholar] [CrossRef]
  240. Lee, J.H.; Long, G.V.; Boyd, S.; Lo, S.; Menzies, A.M.; Tembe, V.; Guminski, A.; Jakrot, V.; Scolyer, R.A.; Mann, G.J.; et al. Circulating tumour DNA predicts response to anti-PD1 antibodies in metastatic melanoma. Ann. Oncol. 2017, 28, 1130–1136. [Google Scholar] [CrossRef]
  241. Bratman, S.V.; Yang, S.Y.C.; Iafolla, M.A.J.; Liu, Z.; Hansen, A.R.; Bedard, P.L.; Lheureux, S.; Spreafico, A.; Razak, A.A.; Shchegrova, S.; et al. Personalized circulating tumor DNA analysis as a predictive biomarker in solid tumor patients treated with pembrolizumab. Nat. Cancer 2020, 1, 873–881. [Google Scholar] [CrossRef]
  242. Marsavela, G.; Lee, J.; Calapre, L.; Wong, S.Q.; Pereira, M.R.; McEvoy, A.C.; Reid, A.L.; Robinson, C.; Warburton, L.; Abed, A.; et al. Circulating Tumor DNA Predicts Outcome from First-, but not Second-line Treatment and Identifies Melanoma Patients Who May Benefit from Combination Immunotherapy. Clin. Cancer Res. 2020, 26, 5926–5933. [Google Scholar] [CrossRef] [PubMed]
  243. Lee, J.H.; Long, G.V.; Menzies, A.M.; Lo, S.; Guminski, A.; Whitbourne, K.; Peranec, M.; Scolyer, R.; Kefford, R.F.; Rizos, H.; et al. Association Between Circulating Tumor DNA and Pseudoprogression in Patients with Metastatic Melanoma Treated with Anti-Programmed Cell Death 1 Antibodies. JAMA Oncol. 2018, 4, 717–721. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Wang, Z.; Duan, J.; Cai, S.; Han, M.; Dong, H.; Zhao, J.; Zhu, B.; Wang, S.; Zhuo, M.; Sun, J.; et al. Assessment of Blood Tumor Mutational Burden as a Potential Biomarker for Immunotherapy in Patients with Non-Small Cell Lung Cancer with Use of a Next-Generation Sequencing Cancer Gene Panel. JAMA Oncol. 2019, 5, 696–702. [Google Scholar] [CrossRef] [PubMed]
  245. Kim, E.S.; Velcheti, V.; Mekhail, T.; Yun, C.; Shagan, S.M.; Hu, S.; Chae, Y.K.; Leal, T.A.; Dowell, J.E.; Tsai, M.L.; et al. Blood-based tumor mutational burden as a biomarker for atezolizumab in non-small cell lung cancer: The phase 2 B-F1RST trial. Nat. Med. 2022, 28, 939–945. [Google Scholar] [CrossRef] [PubMed]
  246. Si, H.; Kuziora, M.; Quinn, K.J.; Helman, E.; Ye, J.; Liu, F.; Scheuring, U.; Peters, S.; Rizvi, N.A.; Brohawn, P.Z.; et al. A Blood-based Assay for Assessment of Tumor Mutational Burden in First-line Metastatic NSCLC Treatment: Results from the MYSTIC Study. Clin. Cancer Res. 2021, 27, 1631–1640. [Google Scholar] [CrossRef]
  247. Chen, X.; Fang, L.; Zhu, Y.; Bao, Z.; Wang, Q.; Liu, R.; Sun, W.; Du, H.; Lin, J.; Yu, B.; et al. Blood tumor mutation burden can predict the clinical response to immune checkpoint inhibitors in advanced non-small cell lung cancer patients. Cancer Immunol. Immunother. 2021, 70, 3513–3524. [Google Scholar] [CrossRef]
  248. Wang, Z.; Zhao, X.; Gao, C.; Gong, J.; Wang, X.; Gao, J.; Li, Z.; Wang, J.; Yang, B.; Wang, L.; et al. Plasma-based microsatellite instability detection strategy to guide immune checkpoint blockade treatment. J. Immunother. Cancer 2020, 8, e001297. [Google Scholar] [CrossRef]
  249. Georgiadis, A.; Durham, J.N.; Keefer, L.A.; Bartlett, B.R.; Zielonka, M.; Murphy, D.; White, J.R.; Lu, S.; Verner, E.L.; Ruan, F.; et al. Noninvasive Detection of Microsatellite Instability and High Tumor Mutation Burden in Cancer Patients Treated with PD-1 Blockade. Clin. Cancer Res. 2019, 25, 7024–7034. [Google Scholar] [CrossRef] [Green Version]
  250. Cabel, L.; Proudhon, C.; Romano, E.; Girard, N.; Lantz, O.; Stern, M.H.; Pierga, J.Y.; Bidard, F.C. Clinical potential of circulating tumour DNA in patients receiving anticancer immunotherapy. Nat. Rev. Clin. Oncol. 2018, 15, 639–650. [Google Scholar] [CrossRef]
  251. Snyder, A.; Morrissey, M.P.; Hellmann, M.D. Use of Circulating Tumor DNA for Cancer Immunotherapy. Clin. Cancer Res. 2019, 25, 6909–6915. [Google Scholar] [CrossRef] [Green Version]
  252. Li, L.; Wang, Y.; Shi, W.; Zhu, M.; Liu, Z.; Luo, N.; Zeng, Y.; He, Y. Serial ultra-deep sequencing of circulating tumor DNA reveals the clonal evolution in non-small cell lung cancer patients treated with anti-PD1 immunotherapy. Cancer Med. 2019, 8, 7669–7678. [Google Scholar] [CrossRef] [PubMed]
  253. Jin, Y.; Chen, D.L.; Wang, F.; Yang, C.P.; Chen, X.X.; You, J.Q.; Huang, J.S.; Shao, Y.; Zhu, D.Q.; Ouyang, Y.M.; et al. The predicting role of circulating tumor DNA landscape in gastric cancer patients treated with immune checkpoint inhibitors. Mol. Cancer 2020, 19, 154. [Google Scholar] [CrossRef] [PubMed]
  254. Takai, E.; Omata, W.; Totoki, Y.; Nakamura, H.; Shiba, S.; Takahashi, A.; Namikawa, K.; Mori, T.; Yamazaki, N.; Shibata, T.; et al. Clonal dynamics of circulating tumor DNA during immune checkpoint blockade therapy for melanoma. Cancer Sci. 2021, 112, 4748–4757. [Google Scholar] [CrossRef] [PubMed]
  255. Rogers, H.W.; Weinstock, M.A.; Feldman, S.R.; Coldiron, B.M. Incidence Estimate of Nonmelanoma Skin Cancer (Keratinocyte Carcinomas) in the U.S. Population, 2012. JAMA Derm. 2015, 151, 1081–1086. [Google Scholar] [CrossRef] [PubMed]
  256. Fania, L.; Didona, D.; Di Pietro, F.R.; Verkhovskaia, S.; Morese, R.; Paolino, G.; Donati, M.; Ricci, F.; Coco, V.; Ricci, F.; et al. Cutaneous Squamous Cell Carcinoma: From Pathophysiology to Novel Therapeutic Approaches. Biomedicines 2021, 9, 171. [Google Scholar] [CrossRef]
  257. Migden, M.R.; Rischin, D.; Schmults, C.D.; Guminski, A.; Hauschild, A.; Lewis, K.D.; Chung, C.H.; Hernandez-Aya, L.; Lim, A.M.; Chang, A.L.S.; et al. PD-1 Blockade with Cemiplimab in Advanced Cutaneous Squamous-Cell Carcinoma. N. Engl. J. Med. 2018, 379, 341–351. [Google Scholar] [CrossRef] [Green Version]
  258. Rischin, D.; Fury, M.G.; Lowy, I.; Stankevich, E.; Han, H.; Porceddu, S. A phase III, randomized, double-blind study of adjuvant cemiplimab versus placebo post-surgery and radiation therapy (RT) in patients (pts) with high-risk cutaneous squamous cell carcinoma (CSCC). J. Clin. Oncol. 2020, 38, TPS10084. [Google Scholar] [CrossRef]
  259. Grob, J.J.; Gonzalez, R.; Basset-Seguin, N.; Vornicova, O.; Schachter, J.; Joshi, A.; Meyer, N.; Grange, F.; Piulats, J.M.; Bauman, J.R.; et al. Pembrolizumab Monotherapy for Recurrent or Metastatic Cutaneous Squamous Cell Carcinoma: A Single-Arm Phase II Trial (KEYNOTE-629). J. Clin. Oncol. 2020, 38, 2916–2925. [Google Scholar] [CrossRef]
  260. Maubec, E.; Boubaya, M.; Petrow, P.; Beylot-Barry, M.; Basset-Seguin, N.; Deschamps, L.; Grob, J.J.; Dreno, B.; Scheer-Senyarich, I.; Bloch-Queyrat, C.; et al. Phase II Study of Pembrolizumab As First-Line, Single-Drug Therapy for Patients with Unresectable Cutaneous Squamous Cell Carcinomas. J. Clin. Oncol. 2020, 38, 3051–3061. [Google Scholar] [CrossRef]
  261. Ferrarotto, R.; Amit, M.; Nagarajan, P.; Rubin, M.L.; Yuan, Y.; Bell, D.; El-Naggar, A.K.; Johnson, J.M.; Morrison, W.H.; Rosenthal, D.I.; et al. Pilot Phase II Trial of Neoadjuvant Immunotherapy in Locoregionally Advanced, Resectable Cutaneous Squamous Cell Carcinoma of the Head and Neck. Clin. Cancer Res. 2021, 27, 4557–4565. [Google Scholar] [CrossRef]
  262. Becker, J.C.; Stang, A.; DeCaprio, J.A.; Cerroni, L.; Lebbe, C.; Veness, M.; Nghiem, P. Merkel cell carcinoma. Nat. Rev. Dis Prim. 2017, 3, 17077. [Google Scholar] [CrossRef]
  263. Harms, P.W.; Vats, P.; Verhaegen, M.E.; Robinson, D.R.; Wu, Y.M.; Dhanasekaran, S.M.; Palanisamy, N.; Siddiqui, J.; Cao, X.; Su, F.; et al. The Distinctive Mutational Spectra of Polyomavirus-Negative Merkel Cell Carcinoma. Cancer Res. 2015, 75, 3720–3727. [Google Scholar] [CrossRef] [Green Version]
  264. Goh, G.; Walradt, T.; Markarov, V.; Blom, A.; Riaz, N.; Doumani, R.; Stafstrom, K.; Moshiri, A.; Yelistratova, L.; Levinsohn, J.; et al. Mutational landscape of MCPyV-positive and MCPyV-negative Merkel cell carcinomas with implications for immunotherapy. Oncotarget 2016, 7, 3403–3415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Lyngaa, R.; Pedersen, N.W.; Schrama, D.; Thrue, C.A.; Ibrani, D.; Met, O.; Thor Straten, P.; Nghiem, P.; Becker, J.C.; Hadrup, S.R. T-cell responses to oncogenic merkel cell polyomavirus proteins distinguish patients with merkel cell carcinoma from healthy donors. Clin. Cancer Res. 2014, 20, 1768–1778. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  266. Knepper, T.C.; Montesion, M.; Russell, J.S.; Sokol, E.S.; Frampton, G.M.; Miller, V.A.; Albacker, L.A.; McLeod, H.L.; Eroglu, Z.; Khushalani, N.I.; et al. The Genomic Landscape of Merkel Cell Carcinoma and Clinicogenomic Biomarkers of Response to Immune Checkpoint Inhibitor Therapy. Clin. Cancer Res. 2019, 25, 5961–5971. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  267. Topalian, S.L.; Bhatia, S.; Amin, A.; Kudchadkar, R.R.; Sharfman, W.H.; Lebbe, C.; Delord, J.P.; Dunn, L.A.; Shinohara, M.M.; Kulikauskas, R.; et al. Neoadjuvant Nivolumab for Patients with Resectable Merkel Cell Carcinoma in the CheckMate 358 Trial. J. Clin. Oncol. 2020, 38, 2476–2487. [Google Scholar] [CrossRef]
  268. Spassova, I.; Ugurel, S.; Kubat, L.; Zimmer, L.; Terheyden, P.; Mohr, A.; Bjorn Andtback, H.; Villabona, L.; Leiter, U.; Eigentler, T.; et al. Clinical and molecular characteristics associated with response to therapeutic PD-1/PD-L1 inhibition in advanced Merkel cell carcinoma. J. Immunother. Cancer 2022, 10, e003198. [Google Scholar] [CrossRef]
  269. Chalmers, Z.R.; Connelly, C.F.; Fabrizio, D.; Gay, L.; Ali, S.M.; Ennis, R.; Schrock, A.; Campbell, B.; Shlien, A.; Chmielecki, J.; et al. Analysis of 100,000 human cancer genomes reveals the landscape of tumor mutational burden. Genome Med. 2017, 9, 34. [Google Scholar] [CrossRef] [Green Version]
  270. Dummer, R.; Ascierto, P.A.; Basset-Seguin, N.; Dreno, B.; Garbe, C.; Gutzmer, R.; Hauschild, A.; Krattinger, R.; Lear, J.T.; Malvehy, J.; et al. Sonidegib and vismodegib in the treatment of patients with locally advanced basal cell carcinoma: A joint expert opinion. J. Eur. Acad Derm. Venereol 2020, 34, 1944–1956. [Google Scholar] [CrossRef]
Figure 1. Timeline of immunotherapy agents’ approval for melanoma and non-melanoma skin cancer treatment. NMSC—non-melanoma skin cancer, cSCC—cutaneous squamous cell carcinoma, BCC—basal cell carcinoma, MCC—Merkel cell carcinoma, T-VEC—Talimogene laherparepvec; IFN-γ—interferon gamma; TCR—T cell receptor; CCR5—chemokine receptor 5; CXCL13—chemokine ligand 13; TLS—tertiary lymphoid structures; TAM—tumor-associated macrophages; MDSC—myeloid-derived suppressor cells; T regs—regulatory T cells; HGF—hepatocyte growth factor; IL6—interleukin 6; TGF-β—Transforming growth factor beta; purple colour for biomarkers: biomarkers of response; blue colour for biomarkers: biomarkers of resistance.
Figure 1. Timeline of immunotherapy agents’ approval for melanoma and non-melanoma skin cancer treatment. NMSC—non-melanoma skin cancer, cSCC—cutaneous squamous cell carcinoma, BCC—basal cell carcinoma, MCC—Merkel cell carcinoma, T-VEC—Talimogene laherparepvec; IFN-γ—interferon gamma; TCR—T cell receptor; CCR5—chemokine receptor 5; CXCL13—chemokine ligand 13; TLS—tertiary lymphoid structures; TAM—tumor-associated macrophages; MDSC—myeloid-derived suppressor cells; T regs—regulatory T cells; HGF—hepatocyte growth factor; IL6—interleukin 6; TGF-β—Transforming growth factor beta; purple colour for biomarkers: biomarkers of response; blue colour for biomarkers: biomarkers of resistance.
Ijms 24 01294 g001
Figure 2. Biomarkers of response/resistance to immunotherapy in melanoma. purple color: markers of response; blue color: markers of resistance.
Figure 2. Biomarkers of response/resistance to immunotherapy in melanoma. purple color: markers of response; blue color: markers of resistance.
Ijms 24 01294 g002
Table 1. Ongoing clinical trials with immunotherapy in advanced cutaneous melanoma (as assessed on 13 November 2022). * Denotes studies that also include patients with tumor types other than melanoma.
Table 1. Ongoing clinical trials with immunotherapy in advanced cutaneous melanoma (as assessed on 13 November 2022). * Denotes studies that also include patients with tumor types other than melanoma.
Type of InterventionTargetsAgentsNumber of PatientsPhaseNCT Number
Novel Immune Pathways
Anti-PD-1 plus anti-TIGIT (IV) or anti-PD-1 + Coxsackievirus (IT) or anti-PD-1 + ILT-4 or anti-PD-1/anti-LAG-3 or anti-PD-1 + retinoidPembrolizumab + vibostolimab or pembrolizumab + gebasaxturev or pembrolizumab + MK-4830 or favezelimab or pembrolizumab + all-trans retinoic acid (ATRA)901/2NCT04303169
Anti-CTLA-4/anti-PD-1 +/− anti-interleukin 8 antibodyIpilimumab/Nivolumab +/− BMS-986253 (HuMax-IL8)3721/2NCT03400332
IL-2Rβ/γ-selective IL-2 variant +/− anti-PD-1 +/− TLR7/8 agonistPembrolizumab +/− TransCon IL-2 β/γ +/− TransCon TLR7/8 agonist317 *1/2NCT05081609
IL-2 superkine +/− ICIMDNA11 +/− ICI100 *1/2NCT05086692
T cell engager +/− anti-PD-1 +/− anti-CTLA-4 Tebentafusp +/− durvalumab +/− tremelimumab3121/2NCT02535078
IL-12 Fc-fusion protein +/− anti-PD-1BMS-986415 +/− Nivolumab473 *1/2NCT04423029
CD40 agonist +/− anti-PD-1APX005M + pembrolizumab411/2NCT02706353
CXCR1/2 inhibitor + anti-PD-1SX-682 + pembrolizumab771NCT03161431
STING agonist + anti-PD-L1 SB 11285 + atezolizumab110 *1NCT04096638
Anti-TIGIT + anti-PD-1AB154 + AB122262NCT05130177
Interleukin (IL-)15 and IL-15 receptor alpha + anti-PD-1NIZ985 +/− spartalizumab110 *1NCT04261439
mAb Specific to B-and T-Lymphocyte Attenuator (BTLA)JS004156 *1NCT04773951
PD-L1/IDO peptide vaccine + anti-PD-1PD-L1/IDO peptide vaccine + nivolumab501/2NCT03047928
Anti-CCR8 +/− anti-PD-1BAY3375968270 *1NCT05537740
TGFβ1 inhibitor +/− anti-PD-L1SRK-181 +/− anti-PD-L1200 *1NCT04291079
Anti-PD-1 inhibitor/OX40 agonist SL-279252 (PD1-Fc-OX40L)87 *1NCT03894618
Granulocyte Macrophage-colony stimulating factor + anti-PD-1Sargramostim + pembrolizumab302NCT04703426
Adoptive cell therapy
PD-1 knock out TILs IOV-400153 *1/2NCT05361174
Adoptive autologous cellular therapy + HDAC + DNMT inhibitorsMAGE-C2/HLA-A2 TCR T cells + azacytidine + valproic acid201/2NCT04729543
CD20 CAR transduced T cellsMB-CART20.1151NCT03893019
Autologous alpha-type-1 polarized dendritic cells (alphaDC1)/TBVA vaccine + cytokine modulating regimenTBVA vaccine intatolimod + IFN-alpha2b + celecoxib242NCT04093323
Autologous neoantigen-specific T-cell productNEO-PTC-01521NCT04625205
Oncolytic viral therapy
Rhinovirus targeting poliovirus receptor CD155 (IT) +/− anti-PD-1Lerapolturev (IT) +/− anti-PD-1562NCT04577807
Oncolytic vesicular stomatitis virus expressing human IFNβ (IT or IV) +/− anti-PD-1 +/− anti-CTLA-4Voyager V1 +/− cemiplimab +/− ipilimumab152 *2NCT04291105
Anti-PD-1 + oncolytic HSV type virus expressing anti-PD-1 antibody and IL-12 (IT)MVR-T3011 (IT) +/− pembrolizumab101/2NCT04370587
Herpes simplex virus type 2 strain HG52 +/− anti-PD-1OH2 +/− nivolumab30 *1/2NCT04386967
Adenovirus with TMA-CD40L and 4-1BBL transgenes + anti-PD-L1LOAd703 + atezolizumab351/2NCT04123470
Great ape Adenoviral (GAd)/Modified Vaccinia Ankara (MVA) boosts with personalised patient neoantigens + anti-PD-1NOUS-PEV + pembrolizumab34 *1NCT04990479
Anti-PD-1 + oncolytic HSV type 1 virus (IT)Pembrolizumab + RP1 300 *2NCT03767348
Anti-PD-1 plus vaccinia virus encoding transgenes for Flt3 ligand, anti-CTLA-4 antibody and IL-12 cytokine, (IT or IV) Pembrolizumab + TBio-6517 (IV or IT)138 *1/2NCT04301011
Targeted treatment plus immunotherapy
PI3K inhibitor + anti-PD-1Duvelisib + nivolumab421/2NCT04688658
ETBR inhibitor +/− anti-PD-1 ENB003 +/− pembrolizumab1301/2NCT04205227
Anti-VEGF + anti-PD-L1Bevacizumab + atezolizumab302NCT04356729
Integrin alpha-V/beta-8 AntagonistPF-069404341221NCT04152018
p38 inhibitor + anti-PD-1 or anti-CTLA-4/anti-PD-1ARRY-614 + nivolumab or nivolumab/ipilimumab144 *1/2NCT04074967
Hedgehog pathway inhibitor + anti-PD-1Sonidegib + Pembrolizumab45 *1NCT04007744
Wnt/β-catenin inhibitor +/− anti-PD-1E7386 +/− pembrolizumab108 *1/2NCT05091346
PARP inhibition + anti-PD-1Olaparib + pembrolizumab412NCT04633902
“Velcrin” that triggers the formation of a complex of two proteins called SLFN12 and PDE3A BAY2666605891NCT04809805
Heparanase inhibitor (TAMs inhibitor) + anti-PD-1 + metronomic chemotherapyPixatimod (PG545) + Nivolumab + cyclophosphamide612NCT05061017
Anti-CTLA-4/anti-PD-1 +/− glutamate modulatorIpilimumab/Nivolumab +/− troriluzole1082NCT04899921
MDM2 inhibitor + anti-PD-1APG-115 + pembrolizumab2241/2NCT03611868
Multi-tyrosine kinase inhibitor + anti-PD-1Cabozantinib + pembrolizumab391/2NCT03957551
BTK inhibitor + anti-PD-1Ibrutinib + pembrolizumab231NCT03021460
ATR inhibitor +/− anti-PD-L1Ceralasertib +/− durvalumab1952NCT05061134
Anti-PD-1 plus NLRP3 inhibitorPembrolizumab + Dapansutrile261/2NCT04971499
Vaccine-based treatment
Melanoma HLA-restricted peptides vaccine +/− anti-CD276MHP +/− CDX-1127301/2NCT03617328
DNA vaccine encoding Tyrosinase-Related Protein 2 (TRP2) and gp100SCIB1 DNA vaccine872NCT04079166
mRNA-encoded with tumor-associated antigens vaccine + anti-PD-1BNT111 + cemiplimab1802NCT04526899
Antibody drug conjugates
ADC against CD228 SGN-CD228A275 *1NCT05571839
ROR2-targeted antibody drug conjugate (CAB-ROR2-ADC) +/− anti-PD-1BA3021 +/− pembrolizumab420 *2NCT03504488
Table 2. Ongoing clinical trial testing immunotherapy strategies in patients with uveal and mucosal melanoma (as assessed on 13 November 2022).
Table 2. Ongoing clinical trial testing immunotherapy strategies in patients with uveal and mucosal melanoma (as assessed on 13 November 2022).
Target PopulationType of TreatmentInterventionNumber of PatientsPhaseNCT Number
Uveal MelanomaLocal treatment + anti-PD-1 and anti CTLA-4Yttrium 90 + Ipilimumab + Nivolumab261/2NCT02913417
Melphalan percutaneous hepatic perfusion + Ipilimumab + Nivolumab831/2NCT04283890
Immunoembolization +IPI/NIVO142NCT03472586
Ipilimumab + Nivolumab + Tumor Treating Fields101NCT05004025
Stereotactic RT + Ipilimumab + Nivolumab402NCT05077280
Anti-PD-1 + Anti-LAG-3Nivolumab + Relatlimab272NCT04552223
Anti-PD-1 and anti CTLA-4 + TLR9 agonistsNivolumab + Ipilimumab + intrahepatic SD-101801NCT04935229
Anti-PD-1 and anti CTLA-4 + anti-arginineADI-PEG20 + Ipilimumab + Nivolumab91NCT03922880
Anti-CTLA-4 + cell therapyCD8+ SLC45A2-Specific T Lymphocytes + Cyclophosphamide, Aldesleukin, and Ipilimumab301NCT03068624
Cell therapyTumor Infiltrating Lymphocytes472NCT03467516
101NCT05607095
Dendritic cells + tumor RNA2003NCT01983748
Anti-PD-1 + epigenetic modulatorsPembrolizimab + Etinostat292NCT02697630
Anti-PD-1 + Anti-VEGFPembrolizumab + Lenvatinib542NCT05282901
302NCT05308901
Anti-PD-1 + Anti-DDRPembrolizumab + Olaparib372NCT 05524935
Mucosal MelanomaAnti-PD-1 + local treatmentAdjuvant Pembrolizumab + hypofractionated RT162NCT04318717
Anti-PD-1 + chemotherapy + local treatmentAdjuvant Toripalimab + chemotherapy or RT452NCT04879654
Anti-PD-1 + Anti-VEGFNeoadjuvant Pembrolizumab + Lenvatinib442NCT05545969
262NCT04622566
SHR-1210 + Apatinib402NCT03986515
Adjuvant Nivolumab +/− Cabozantinib992NCT05111574
Nivolumab + Axitinib202NCT05384496
Toripalib + Axitinib302NCT04180995
992NCT03941795
Anti-PD-1 + anti-TGFβPembrolizumab + Vactosertib302NCT05436990
Anti-PD-1 + anti CD40 + chemotherapyYH003 + Pembrolizumab + Paclitaxel432NCT05420324
Anti-PD-1 + chemotherapy Nivolumab + Decitabine/Cedazuridine301/2NCT05089370
IL-2 agonistsALKS 4230 1102NCT04830124
Table 3. Ongoing clinical trials with different types of ICI in NMSC (as assessed on 13 November 2022).
Table 3. Ongoing clinical trials with different types of ICI in NMSC (as assessed on 13 November 2022).
Study TitleTumor TypeInterventionPhaseNCT Number
Neo-adjuvant Nivolumab or Nivolumab With Ipilimumab in Advanced Cutaneous Squamous Cell Carcinoma Prior to SurgerycSCCIpilimumab Nivolumab2NCT04620200
Neoadjuvant Plus Adjuvant Treatment With Cemiplimab in Cutaneous Squamous Cell CarcinomacSCCCemiplimab2NCT04632433
Neoadjuvant Pembrolizumab in Cutaneous Squamous Cell CarcinomacSCCPembrolizumab2NCT05025813
Atezolizumab Before Surgery for the Treatment of Regionally Metastatic Head and Neck Squamous Cell Cancer With an Unknown or Historic Primary SitecSCCAtezolizumab2NCT05110781
Avelumab With or Without Cetuximab in Treating Patients With Advanced Skin Squamous Cell CancercSCCAvelumab
Cetuximab
2NCT03944941
Phase II Study of Peptide Receptor Radionuclide Therapy in Combination With Immunotherapy for Patients With Merkel Cell CancerMCCPembrolizumab, Lutetium Lu 177 dotatate2NCT05583708
Neoadjuvant Lenvatinib Plus Pembrolizumab in Merkel Cell CarcinomaMCCPembrolizumab, Lenvatinib2NCT04869137
Navtemadlin (KRT-232) With or Without Anti-PD-1/Anti-PD-L1 for the Treatment of Patients With Merkel Cell CarcinomaMCCKRT-232, Avelumab 2NCT03787602
Adjuvant Avelumab in Merkel Cell CancerMCCAvelumab, placebo2NCT03271372
Testing Pembrolizumab Versus Observation in Patients With Merkel Cell Carcinoma After Surgery, STAMP StudyMCCPembrolizumab, placebo3NCT03712605
Intralesional Cemiplimab for Patients With Cutaneous Squamous Cell Carcinoma or Basal Cell CarcinomacSCC, BCCCemiplimab intralesional1NCT03889912
Nivolumab Alone or Plus Relatlimab or Ipilimumab for Patients With Locally-Advanced Unresectable or Metastatic Basal Cell CarcinomaBCCNivolumab, Relatlimab2NCT03521830
Neoadjuvant-Adjuvant Pembrolizumab in Resectable Advanced Basal Cell Carcinoma of H&NBCCPembrolizumab1NCT04323202
Anti-PD1-antibody and Pulsed HHI for Advanced BCCBCCCemiplimab, Sonidegib2NCT04679480
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

Spiliopoulou, P.; Vornicova, O.; Genta, S.; Spreafico, A. Shaping the Future of Immunotherapy Targets and Biomarkers in Melanoma and Non-Melanoma Cutaneous Cancers. Int. J. Mol. Sci. 2023, 24, 1294. https://doi.org/10.3390/ijms24021294

AMA Style

Spiliopoulou P, Vornicova O, Genta S, Spreafico A. Shaping the Future of Immunotherapy Targets and Biomarkers in Melanoma and Non-Melanoma Cutaneous Cancers. International Journal of Molecular Sciences. 2023; 24(2):1294. https://doi.org/10.3390/ijms24021294

Chicago/Turabian Style

Spiliopoulou, Pavlina, Olga Vornicova, Sofia Genta, and Anna Spreafico. 2023. "Shaping the Future of Immunotherapy Targets and Biomarkers in Melanoma and Non-Melanoma Cutaneous Cancers" International Journal of Molecular Sciences 24, no. 2: 1294. https://doi.org/10.3390/ijms24021294

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