Skip to main content

Recent developments in immunotherapy for gastrointestinal tract cancers

Abstract

The past few decades have witnessed the rise of immunotherapy for Gastrointestinal (GI) tract cancers. The role of immune checkpoint inhibitors (ICIs), particularly programmed death protein 1 (PD-1) and PD ligand-1 antibodies, has become increasingly pivotal in the treatment of advanced and perioperative GI tract cancers. Currently, anti-PD-1 plus chemotherapy is considered as first-line regimen for unselected advanced gastric/gastroesophageal junction adenocarcinoma (G/GEJC), mismatch repair deficient (dMMR)/microsatellite instability-high (MSI-H) colorectal cancer (CRC), and advanced esophageal cancer (EC). In addition, the encouraging performance of claudin18.2-redirected chimeric antigen receptor T-cell (CAR-T) therapy in later-line GI tract cancers brings new hope for cell therapy in solid tumour treatment. Nevertheless, immunotherapy for GI tumour remains yet precise, and researchers are dedicated to further maximising and optimising the efficacy. This review summarises the important research, latest progress, and future directions of immunotherapy for GI tract cancers including EC, G/GEJC, and CRC.

Background

Worldwide, gastrointestinal (GI) tract cancers are the most prevalent malignancies with high morbidity and mortality [1]. According to the Global Cancer Observatory 2022 estimates, East Asia registered 1,469,225 new patients with GI tract cancers in 2022, accounting for 43.1% of the global incidence of GI tract cancers and 837,360 deaths, amounting to 41.7% of cancer-related deaths of GI tract cancers. Notably, the incidence of GI tract cancers is sharply increasing among young adults [2]. However, the etiological factors for GI tract cancers encompass infectious agents (such as HP, EBV infections), genetic factors (such as CDH1 mutations), and environmental factors (such as poor dietary habits) which have not been completely eliminated. In certain high-risk regions, routine endoscopic screening has not yet been universally implemented, severely affecting the health of populations and imposing economic burdens globally. Immunotherapy, one of the most recently developed therapies based on the theory of cancer immunoediting which indicated the crucial role of immune escape in tumour development and growth, has wrought a profound revolution in cancer treatment [3]. Immunotherapy was defined as therapeutic methods restoring the normal anti-tumour immune response, restarting the tumour-immune and further eliminating tumour cells [4]. Immune checkpoint inhibitors (ICIs), cancer vaccines, cell therapy, oncolytic virus (OVs) were in the category of immunotherapy.

Recent clinical trials, especially of ICIs targeting programmed death protein 1 (PD-1) and PD ligand-1(PD-L1), have shown their noteworthy efficacy in GI tract cancers and have contributed to a paradigm shift in treatment principles. Notwithstanding the anti-PD-1/PD-L1 treatment patterns of second- or later-line GI tract cancers, the success of Chekmate-649 [5], ORIENT-16 [6], Keynote-177 [7], and Keynote-590 [8] enabled anti-PD-1 antibodies plus chemotherapy be involved into the first-line mainstay treatment for esophagogastric junction (EGJ)/gastric cancer (GC), mismatch repair-deficient (dMMR)/microsatellite instability-high (MSI-H) colorectal cancer (CRC) and advanced oesophageal cancer (EC). Besides, Keynote-811 promoted the combination of trastuzumab, pembrolizumab, and chemotherapy as the standard first-line treatment for human epidermal growth factor receptor 2 (HER-2) positive GC [9]. Moreover, the long-lasting survival benefit of above large-scale clinical trials is under observation. The remarkable improvement in response to first-line therapy has boosted a surge of clinical trials focused on perioperative anti-PD-1/PD-L1 therapy in GI tract cancers [10,11,12,13,14,15,16,17,18,19,20,21,22,23]. Additionally, chimeric antigen receptor T-cell (CAR-T) therapy, exerts potent anti-tumour efficacy against GI cancers. According to phase I results in claudin18.2-redirected CAR-T cells (CT041) in patients with GI cancers who failed to respond after at least two prior lines of therapy, CT041 reached an objective response rate (ORR) and Disease control rate (DCR) of 48.6% and 73.0%, respectively. What’s more, the 6-month response rate was 44.8% [24]. The initial triumph of the CT041 phase I clinical trial was a milestone marking the entry of CAR-T therapy into GI tumour treatment. Therefore, an increasing number of clinical trials and translational studies concerning novel immunotherapies, combination treatments, and therapy modes in distinct lines for GI are being performed.

Despite the significant therapeutic progress in immunotherapy, many challenges persist: the biomarkers with definite cut-off values remain unknown; the majority of patients with GI tract cancers still suffer from primary or secondary resistance; the strategies for patients with specific subtypes have not been identified; and the management of adverse effects is yet to be standardized. Here, we focused on the EC, G/GEJC, and CRC and systematically summarised the pivotal clinical trials and discussed the latest advances in the management of these cancers. Furthermore, we outlined the progress and challenges in the realisation of precision immunotherapy, including the exploration of biomarkers, investigation of resistance mechanisms, and strategies to optimise the development and efficacy of novel immunotherapies.

Current immunotherapy landscape

Oesophageal cancer

History and current situation of immunotherapy for EC

Approximately 70% of patients with EC are diagnosed at an advanced stage, with a median overall survival (OS) of 7–13 months under standard chemotherapy [25, 26] and a 5-year survival rate of 15–20% [27].

The histological subtypes of EC vary significantly across different regions. Oesophageal squamous cell carcinoma (ESCC) accounted for 85% of all EC worldwide, predominating in Eastern Europe and Asia. In contrast, oesophageal adenocarcinoma (EAC) comprised about 14% of cases and is more prevalent in Western Europe and North America [28]. EAC typically involves the lower third of the oesophagus and the GEJ, associated with Barrett’s oesophagus, history of gastroesophageal reflux disease, obesity, and tobacco usage, sharing molecular characteristics similar to those in GC [27,28,29,30]. Conversely, ESCC generally affects the upper two-thirds of the oesophagus and is linked with both smoking and alcohol consumption, exhibiting molecular features more resemble those observed in head and neck squamous cell carcinoma. Despite the advances in treatment, there remains a lack of phase III clinical trials focusing exclusively on immunotherapy for EAC. Currently, the efficacy data for immunotherapy in EAC is primarily derived from subgroup analyses of related studies. Previous studies have compared second-line single-agent immunotherapy with chemotherapy and reported positive outcomes, preliminarily establishing the role of immunotherapy in ESCC [31]. Immunotherapy has therefore advanced towards first-line treatment (Table 1).

Table 1 Summary of key clinical trials of immunotherapy in EC

To date, data from several phase III clinical trials, including KEYNOTE-590 [8], ESCORT-1st [26], CheckMate-648 [45], ORIENT-15 [37], JUPITER-06 [38], RATIONALE-306 [39], GEMSTONE-304 [46] and ESCORT-RWS [47] demonstrated that the first-line combination of chemotherapy and PD-1/PD-L1 blockade (chemo + anti-PD-1/PD-L1) could extend the OS of patients with ESCC from less than 12 months with chemotherapy alone to approximately 16 months. These findings have solidified the premier role of immuno-chemotherapy as the first-line treatment for advanced ESCC. And subsequently led to the exploration of immunotherapy for locally advanced diseases. In terms of adjuvant treatment, the phase III study CheckMate-577 [43] revealed significant benefits for patients with EC or GEJC (EAC accounting for 71%) underwent neoadjuvant chemoradiotherapy (nCRT) followed adjuvant nivolumab therapy, showing improved disease-free survival (DFS) (22.4 months vs. 11.0 months, HR 0.69, P < 0.001). Currently, most clinical trials of different neoadjuvant immunotherapy combined with chemotherapy (CT) or chemoradiotherapy (CRT) regimens were phase I or II studies, which demonstrated favourable pathologic complete response (pCR) rates ranging from 17 to 56% [48,49,50,51,52,53,54,55,56,57,58,59]. ESCORT-NEO [44] was the first phase III study to evaluate the role of neoadjuvant immunotherapy for resectable locally advanced ESCC (LA-ESCC) demonstrated that neoadjuvant camrelizumab plus chemotherapy can significantly improve pCR rate from 4.7 to 28% (P < 0.0001) compared to chemotherapy alone. Moreover, the phase II study NICE [60] recently updated its 2-year follow-up data, showing that 37.3% of patients experienced disease recurrence with distant metastasis accounting for 27.4% of these cases, being the predominant recurrence pattern. The 2-year OS and progression-free survival (PFS) rates were 78.1% and 67.9%, respectively. Although these trials underline the short-term benefits of neoadjuvant immunotherapy plus chemotherapy, further data are essential for evaluating long-term survival outcomes. Accordingly, National Comprehensive Cancer Network (NCCN) guidelines recommend multiple immunotherapy combinations with chemotherapy as the first-line treatment for ESCC, with nivolumab specifically endorsed for adjuvant treatment.

Challenges in immunotherapy for ESCC

Perioperative immunotherapy necessitates comprehensive investigation

Despite ongoing research into various regimens, there are currently no standard guidelines for neoadjuvant immunotherapy. Based on CROSS [61] and NEOCRTEC5010 [62], nCRT could improve survival over surgery alone among LA-ESCC patients and increase pCR rate to 29.0–43.2%. Further, combining nCRT with PD-1 blockade, as demonstrated in the phase II PALACE-1 [59] and NEOCRTEC1901 [63] studies, can increase the pCR rate to 50.0–55.6%. Neoadjuvant PD-1 blockade with chemotherapy also showed a significant increase in pCR rates over chemotherapy alone [44]. However, whether these higher pCR rates contribute to prolonged survival remains unclear and requires further validation in phase III randomized controlled trials (RCTs). Moreover, there is a notable absence of comparative studies between immunotherapy combined with nCRT versus nCT alone. Additionally, the role of neoadjuvant radiotherapy in the era of immunotherapy requires further clarification. Currently, there was only trials compared the nCRT and nCT. A Japanese study [64] showed that nCRT yielded superior OS compared to nCT in patients undergone R0 resection (mOS: not reached vs. 20.2months, P = 0.028).In contrast, a Chinese study [65] indicated that nCRT did not significantly improve OS over nCT (HR 0.82, 95%CI 0.58–1.18, P = 0.28), which may attribute to the inconsistency to differences in T stage and radiotherapy dosage among the study populations. Considering that combined radiotherapy may not substantially improve efficacy and that treatment-related deaths have been reported in studies including PALACE-1 [59], NEOCRTEC1901 [63] and EC-CRT-001 [66], the treatment regimens of immunotherapy combined with nCRT require further optimisation. For example, researchers are exploring whether reducing the dose of radiation or chemotherapy could achieve effective outcomes with reduced toxicity. As phase I SCALE-1 [67] study presented at the 2022 ASCO indicated, a short course of nCRT combined with toripalimab exhibited promising efficacy and tolerability in LA-ESCC.

To improve the effectiveness of the above-mentioned treatments, several promising directions are being explored: (1) Optimising combination chemotherapy regimens. Different chemotherapy agents affect tumour immune microenvironment (TIME) differently. For example, a meta-analysis, showed that the combination of taxane plus platinum (TP) with ICI exert higher survival rates compared to fluorouracil plus platinum [68]. Studies showed that taxanes could induce immunogenic death of tumour cells and activate TIME [69] making the taxane-based regimen a frequent choice in ongoing studies. Taxanes such as albumin-bound paclitaxel and paclitaxel are commonly used in EC clinical trials. The ESCORT-NEO [44] study compared the differences between albumin-bound paclitaxel and paclitaxel: group A (camrelizumab, albumin-bound paclitaxel, and cisplatin) and B (camrelizumab, paclitaxel, and cisplatin). Results showed that group A exhibited a higher pCR rate of 28.0%, compared to 15.4% in group B. This superior outcome in group A may be attributed to the "selective tumour local enrichment" effect of albumin-bound paclitaxel, which allows higher concentration in tumour tissues, potentially minimizing immune system damage. Additionally, the application of albumin-bound paclitaxel avoids the negative immune regulatory effects caused by glucocorticoids. (2) Optimising the sequence of administering chemotherapy and immunotherapy. Delaying PD-1 blockade until 3 days after administrating chemotherapy, as reported in a phase II study [51] could possibly enhance pCR rates by allowing chemotherapeutic agents to clear from the body before receiving anti-PD-1 agents, thus sparing T cells and maximizing cancer cell eradication. Thus, it is essential to explore the sequence of regimens in combination strategies by RCTs. (3) Determining the optimal duration of neoadjuvant therapy. The optimal mumber of neoadjuvant immunotherapy cycles remains uncertain. A prospective study investigating the combination of nCT and camrelizumab for the treatment of ESCC revealed that the pCR rates within the four-cycle group (50.0%, 3/6) were not significantly superior to the two-cycle group (46.7%, 7/15) [70]. Excessively prolonged treatment may result in missing the optimal surgical timing, whereas insufficient treatment could result in inadequate tumour shrinkage. Thus, it is essential to conduct prospective RCTs to determine the most effective treatment cycle and customise the number of treatment cycles for each patient. (4) Identifying patients who would most benefit from immunotherapy. Notably, the 2-year OS rate for patients treated with neoadjuvant adebrelimab is 92% [71], higher than that of immunotherapy plus nCT. Thus, it is also vital to identify whether it is enough to receive neoadjuvant mono-immunotherapy for patients with specific features.

Efficacy-enhancing strategy remains unclear

At present, several first-line combined immunotherapy treatments have shown promising outcomes, yet they failed to meet clinical demands. The challenges of improving the efficacy persist. In recent years, the number of clinical studies related to immunotherapy combination therapy has increased rapidly. CheckMate-648 [36] showed that nivolumab plus ipilimumab extended OS compared to chemotherapy alone, without new safety concerns identified, providing new chance for ESCC patients with chemotherapy intolerance. Additionally, the feasibility of combining PD-1/PD-L1 inhibitors with TIGIT monoclonal antibody has also been explored. The phase I study GO30103 [72] and phase II study AdvanTIG-203 [73] initially assessed the safety and efficacy of this combination. Furthermore, the phase III study SKYSRAPER-08 [42] compared the efficacy of tiragolumab (tira) + atezolizumab (atezo) + chemotherapy (CT) and placebo (pts) + CT, suggesting significantly improved PFS (6.2 months vs. 5.4 months) and OS (15.7 months vs. 11.1 months). The MORPHEUS-EC study [74] included the comparison of mono-immunotherapy (tira + atezo + CT vs. atezo + CT vs. CT alone) and showed that the OS in three groups were 16, 13.1, and 9.9 months, respectively. Although the combination of PD-1/PD-L1 blockade and TIGIT mAb did not demonstrate a clear numerical advantage over the 16-month OS of the first-line immunotherapy, there are notable differences in the efficacy and prognosis between monotherapy and combination immunotherapy in the phase Ib/II MORPHEUS-EC study [74]. Therefore, blockade of PD-1/PD-L1 combined with TIGIT mAb presents a promising strategy for ESCC. Additionally, given the unique anatomical site of EC, investigation into first-line systemic therapy combined with radiotherapy is warranted. Both the ESO-Shanghai 13 study [75] and the 2023 ESMO 1576P study [76] showed improvements in prognosis. More ongoing clinical studies on immunotherapy for esophageal cancer have been summarized in Table 2.

Table 2 Summary of ongoing key phase III clinical trials of immunotherapy in EC

In summary, immunotherapy strategies of EC vary by tumour sites. The management of precise perioperativeimmunotherapy, the strategies to improve efficacy of diverse tumour sites, potential beneficiary population features in perioperative and advanced-stage settings are important directions for future development.

Gastric and gastroesophageal junction cancer

History and current situation of immunotherapy for G/GEJC

Despite the continuous advancement and optimisation of chemotherapy regimens for advanced G/GEJC, the efficacy of first-line chemotherapy for advanced G/GEJC remains poor, with an OS of no longer than 12 months [77]. Immunotherapy has achieved satisfactory results in the treatment of advanced G/GEJC (Table 3), breaking through the long-standing treatment bottleneck of short survival with traditional chemotherapy.

Table 3 Summary of key clinical trials of immunotherapy in G/GEJC

Immunotherapy initially began after the reporting of positive results from the ATTRACTION-2 [79] and KEYNOTE-059 [95] studies, which suggested that the single-agent immunotherapy in the later-line (≥ 3) treatment of advanced G/GEJC could significantly improve ORR and OS. Despite the KEYNOTE-061 [80] in the second-line treatment of advanced G/GEJC were failed to meet the primary endpoint, however the post-hoc analysis found pembrolizumab in PD-L1 CPS ≥ 10 patients had a better outcome than chemotherapy group, suggesting a potential benefit from immunotherapy. KEYNOTE-062 [96] initially explored the role of first-line immunotherapy in (HER-2-negative G/GEJC across three cohorts, including pembrolizumab monotherapy, pembrolizumab combined with chemotherapy and chemotherapy alone, with primary endpoints being OS and PFS in patients with PD-L1 CPS ≥ 1 or ≥ 10. Despite failing to meet its primary endpoint, an OS benefit was observed for pembrolizumab monotherapy over chemotherapy in PD-L1 CPS ≥ 10 patients (17. 4 months vs 10. 8 months; HR 0. 69; 95%CI 0. 49–0. 97). However, this finding was not statistically tested. The study's shortcomings are presumably attributed to its complex statistical design and the variability in chemotherapy regimens. Concurrently, the CheckMate-649 study [5] of first-line treatment using nivolumab for HER-2-negative G/GEJC enrolled 2687 patients across three cohorts, focusing on OS and PFS in the PD-L1 CPS ≥ 5 patients. This study demonstrated that nivolumab combined with chemotherapy increased mOS by 3.3months and decreased the risk of death by 29% compared to chemotherapy alone in the CPS ≥ 5 patients, with similar trends across the entire cohort. Subsequently, ORIENT-16 [6], KEYNOTE-859 [82], RATIONALE-305 [97], and GEMSTONE-303 [85] showed that combination chemotherapy with sindilizumab, pembrolizumab, tirelizumab, and suglizumab (PD-L1 mAb) significantly increased mOS to 13 ~ 15 months in first-line treatment for HER-2-negative G/GEJC, especially in patients with high PD-L1 expression, where mOS could reach 15 ~ 18 months, and ORR was approximately 60%. Notably, as a biomarker-driven clinical trials, GEMSTONE-303 [85] specifically enrolled individuals with PD-L1 expression ≥ 5%, thus resulting in relatively high efficacy and survival benefits.

Based on these promising results, PD-1/PD-L1 mAb combined with platinum-containing chemotherapy has become the standard recommended regimen for the first-line treatment of G/GEJC, as incorporated into NCCN guidelines. Besides, it is noteworthy that long-term follow-up data from Checkmate-649 revealed a marked survival advantage in the Chinese population when compared to the global average, particularly notable in patients with a CPS ≥ 5, where the 3-year OS rates are 31% [98] and 17% [99] respectively. This disparity likely stems from genetic and lifestyle differences between Eastern and Western populations. Additionally, molecular characteristics, the TME, and health economics concerns should also be considered. The shared dining customs among Chinese may increase the transmission of Helicobacter pylori (HP), a known precursor for G/GEJC [100, 101]. A comprehensive meta-analysis involving 11 studies showed that G/GEJC cases with HP positivity are associated with higher PD-L1 expression levels, which correlates to better responses to immunotherapy [102, 103]. A study by Jia et al. [104] demonstrated that patients with HP-positive, EBV-negative, and MSS-positive G/GEJC achieved significantly longer immune-related PFS (irPFS) (6.97 months vs 5.03 months, HR 0.76, p < 0.001) and showed a tendency toward four months extended irOS in comparison to those in the HP-negative group (18.3 months vs 14.2 months, p = 0.105). Therefore, considering the characteristics of GC populations in different regions, the high spatiotemporal heterogeneity of GC, and the complexity of the microenvironment, researchers around the world have been exploring personalized and precise immunotherapy strategies for GC patients with specific features in recent years.

Current situation of immunotherapy for G/GEJC with specific subtypes

G/GEJC is characterized by significant heterogeneity and a diversity of molecular subtypes. Several studies concentrating on specific molecular subtypes of G/GEJC have demonstrated that combination immunotherapy may offer survival advantages, especially in patients with MSI-H and EBV infection. These insights are often derived from subgroup analyses in combination immunotherapy clinical trials [80, 96], or from retrospective evaluations of clinical data [105]. The discussion below will detail the investigation of immunotherapy for particular subtypes of G/GEJC.

HER-2

HER-2 is a classic target for G/GEJC. ToGA study had established trastuzumab combined with chemotherapy as the first-line standard treatment for HER-2-positive G/GEJC [106]. Since then, despite numerous failures in the development of new drugs targeting HER-2, the emergence of immunotherapy and antibody–drug conjugates (ADCs) has dramatically altered the treatment landscape for HER-2-positive GC. KEYNOTE-811 evaluated trastuzumab plus chemotherapy with or without pembrolizumab in the first-line treatment of HER-2-positive G/GEJC. The third interim analysis (follow-up 38.5 months) [88] showed significant improvement in mPFS and mOS for patients with ITT and PD-L1 CPS ≥ 1 when combined with ICIs. However, for patients with CPS < 1, there was no difference in mPFS and mOS was worse when combined immunotherapy.Given the promising results, NCCN guidelines have restricted the indication for combined immunotherapy to HER-2-positive patients with CPS ≥ 1. The lack of survival benefit of combination immunotherapy in patients with CPS < 1 may stem from a higher percentage of patients in the control group receiving subsequent treatments (47% vs 39%). Additionally, HER-2 may up-regulate PD-L1 expression [107], and the status of immunotherapy included in follow-up treatment after progression in the control group is unknown, which could also affect the outcome. Besides trastuzumab combined with immunotherapy, HER-2-targeted ADCs have also achieved satisfying efficacy in clinical application of G/GEJC. The results from the Chinese mul-ticenter phase I study C103 [108] showed that in HER-2-expressing G/GEJC who progressed after at least first-line treatment, the ORR for disitamab vedotin (RC-48) combined with toripalimab was 43%, with an mPFS of 6.2 months, and an mOS of 16.8 months. Clinical benefits were also observed in HER-2-positive and low-expressing patients at the recommended phase II dose, with ORR of 56% and 46%, respectively, and mPFS of 7.8 months and 5.8 months, respectively. These outcomes represent a significant improvement compared over previous data for RC-48 monotherapy. In addition, duration of response (DoR) in second-line treatment was significantly outperformed that in third-line and subsequent treatments, 15.6 months and 3.6 months respectively, suggesting that the advance of the combined treatment regimen could yield superior benefits. A clinical trial evaluating RC-48 combined with toripalimab and chemotherapy or trastuzumab in first-line HER-2-expressing G/GEJC is ongoing (NCT05980481), with results eagerly anticipated.

Claudin18. 2

According to the SPOTLIGHT [109] and GLOW [110], Zolbetuximab, a CLDN18.2 mAb in combination with chemotherapy provided a survival benefit in first-line Claudin18.2-positive (≥ 75% of tumor cells exhibit moderate-to-strong membranous CLDN18.2 staining) and HER-2-negative G/GEJC. Preclinical study [111] demonstrated that zolbetuximab combined with chemotherapy can significantly promote CD8 + T cell infiltration. Moreover, when combined with PD-1 mAb, zolbetuximab more effectively inhibited tumour growth, and significantly increased the number of mice achieving complete response (CR), indicating a synergistic effect. The phase 1 study of QLS31905 [112] a CLDN18.2/CD3 bispecific antibody, in advanced solid tumours, presented at the 2023 ESMO, included 31 patients with G/GEJC and initially reported safety and efficacy data. Further clinical studies focusing on CLDN18.2 combination immunotherapy are currently underway (NCT06206733, NCT05964543).

dMMR/MSI-H

Notable characteristics of dMMR/MSI-H G/GEJC include a favourable prognosis, chemotherapy insensitivity and obvious benefits from immunotherapy [113]. Due to the low incidence, high-level evidence from large sample studies is absent. Subgroup analyses from the KEYNOTE-059, KEYNOTE-061 and KEYNOTE-062 studies initially suggested that PD-1 mAb exhibits a high response rate and prolonged survival in dMMR/MSI-H G/GEJC [114]. Preliminary findings from the phase II prospective single-arm NO LIMIT study [115] indicate that nivolumab combined with ipilimumab as a first-line treatment results in an ORR of 62.1%, with 3 patients (10%) achieving CR, a DCR of 79.3%, and a mPFS of 13.8 months, aligning with the established safety profile of dual immunotherapy. In addition, results from multiple prospective phase II studies have shown therapeutic efficacy of envafolimab [116], tislelizumab [117], serplulimab [118], and pembrolizumab [119] in the second-line treatment of dMMR/MSI-H advanced G/GEJC patients who have not previously received immunotherapy..

EBVaGC

EBV associated GC (EBVaGC) patients, characterized by highly active immune-active TME [120], may benefit from immunotherapy. However, the efficacy of immunotherapy in EBV-positive G/GEJC are inconsistent [105, 121, 122]. At present, a clinical trial investigating double-immunotherapy in EBV-positive G/GEJC patients is nearing completion (NCT04202601), with results eagerly anticipated.

AFPGC

Alpha fetoprotein-producing G/GEJC (AFPGC), a distinct subtype of G/GEJC, comprises approximately 6% and is known for its predisposition to liver and lymph node metastases, elevated AFP levels, and a poor prognosis [123]. A prospective phase II study [124] demonstrated that a combination of camrelizumab, apatinib and SOX achieved promising outcomes in advanced G/GEJC patients with AFP > 2 × ULN or AFP positive, with an ORR of 55.6%, a 12-month PFS rate of 42.1%, and a 12-month OS rate of 63.7%.

The results presented above indicate that G/GEJC, which expresses numerous specific targets, can be effective in combination with immunotherapy. The subsequent steps involve further validating the efficacy through additional phase III studies and conducting translational research to elucidate the precise mechanisms impacting immunotherapy, aiming to facilitate its clinical application. Furthermore, it is important to note that while G/GEJC expressing different targets can receive targeted therapy, multiple targets are often expressed simultaneously. In this regard, determining the appropriate weighting and sequencing of different targeted therapies is crucial, necessitating a deeper understanding of tumours and the TME.

With the continuous exploration of new targets, molecular typing is becoming increasingly precise, and treatment options tending to be diversified. Compared to EC and CRC, molecular targets that have been discovered and successfully translated into clinical practice are enriched in G/GEJC. Therefore, an increasing number of biomarker-driven gastric cancer-related clinical studies are underway, covering populations with dMMR/MSI (NCT 06346197), HER-2 positive (NCT05152147), CLDN18.2 positive (NCT06093425, NCT06206733), PD-L1 positive (NCT06346197, NCT06093425, NCT06206733), FGFR2b positive (NCT05111626) GC. For patients expressing specific targets, a major goal of biomarker-driven research is to further explore whether immune combination targeted therapy produces synergistic anti-tumour effects and whether combination therapy can help these patients achieve a chemo-free state. Additionally, accompanying clinical research to further explore biomarkers of patient efficacy and survival, and to uncover resistance mechanisms and reversal strategies, will aid in achieving a sustained refinement of GC stratification (Table 4).

Table 4 Summary of ongoing key phase III clinical trials of immunotherapy in G/GEJC

Challenges in immunotherapy for G/GEJC

Efficacy-enhancing strategy remains unclear

Despite the advancements of ICIs as first-line regimen in G/GEJC, approximately 40% of patients still could not benefit from combination therapy with single-agent ICIs. The mOS shows a modest extension of merely 1–2 months for intention-to-treat (ITT) patients and 4–5 months for those with high PD-L1 expression, which is insufficient to meet clinical expectations.

Clinical trials investigating combination therapy never stop. The dual blockade of nivolumab plus ipilimumab cohort in CheckMate-649 [5] did not demonstrate prolonged survival compared to chemotherapy alone, indicating that chemo-free regimens may not be suitable for all patients with advanced G/GEJC. COMPASSION-15 [125] showed that cadonilimab (AK104, a PD-1/CTLA bispecific antibody) combined with chemotherapy significantly improved mOS compared with placebo combination chemotherapy (15.0 months vs 10.8 months, HR 0.62), even in PD-L1 low expression patients (PD-L1 22C3 CPS < 5), indicating that cadonilimab combined with chemotherapy benefits patients regardless of PD-L1 expression status, providing a new treatment approach for HER-2 negative and PD-L1 low expression G/GEJC. Additionally, The sequence of medications in treatment protocols warrants careful consideration. The JAVELIN Gastric 100,103 study [126] evaluated the effect of fusing Avelumab, a PD-L1 monoclonal antibody, as maintenance therapy after a minimum of 12 weeks of first-line chemotherapy. While OS benefits were not observed universally, notable differences were noted among patients without metastatic sites after induction chemotherapy and a small subset with MSI-H. What’s more, avelumab showed favourable safety profiles compared to continuous chemotherapy, offering new insights for maintenance therapy in selected subgroups of G/GEJC patients.

Perioperative immunotherapy strategies were yet established

Encouraged by the results of immunotherapy in advanced G/GEJC [98], the integration of immunotherapy in frontline treatments is under investigation recently.

Several phase II studies have demonstrated that neoadjuvant chemotherapy combined with immunotherapy can improve pCR or major pathologic response (MPR) rate [127,128,129]. KEYNOTE-585 [16] revealed that neoadjuvant pembrolizumab combined with chemotherapy significantly improved pCR rate. Although long-term follow-up did not demonstrate statistical improvement in event-free survival (EFS), there was a noticeable trend towards delaying disease recurrence. The near-significant P-value highlights the importance of statistical design in clinical trial research. Further long-term follow-up data from MATTERHORN [91] and HLX-10 (NCT04139135) neoadjuvant studies are expected.

ATTRACTION-5 [92] evaluated the efficacy of nivolumab in combination with chemotherapy as adjuvant therapy for phase III G/GEJC, revealing no significant improvement in relapse-free survival (RFS). However, subgroup analysis indicated benefits from immunotherapy for patients with ECOG 1, postoperative pathology stage IIIc, and PD-L1 tumour proportion score ≥ 1.

For the perioperative treatment for the special MSI-H G/GEJC, GERCOR NEONIPIGA [11] and INFINITY [130] preliminarily explored the feasibility of perioperative immunotherapy alone for locally advanced MSI-H G/GEJC without chemotherapy, and achieved satisfactory pCR rates of 58.6–60.0%.

All in all, the significant heterogeneity of G/GEJC and the complexity of its microenvironment present formidable challenges in advancing immunotherapy. In the era of precision medicine, developing strategies to overcome this heterogeneity and achieve “high efficiency and low toxicity”, comprehensive management of immunotherapy will be crucial for future advancements.

Colorectal cancer

History and current state of immunotherapy for CRC

Immunotherapy has been utilised against dMMR/MSI-H CRC (Table 5). The efficacy of this cancer subtype is largely due to the high tumour mutational burden, abundance of neoantigens, and increased immune cell infiltration [131, 132]. Compared with that of other tumour types, the incidence of MSI-H/dMMR in CRC is relatively high (~ 15%) [132, 133]. MSI-H/dMMR CRC tends to be less sensitive to chemotherapy-based treatments and has a lower ORR than that of microsatellite stable (MSS)/proficient mismatch repair (pMMR) CRC [134,135,136,137]. For later-line treatment of MSI-H/dMMR metastatic CRC (mCRC), multiple studies have confirmed the efficacy of immunotherapy, with PD-1 monotherapy showing an ORR of 32.8–40% [138,139,140]. The results from these clinical studies have facilitated the exploration of immunotherapy as a first-line option against MSI-H/dMMR mCRC. The KEYNOTE-177 evaluated the antitumor activity of pembrolizumab (pembro) vs. chemotherapy ± bevacizumab or cetuximab (chemo) and showed that pembro was superior to chemo for mPFS (16.5 m vs. 8.2 m; HR 0.60) [141]. Additionally, CheckMate-8HW compared nivolumab (NIVO) + ipilimumab (IPI) with NIVO or chemotherapy (chemo); NIVO + IPI demonstrated clinically meaningful and statistically significant improvement in PFS vs. chemo, with a 79% reduction in the risk of disease progression or death [142]. These trials demonstrated that both mono- and dual-immunotherapy were superior to chemotherapy in terms of efficacy and reducing associated AEs.

Table 5 Summary of key clinical trials of immunotherapy in CRC

Challenges in immunotherapy for MSI-H/dMMR CRC

Efficacy-enhancing strategy remains unclear

The retrospective study by Chen et al. demonstrated that chemo-anti–PD-1/PD-L1 therapy was more effective than anti–PD-1/PD-L1 alone in treating MSI/dMMR GI cancers, offering better survival benefits [145]. Another phase III trial, COMMIT, is currently assessing the safety and effectiveness of atezolizumab alone versus its combination with mFOLFOX6 and bevacizumab as a first-line treatment for dMMR/MSI-H mCRC [146]. Wu et al. investigated the efficacy of combining COX inhibitors with ICIs for the treatment of dMMR/MSI-H mCRC. The results indicated that the mPFS and mOS were not reached, with an ORR of 73.3% [147]. Additionally, the results of ongoing CheckMate-8HW study comparing NIVO + IPI to Nivolumab monotherapy will provide insight into whether first-line NIVO + IPI can enhance the therapeutic efficacy over NIVO for dMMR/MSI-H mCRC [142]. More clinical trials on ongoing are detailed in the Table 6.

Table 6 Summary of ongoing key phase III clinical trials of immunotherapy in CRC
Optimal treatment strategies following progression necessitates explored

A retrospective study analysed 51 patients with dMMR/MSI-H GI cancers who continued to receive antitumor therapy after progression on immunotherapy. Of these patients, 35 cases (68.6%) were mCRC. The study concluded that continuing with anti-PD-1/PD-L1 therapy along with other drugs significantly improved the disease control rate, PFS, and OS compared with receiving chemotherapy alone or with targeted therapy [148]. The ongoing NIPIRESCUE study (NCT05310643) evaluates the efficacy of nivolumab and ipilimumab in MSI/dMMR mCRC patients following resistance to PD-1 monotherapy. The results are eagerly anticipated.

Perioperative immunotherapy strategies require further validation

Despite no statistical difference in OS between the Pembrolizumab group and the Chemotherapy group in KEYNOTE-177 [141], a trend towards decreased mortality risk was observed, indicating that immunotherapy should be utilised early and encouraging further investigation of immunotherapy in the perioperative treatment of CRC. The NICHE-1 [149] and NICHE-2 trial [150], demonstrated effective neoadjuvant outcomes for NIVO + IPI, showing pCR rates of 60–67% and MPR rates of 95–97%.

Eventhough the long-term follow-up data was absent, the high pCR rates notably preserved organ function and enhanced Quality of Life, making this regimen widely accepted. Nevertheless, there are still challenges. Given the AEs associated with dual immunotherapy, the reliability of mono-immunotherapy and the appropriate treatment duration remain unclear. Efforts to resolve these challenges never stop. A phase II MSKCC study explored the PD-1 antibody, dostarlimab, for a total of 6-months neoadjuvant the results showed that all patients achieved cCR (12/12), with manageable safety [151].

Explorations of other combined strategies are ongoing. For example, The phase II study PICC showed the possibility of toripalimab in combination with celecoxib (COX-2 inhibitor) [152]. What’s more, research is also being conducted on adjuvant immunotherapy regimens for dMMR/MSI-H CRC [153, 154], and the findings from these studies will shed light on whether adjuvant immunotherapy can boost survival rates in stage III dMMR/MSI-H or POLE-mutated CRC, which will significantly influence future treatment approaches.

Challenges in immunotherapy for pMMR/MSS CRC

Most patients with mCRC have pMMR/ MSS tumours, which are considered “cold tumours”.The low immunogenicity of these tumours makes them less recognizable by CD8 + T cells. Cold tumours are characterized by the overexpression of innate immune inhibitory oncogenic pathways and the presence of numerous immunosuppressive factors in the tumour microenvironment [155]. Consequently, single-agent immunotherapy often yields unsatisfactory results, necessitating the exploration of new treatment strategies and the identification of populations likely to benefit from immunotherapy.

Treatment strategies remain controversial

The primary current explorations mostly focus on combined treatments enhancing the sensitivity of immunotherapy [156]. Nevertheless, most of the research is still in the early exploratory stage with small samples, and the results remain further validation. The LEAP-017 study suggested that the lenvatinib plus pembrolizumab showed no improvement in survival in the pMMR/MSS mCRC patients compared with the standard of care group, with OS being 9.8 vs. 9.3 months (HR 0.83, 95% CI 0.68–1.02, P = 0.0379) [157]. A study included 39 pMMR/MSS mCRC patients who experienced progression following standard chemotherapy and received the RIN regimen (regorafenib + ipilimumab + nivolumab). The RP2D cohort's patients had an ORR, mPFS, and mOS of 27.6%, 4 months, and 20 months, respectively, which demonstrated promising prospective of clinical application [158]. Segal et al. included 24 chemotherapy-refractory pMMR mCRC patients and administered durvalumab and tremelimumab with concurrent radiotherapy, with an ORR of 8.3%. The mPFS and mOS were 1.8 month and 11.4 months correspondingly, which did not meet the prespecified endpoint [159]. Meanwhile, Thibaudin et al. included 57 patients with RAS-mutated mCRC treated with first-line durvalumab and tremelimumab plus mFOLFOX6. The 3-month PFS for MSS patients was 90.7%, with a response rate of 64.5%; mPFS was 8.2 months, while OS was not reached [160]. The CAPability-01 study evaluated the efficacy of romidepsin, a histone deacetylase inhibitor, plus sintilimab and bevacizumab or romidepsin plus sintilimab in treating patients with advanced or metastatic pMMR/MSS CRC who have failed at least two lines of prior treatment and reported an ORR of 44.0% and a PFS of 7.3 months for the triplet group [161]. More clinical trials on ongoing are detailed in the Table 6.

Features of potential beneficiaries remain unclear

A study that explored the efficacy of treatment with PD-1, BRAF, and MEK inhibition in BRAFV600E CRC patients exhibited an ORR of 25% and a DCR of 75%, with a mPFS of 5 months in MSS CRC patients [162]. Acquired resistance to temozolomide may be associated with the onset of hypermutation, facilitating immune sensitisation. The MAYA study included 33 previously treated O6-methylguanine–DNA methyltransferase silenced MSS mCRC patients who received temozolomide followed by combination with low-dose ipilimumab and nivolumab. The results reported mPFS and mOS of 7.0 and 18.4 months, respectively, and an ORR of 45% [163]. The biomarker analysis from the AtezoTRIBE study indicated that patients with high immunomodulatory signature scores within the pMMR subgroup derived superior benefits from atezolizumab [164]. The CheckMate-9X8 study reported that MSS/pMMR mCRC patients with consensus molecular subtype (CMS) 1 and CMS3 at baseline had a higher probability of being progression-free at 12 months when treated with nivolumab in combination with the standard treatment regimen [165].

Anal carcinoma

Anal carcinoma (AC) is a rare malignant tumour of the digestive system, accounting for approximately 2% of all GI cancers, with anal squamous cell cancer (ASCC) constituting the majority (approximately 80%) of AC [166]. The treatment options for advanced ASCC are limited, predominantly relying on chemotherapy; therefore, new therapeutic approaches are warranted [167,168,169]. Unlike EC, G/GEJC, and CRC, human papillomavirus (HPV) infection is closely associated with the occurrence of ASCC. The HPV oncoproteins are immunogenic and can trigger the host's anti-tumour immune response. Therefore, immunotherapy holds promise as a treatment strategy for ASCC. However, due to its low incidence, only preliminary exploration has been conducted. The results of the studies KEYNOTE-158, POD1UM-202, and CARACAS, have demonstrated the efficacy of single-agent ICIs in treating metastatic ASCC that has failed previous treatments. The ORR ranged from 10–24%, PFS from 2.0–4.1 months, and OS from 10.1–12.8 months in these studies [170,171,172,173]. There has also been some preliminary exploration into combination therapies in populations with failed prior treatments.

The NCI ETCTN (NCI9673 Part B) study showed a trend towards prolonged PFS and OS with dual immunotherapy (nivolumab + ipilimumab) compared with nivolumab monotherapy for metastatic ASCC with failed previous treatments. However, the differences were not statistically significant(PFS: 2.9 vs. 3.7 months, HR 0.80, 95% CI 0.51–1.24; OS: 15.4 months vs. 20.0 months, HR 0.86, 95% CI 0.51–1.47) [174]. The CARACAS study also indicated that for advanced ASCC with failed previous treatments, the ORR for PD-L1 antibody avelumab combined with cetuximab was 17%, with PFS at 3.9 months and median OS at 7.8 months [171]. Atezolizumab, another PD-L1 antibody, combined with bevacizumab, showed no superior results to single-agent ICI in treating advanced ASCC with failed previous treatments, with an ORR of 10%, PFS of 4.1 months, and OS of 11.6 months [175]. In the first-line treatment setting, the SCARCE study was the first to validate the combination of ICI with chemotherapy for first-line treatment of advanced ASCC, although failed to meet the primary endpoint, with 1-year PFS rates of 45% for atezolizumab + mDCF (docetaxel + cisplatin + fluorouracil) and 43% for mDCF alone, 12-month OS rates of 77% and 81% respectively, and ORRs of 75% and 78% [176]. Several phase III RCT studies are currently ongoing [177, 178] to evaluate the role of chemotherapy combined with ICI in the first-line treatment of advanced ASCC.

In summary, ICI therapy can improve survival for patients with advanced ASCC who have previously received treatment. However, the role of ICI in the first-line treatment of advanced ASCC remains unclear. Furthermore, identification of the population that would benefit from the treatment remains necessary, as does exploration of the optimal combination treatment strategy. For the treatment of locally advanced ASCC, several studies are currently in the early stages of exploration (NCT04230759 and NCT03233711).

Collectively, immunotherapy has advanced the treatment for EC and G/GEJC with significant breakthroughs in CRC, although several challenges remain to be addressed, including the diversity and complex molecular types in GC, the optimal treatment regimen for MSI-H/dMMR CRC, the potential for immunotherapy in MSS/pMMR, and nutritional issues in EC. In addition to the specific limitations associated with each tumour type, other challenges, including identifying biomarkers and the mechanisms underlying resistance to immunotherapy, optimisation of methods for assessing the efficacy of immunotherapy, determination of the best combination treatment modalities, and development of new immunotherapeutic agents remain to be addressed. With more studies, the application of immunotherapy in the field of GI cancers will be further optimised, thereby improving clinical outcomes.

Biomarkers of immunotherapy

The exploration of biomarkers is crucial for realising precise and individualised immunotherapy and has witnessed significant advancements in recent years. This progress is primarily evident in two aspects. First, the predictive efficacy of classical markers has been validated and discussed in large-scale, multi-dimensional studies, thereby elucidating their strengths and weaknesses. Second, with the development and cost reduction of cutting-edge technologies, including high-throughput detection, visualisation, liquid biopsy, and artificial intelligence, researchers have shifted their focus from tumour cells to the entire tumour microenvironment, from single to multiple dimensions, and from a static to a dynamic pattern. Here, we elaborate on the progress of both classical and novel biomarkers based on these two perspectives (Fig. 1).

Fig. 1
figure 1

Biomarkers under-investigation for immunotherapy in GI tract cancers. The pattern diagram summarized immunotherapeutic biomarkers under-investigations in GI tract cancers from three perspectives including tumour cells and tumour microenvironment, peripheral components and etiology

Classic biomarkers

PD-L1

PD-L1 positivity in EC, G/GEJC, CRC, and immune cells has shown variability, ranging from 17.4% to 43.5%, 12.0% to 43.6%, 8.1% to 44.0%, and 15.3% to 69.0%, respectively [179]. While the predictive utility of PD-L1 expression assessed using the CPS appears limited in CRC immunotherapy, it demonstrates a robust predictive value in EC and G/GEJC. The cut-off values of CPSs in clinical trials are commonly defined as 1, 5, and 10. The ATTRACTION-03 and CheckMate-648 trials revealed nivolumab benefited the entire population for advanced EC regardless of CPS [31, 180]. Conversely, the KEYNOTE series trials of EC demonstrated that pembrolizumab treatment predominantly benefited patients with high PD-L1 expression (CPS ≥ 10) [181]. Similar to studies on EC, the KEYNOTE series of trials on advanced GC has highlighted the therapeutic advantages of immunotherapy in populations with high PD-L1 expression (CPS ≥ 10) [80], while the ChecMate-649 [5, 182] and ORIENT-16 trials [6] validated CPS ≥ 5 for identifying patients that would benefit from chemotherapy combined with nivolumab or sintilimab, respectively.

However, the use of different detection antibodies and assay platforms in various clinical trials has led to confusion and a lack of uniformity in the CPS cut-off values used to identify the beneficiary population, and conclusions drawn from existing studies have limited applicability. Several studies have attempted to address this variability. A meta-analysis incorporating 6,488 cases of advanced GC immunotherapy for PD-L1 demonstrated significant survival benefits in patients with CPS ≥ 1, regardless of whether they underwent monotherapy or combination immunotherapy (single-agent immunotherapy: OS (hazard ratio (HR) 0.84, 95% confidence interval (CI) 0.74–0.96); combination immunotherapy: OS (HR 0.81, 95% CI 0.71–0.92) and PFS (HR 0.77, 95% CI 0.69–0.86)) [183]. However, another retrospective study analysing low-CPS subgroups in CheckMate-649, KEYNOTE-062, and KEYNOTE-590 (CPS: 1–4; CPS: 1–9) observed that patients with low PD-L1 expression did not benefit from chemotherapy combined with ICIs [184]. Notably, these findings pertained specifically to advanced G/GEJC. Furthermore, the PD-L1 CPS faces additional challenges as a critical biomarker of immunotherapy efficacy. First, PD-L1 expression exhibits considerable heterogeneity [185] and can be dynamically influenced by various therapeutic approaches [186, 187], affecting its reliability as a biomarker of efficacy. Second, there is no standardised method for PD-L1 testing, leading to low consistency among commercially available antibodies. For examples, PD-L1 positivity with the Dako 28–8 antibody, which is used in the CheckMate-649 trial, is higher than that with the Dako 22C3 antibody, which is used in the KEYNOTE series of trials. This suggests that with a CPS cut-off value of 5, the positivity for Dako 28–8 is approximately twice that for Dako 22C3 [188].

Therefore, establishing a consensus on PD-L1 testing is essential to standardise diagnostic platforms across different clinical trials and accelerate the identification of biomarkers for different stages. Integrating various therapeutic drugs, patient clinical characteristics and other relevant biomarkers is key to achieving breakthroughs in this field.

MSI-H/dMMR

MSI-H and dMMR are known to increase somatic cell mutations and neoantigen production, often resulting in extensive lymphocytic infiltration and elevated expression of immune checkpoints in the tumour microenvironment (TME) [189, 190]. Consequently, patients with MSI-H/dMMR exhibit poor response to cytotoxic drugs and high sensitivity to immunotherapies. Emerging evidence suggests that the predictive value of MSI-H/dMMR for immunotherapy efficacy in GI tract cancers is significantly superior to that of other biomarkers. Nevertheless, the incidence of MSI-H/dMMR in EC is relatively low. Currently, there are no large-scale clinical trial data available, and most studies have focused on GC and CRC.

A retrospective analysis [114] of three trials (KEYNOTE-059, 061, and 062) revealed that patients with MSI-H advanced G/GEJC benefited significantly from first-line to third-line pembrolizumab monotherapy, establishing MSI-H as a predictive biomarker for efficacy. Similarly, in the KEYNOTE-158 trial, all 24 patients with GC and MSI-H/dMMR benefited from pembrolizumab treatment [191]. The 3-year follow-up results from the CheckMate-649 trial [180] showed that in unselected MSI-H subgroup of patients, nivolumab combined with chemotherapy versus chemotherapy alone led to an improved median OS, further supporting MSI-H as an effective biomarker for identifying advanced patients with GC for immunotherapy. Several clinical trials on perioperative immunotherapy for GC are ongoing, and the interim results suggest that MSI status, not PD-L1 CPS, is the most crucial biomarker for predicting the benefit of perioperative pembrolizumab with chemotherapy in locally advanced GC [11, 15, 16]. Therefore, MSI-H can also serve as a biomarker for the efficacy of perioperative immunotherapy in GC, laying the foundation for the exploration of precise treatment patterns during the perioperative period with promising application prospects.

To date, MSI-H/dMMR is the most important biomarker for predicting the efficacy of immunotherapy in CRC [138, 139, 192,193,194], as confirmed by the 5-year OS results from KEYNOTE-177 [195] and trials such as NICHE-2 [20], NICHE-3 [196], and PICC [21] which demonstrate the significant role of MSI-H/dMMR as a biomarker of immunotherapy efficacy across different CRC stages.

However, several challenges remain for the use of MSI-H/dMMR as a biomarker of immunotherapy efficacy. First, MSI-H/dMMR is underrepresented in the population, accounting for only 15% of CRC, 7% of GC, 4% of GEJC, 0.4% of EAC, and 0% of ESCC cases [197, 198]. Moreover, although the overall percentage of patients with MSI-H/dMMR in CRC is 15%, only 5% of patients with advanced CRC have MSI-H, greatly limiting its clinical practice as a biomarker of immunotherapy efficacy. Second, the accuracy of MSI detection and interpretation needs to be improved. Currently, immunohistochemistry (IHC) is the main method for detecting MMR protein expression, whereas polymerase chain reaction (PCR) and next-generation sequencing (NGS) are the main methods used to identify MSI status. PCR is the only approved method for detecting MSI in CRC, and NGS-based MSI detection methods need to be verified using more data. Additionally, MSI detection methods for G/GEJC need to be supported by a large sample size and reliable data. In 2022, the College of American Pathologists (CAP) released the recommended examination protocols for MSI/MMR in different tumour types [199]: (1) For patients with CRC, the use of either IHC to detect MMR or PCR to detect MSI is recommended. Although these methods are preferred, an NGS assay that has been validated for dMMR could also be used to determine the MSI status. (2) For patients with gastroesophageal or small intestinal cancer, IHC should be used to identify MMR or PCR should be used to identify MSI. Although the consistency between MMR-IHC and MSI-PCR testing results for various solid cancers is 90.3–99.4% [200,201,202,203], the accuracy of MMR testing and interpretation is affected by several factors, including tumour heterogeneity (such as variability between different lesions and intra-tumoural heterogeneity), antitumour therapies, the experience of the pathologist, and detection platforms [204]. Currently, any evidence of dMMR or MSI should be interpreted as a positive result, thereby qualifying patients for immunotherapy. Additionally, inconsistent results should be rigorously reviewed to ensure they are not due to misjudgement.

Tumour mutational burden

Tumour mutational burden-high (TMB-H) is defined as 10 or more mutations per Mb, with an incidence of 5% in patients with CRC, 8% in patients with GC, 3% in GEJC, 2% in patients with EAC, and 3% in those with ESSC [198, 205]. Given its potential to generate immunogenic neoantigens, the application of TMB status as a biomarker of immunotherapy efficacy has been widely studied. However, these conclusions remain controversial [206,207,208].

Based on the findings of KEYNOTE-158 study [191], in 2020, the FDA approved the use of pembrolizumab in patients with TMB-H cancers. However, a retrospective analysis of data from 10,000 solid tumour samples included in The Cancer Genome Atlas Program (TCGA) [209] revealed that TMB status is not correlated with immunotherapy efficacy in GI tract cancers, indicating distinct immune patterns of GI tract cancers. Furthermore, the study divided cancers into two categories based on whether CD8 + T cell infiltration was positively correlated with the generation of tumour neoantigens. Category I cancers were defined as cancers with CD8 + T cell infiltration positively correlated with neoantigens productions, while category II exhibited reverse relationship.TMB-H could facilitate the selection of patients that might benefit from immunotherapy for Category I tumour but not for Category II cancers; notably, most GI cancers were Category II. Therefore, the infiltration of antigen-specific immunocytes may be a decisive factor underlying the influence of TMB-H on immunotherapy efficacy. Moreover, a retrospective study of 48,606 GI tract cancers exploring the characteristics of gene mutations associated with TMB-H independent of dMMR/MSI [205] found that not all gene mutations related to TMB-H could play a role in influencing immunotherapy efficacy. Based on the mutations that influence immunotherapy efficacy, this study established an mTMB model and discovered an increase in the infiltration of cells associated with antitumour immunity, such as M1 and CD8 + T cells, among patients with mTMB-H. Collectively, further division based on immune cell infiltration and gene variation in patients with GI tract cancers and different TMB statuses is warranted.

Novel biomarkers under-investigation

Virus and microbiome

EBV

9% of patients with GC have EBVaGC, and 80% of them harbour an immune-inflamed microenvironment with enriched T-cell and B-cell infiltration [101, 210,211,212]. Consequently, EBV is considered a potential marker for the efficacy of GC immunotherapy.

While clinical trials with large sample sizes specifically focusing on EBVaGC are mostly ongoing, certain small-sample studies have indicated that EBVaGC has a relatively high response rate to immunotherapy. For instance, a phase II clinical study conducted a retrospective analysis on the use of pembrolizumab for treating advanced GC, reporting a 100% response rate and an mOS time of 8.5 months among six cases of EBVaGC [213]. In contrast, a prospective clinical study enrolling nine patients with advanced EBVaGC treated with a combination of PD-1 and CTLA-4 inhibitors (including two patients receiving first-line treatment, four receiving second-line treatment, and three receiving third-line treatment) found an ORR and DCR of 55.6% and 88.9%, respectively [214]. A retrospective study involving 66 patients with advanced GC treated with PD-1 and CTLA-4 inhibitors indicated that patients with EBV + /pMMR (n = 22) displayed significantly superior outcomes than those with EBV-/pMMR (n = 44) (ORR: 54.6% vs 17.7%, P = 0.008; mPFS: 8.5 months vs 2.0 months, P < 0.001; mOS: not reached vs 5.0 months, P = 0.002) [105]. Nevertheless, the conclusions drawn by these studies were limited in generalisability owing to factors such as sample size and the presence of other molecular markers (such as PD-L1 and MSI), necessitating further validation.

Helicobacter pylori

Helicobacter pylori (H.pylori) considerably contributes to the pathogenesis of GI tract cancers by producing the cytotoxins VacA and CagA [215,216,217]. Among patients with GI tract cancers receiving immunotherapy, those with H.pylori infection exhibited shorter PFS and OS [218,219,220]. However, these studies have several limitations, including the small size of the H.pylori infection samples, inconsistent methods used to detect H.pylori infection, and varied standards for diagnosing H.pylori positivity, that dramatically influenced the reliability of the conclusions. In a recent prospective observational study involving 10,122 patients with GI tract cancers [104], the incidence rates of H.pylori infection, as detected using the 13C breath test, were 44.19%, 33.24%, and 42.35% for patients with EC, GC, and CRC, respectively. Among the 636 patients with EBV- MSS GC receiving anti-PD-1/PD-L1 treatment enrolled in the study, patients with H.pylori infection showed a significantly better response [PFS: 6.97 months vs. 5.03 months; median immune-related overall survival (irOS): 18.30 months vs. 14.20 months P < 0.01]. of note, the impact of H.pylori infection on immunotherapy efficacy appears to be organ-specific. Patients with H.pylori infection and dMMR/MSI-H CRC and ESSC displayed shorter irPFS (dMMR/MSI-H CRC: 16.13 months vs. not reached, P = 0.042; ESSC: 5.57 months vs. 6.97 months, P = 0.029). The heterogeneity of the modulation of the gut microbiome and TME may explain the contrasting role of H.pylori in distinct GI tract cancers. Although this study provides powerful evidence for H.pylori as a biomarker of immunotherapy efficacy for GI tract cancers, the underlying mechanisms remain unclear. Considering the diversity and specificity of the microbiome in GI tract cancers, further studies should focus on the H.pylori function on the microbiome and tumour metabolic features.

Microbiome

The human microbiome, which resides primarily in the respiratory, digestive tracts, and skin, finds its highest abundance and diversity within the gut. Recent advancements in culture-independent microbial analysis methods have deepened our understanding of the gut microbiome's role in cancer's onset, development and treatment. Research increasingly shows that its composition, diversity, and specific communities modulate cancer patients' immune status, influencing immunotherapy outcomes. Compared to healthy individuals, patients with digestive tract cancers exhibit markedly diverse gut microbiome richness and diversity [221,222,223]. A reduction in microbial diversity and stability can compromise local immunity in the intestinal mucosa and trigger systemic immune responses via immune cells. This impairs both local and systemic immune functions, damages the mucosal barrier, and allows microbial elements to enter the systemic circulation, altering cytokine profiles throughout the body [224]. Moreover, metabolic byproducts from the gut microbiome, such as short-chain fatty acids (SCFAs), indole propionic acid, serotonin, and secondary bile acids, play crucial roles in regulating bacterial composition and activity [225]. These can permeate the intestinal barrier, impact host physiology, and activate immune responses, with SCFAs being particularly influential on host immunity. Studies highlight that the gut microbiome and its metabolites are pivotal in modulating innate immunity (including dendritic cells, macrophages, and NK cells), adaptive immunity (involving CD8 + T and CD4 + T cells), and tumour cell immunogenicity, thus affecting immunotherapy's effectiveness [226]. Research by Zhang et al. [227] demonstrated the impact of PD-1 inhibitors in CRC, which is highly influenced by gut microbiome diversity. Their use of broad-spectrum antibiotics to erase endogenous bacteria in mice diminished the PD-1 inhibitors' tumuor-suppressing effects, indicating a dependency on microbial diversity. Furthermore, studies have found a positive correlation between the abundance of Joshi's lactobacillus in the mouse gut microbiome and the response to anti-PD-1 immune checkpoint therapy. Enhancements in CD8 + T cell-mediated anti-PD-1 therapy efficacy were observed when supplemented with Joshi's lactobacillus or the tryptophan-derived metabolite indole-3-propionic acid [228]. An analysis of fecal samples from 106 patients with various rare cancers revealed that 22 strains, mainly from the phylum Firmicutes, significantly predict responses to combined anti-PD-1 and anti-CTLA-4 therapy. Strains that indicate better treatment responses are predominantly from the Ruminococcaceae family and the Faecalibacterium genus [229]. Additionally, the gut microbiome's predictive capabilities have also been validated in GI tract cancer cohorts. For instance, a study involving 74 patients with advanced digestive tract cancers undergoing PD-1/PD-L1 inhibitor therapy found a positive correlation between treatment response and levels of true bacteria, lactobacilli, and streptococci [230]. Similarly, another study with 117 patients suffering from HER-2-negative advanced GC or GEJC showed those with higher lactobacilli abundance experienced greater microbial diversity and improved response to anti-PD-1/PD-L1 therapy, often achieving better PFS [231]. The potential of using the gut microbiome for precise subtyping of digestive tract tumors was also explored in a study covering 77 advanced MSI-H/dMMR GI tract patients. By analyzing gut microbiomes, blood metabolomes, and cytokine and chemokine profiles, researchers identified specific microbes like B. caccae, V. parvula, V. atypica, and Clostridiales bacterium as potential subtyping markers for MSI-H/dMMR GI tract cancers, offering predictive insights into immunotherapy responses [232]. Despite its promise, the use of gut microbiome characteristics as markers for predicting the efficacy of ICIs encounters challenge of reproducibility issues, often exacerbated by factors like diet, regional differences, and ethnicity. Furthermore, the specificity of microbiome traits to particular treatment regimens [229] limits their clinical utility as reliable markers. Therefore, a comprehensive exploration of the gut microbiome spectrum and abundance in GI tumours, identification of broad-spectrum efficacy characteristic genera, and quantification of recognized beneficial or harmful bacteria proportions remain critical areas for future research, aiding in the realization of precision therapy.

Tumour microenvironment

The TME encompasses the internal environment where tumour cells develop, including all non-tumour components, metabolites, and secretions. These components include the extracellular matrix, fibroblasts, endothelial cells, immune cells, mesenchymal stem cells, cytokines, and metabolites [183]. The TME is complex, exhibiting both tumour-promoting and tumour-inhibiting effects [184,185,186,187,188,189]. Currently, identifying different TME statuses from multiple dimensions as biomarkers for the efficacy of immunotherapy is a focal area of research, with particular emphasis on the TIME.

TILs

Tumour-infiltrating lymphocytes (TILs) are integral components of the TME. With the integration of spatial information in histopathology and other technologies, the predictive value of TILs for the response to immunotherapy in GI tract cancers has gained attention, particularly focusing on CD8 + and CD4 + T cells.

The importance of CD8 + T cells in predicting therapeutic efficacy has been demonstrated in several clinical research cohorts. For example, the EC-CRT-001 study applied multipanel multiplex immunofluorescence to a cohort of patients with EC undergoing immunotherapy, revealing a significantly higher density of CD8 + T cells in the CR group than in the non-CR group [66]. However, with advancements in multi-omics technologies, it has become clear that a single marker is insufficient to represent the functional characteristics of a cell population. Further in-depth studies have attempted to reveal the functional CD8 + T cell subsets that play an active role in the response to immunotherapy. In the PANDA study, for instance, researchers discovered significantly higher densities of PD-1 + CD8 + T cells in responders to immunotherapy than in non-responders; such differences were not observed for other subsets of CD8 + T cells [233]. The NICHE study also identified the infiltration of PD-1 + CD8 + T cells as a predictor of response to immunotherapy in pMMR cancers [19]. These findings indicate that in the context of GI tract cancers, which are highly heterogeneous, CD8 + T cells can be predictive markers for the efficacy of immunotherapy and that differentiating these cells into subsets can provide more precise guidance for predictions [71]. CD4 + T cells primarily mediate the function and maintenance of CD8 + T cells [234]. FOXP3 + CD4 + T cells (Tregs) are believed to contribute to the inhibition of antitumour immunity and maintenance of immunological tolerance via immune checkpoints and high levels of anti-inflammatory cytokines, thus exerting adverse effects on prognosis [235]. However, further research has shown that an increased abundance of Tregs is not necessarily a poor prognostic indicator in GI tract cancers; in many cases, it is predictive of a superior prognosis. The SPACE study reported a correlation between high Treg infiltration and a more favourable prognosis [236]. Conversely, other clinical studies argued that Tregs level is not an effective predictor of immunotherapy efficacy. For instance, the NICHE study found no difference in baseline Tregs densities between responders and non-responders [19], and the PANDA study illustrated a lack of correlation between FOXP3 expression and therapeutic responses [233]. As these contradictory results indicate that Tregs consist of multiple subsets that vary substantially in terms of function and influence treatment responses, further dividing Tregs into more precise subsets may be necessary to better understand their impact on treatment outcomes. For example, Masuda et al. conducted single-cell sequencing of CRC samples and found that while total Tregs are associated with a more favourable prognosis, CD38 + Tregs exhibit a highly suppressive phenotype and are associated with a poor prognosis [237]. In an early clinical study of durvalumab and tremelimumab combined with chemotherapy, FOXP3- CD25 + CD4 + T cells were the only subset of CD25 + CD4 + T cells that correlated with survival;FOXP3 + CD25 + CD4 + T cells did not correlate with treatment outcomes [238].

Despite extensive research on TILs in GI cancer immunotherapy and the identification of numerous TIL subgroups with established links to immunotherapy response, TILs have yet to be used in clinical practice like PD-L1, MMR, and TMB. This can be attributed to several factors: (1) The plethora of detection methods for TILs complicates standardization. Current techniques include hematoxylin and eosin (HE) staining, IHC, mIHC, tissue microarray technology, multiphoton microscopy. Crucial yet rare TIL subgroups require identification and detection through advanced methods such as single-cell RNA sequencing and tumor tissue flow cytometry. Although HE and IHC staining are routinely used clinically, other methods demand high sample volume and viability, are expensive, and thus limit clinical application. Additionally, these methods involve lengthy testing and analysis cycles, taking at least a month from sample collection to result interpretation, which fails to meet urgent patient treatment timelines. (2) The challenge of standardizing interpretation methods and criteria. While HE staining and IHC are commonly employed in clinical settings, standardized interpretation methods and criteria are lacking. Efforts to standardize the clinical application of TILs commenced with breast cancer, highlighted by the International TILs Working Group’s 2014 consensus on manual interpretation standards based on HE staining [239]. This consensus outlined experimental procedures, calculation formulas, and interpretation criteria. Furthermore, studies indicated that stromal TILs (sTILs) are more abundant and consistently measurable than intratumoral TILs (iTILs), leading to recommendations from the International Immuno-Biological Markers Collaboration to prioritize sTILs in experimental research and clinical applications [240]. The combinatory predictive value of TILs and PD-L1 has been validated in multiple clinical trials, leading to their inclusion in the NCCN guidelines for breast cancer, offering insights for GI tract tumours. With advancements in visualization technology and artificial intelligence (AI), developing algorithms for analyzing tumour TILs based on pathological slice images and supporting pathologists in interpretation through computer assistance remains an active area of research.

Cytokines

Cytokines—including interferons (IFNs), interleukins (ILs), tumour necrosis factors (TNFs), and chemokines—produced by multiple immune cells have been suggested to be closely linked to immunotherapy for GI tract cancers.

IFNs are crucial inflammatory cytokines that activate the PD-1 signalling axis of tumour and stromal cells by upregulating PD-L1 expression [241]. A phase Ib study on neoadjuvant adebrelimab treatment for locally advanced ESCC reported a significantly higher level of IFN-γ expression among responders, along with a notable correlation between the IFN score and pathological regression [71]. Similar conclusions were drawn for EAC based on the findings of the PERFECT study [65]. Additionally, the KEYNOTE-028 study analysed the IFN-γ gene-expression profile of six genes and demonstrated that patients with higher expression could benefit more from pembrolizumab treatment [242].

ILs are soluble proteins secreted by leukocytes that are primarily involved in regulating immune cell functions. A meta-analysis summarising 14 studies on immunotherapy involving 3,190 participants indicate that individuals with high IL-8 levels exhibited lower ORR, OS, and PFS than those with low IL-8 levels [243]. Similarly, a phase IB/II clinical study of durvalumab and tremelimumab plus chemotherapy for advanced CRC revealed that high levels of IL-6 and IL-8 were correlated with lower response rates [238]. Additionally, a phase II clinical study of combined anti-PD-1 and anti-VEGF therapy for pMMR CRC showed a significantly increased level of cytokine signalling in responders compared with in non-responders [244].

Chemokines primarily coordinate cell movement during inflammation and induce directional migration of leukocytes and endothelial cells. However, studies on the relationship between chemokines and immunotherapeutic responses in GI tract cancers are currently limited. The MEDITREME study, which included patients with mCRC undergoing immunotherapy, reported that CXCL9, CXCL10, and CXCL11 expression levels were correlated with a more favourable PFS [238]. Similarly, in the PANDA study, responders exhibited a significant increase in CXCL13 expression compared with non-responders [233].

Currently, research on cytokines and chemokines primarily uses peripheral blood from patients, which necessitates small sample sizes and relatively uncomplicated detection techniques, predominantly ELISA, ELISPOT, and microscopy. These approaches facilitate dynamic and real-time monitoring of therapeutic efficacy. However, unlike emerging efficacy biomarkers like TILs under investigation, the utility of cytokines and chemokines in guiding the clinical practice of immunotherapy for GI tract cancers is limited by their low specificity and the challenges associated with establishing abnormal cut-off values. The levels of cytokines and chemokines are affected by a variety of factors, including infectious diseases, hormonal changes, trauma, general patient status, and treatment regimen, which significantly diminishes their specificity in reflecting tumour status. Moreover, current research predominantly relies on small cohorts and lacks extensive expression profiles of cytokines and chemokines in healthy populations and cancer patients, including the range of fluctuations following different treatments. Consequently, the application prospects of cytokines and chemokines should primarily focus on supplementing mainstream biomarkers to predict therapeutic efficacy and monitor disease status.

Owing to the complexity and diverse functions of the TME, a single-marker classification is insufficient to predict the efficacy of immunotherapy in patients with GI tract cancers. To address this, researchers have utilised data from the TCGA database of GI cancers, classifying the TME based on immune cell infiltration or distinct gene expression features, and developed a TME score to predict immunotherapy outcomes. Patients who benefit from immunotherapy typically harbour TME subtypes characterised by M1 macrophages, infiltration of CD8 + T cells, elevated expression of immune checkpoints, and robust T cell activity [245,246,247,248,249,250,251]. Tertiary lymphoid structures (TLS), ectopic lymphoid organs comprising B cells, plasma cells, CD4 + T cells, CD8 + T cells, DCs, macrophages, and neutrophils, are associated with improved OS in patients with GI tract cancers characterised by high infiltration of CD3 + T cells, CD8 + T cells, and M1 macrophage infiltration [252]. However, conclusive evidence based on large-scale studies correlating TLS with immunotherapy efficacy in GI tract cancers is still lacking [253, 254]. Additionally, the spatial distribution of immune cells relative to the immunotherapy response has been increasingly emphasised. Chen et al. analysed immune cell infiltration and its two-dimensional spatial distribution characteristics related to immunotherapy efficacy in 80 patients with GC using multiple immunofluorescence markers (immune checkpoints including PD-1, CTLA-4, TIM-3, LAG-3, and immune cell infiltration markers such as T, B cells, and macrophages). They established an infiltrating immune cells (TIICs) model, which included CD4 + FoxP3-PD-L1 + , CD8 + PD-1 -, the density of CD8 + PD-1-LAG3 -, and CD68 + STING + cells, as well as the spatial distribution of CD8 + PD-1 + LAG3-T cells. This study revealed that patients with low TIIC scores have better immunotherapy outcomes [10]. Moreover, in dMMR CRC, the proximity between PD-1 + cells and PD-L1 + cells can also serve as a positive marker of immunotherapy efficacy [255]. Furthermore, a study on the construction of a mathematical framework for cancer immunotherapy outcomes found that the increase in oxygen content caused by vascular normalisation promoted the polarisation of M1 cells and infiltration of immune cells, thereby promoting tumour cell killing and improving immunotherapy efficacy. Hence, mathematical models incorporating tumour cells, immune cells (M1, M2, CD4 + T, CD8 + T, Tregs), vascular and perivascular cells, and anti-angiogenic/pro-angiogenic-related molecules (including Ang1, Ang2, PDGF-B, VEGF, and CXCL12) effectively predict immunotherapy efficacy [256].

In addition to the TIME, the TME of GI tract cancers is also affected by perineural invasion, which is closely related to immunotherapy efficacy. A study involving 3,236 patients with GI cancers constructed a neuroinflammatory infiltration (NII) scoring system based on 26 PNI-related specific inflammatory genes. This study found that cancers with low NII scores infiltrated more immunosuppressive cells. Patients with PNI and high NII scores usually benefit from immunotherapy [257].

The advancement of cutting-edge techniques has enabled researchers to comprehensively characterise the TME in patients with GI. For instance, single-cell detection technology has been utilised to identify new cell subsets related to immunotherapy efficacy [212, 253]. High-resolution technology has facilitated the exploration of immune cell infiltration and spatial distribution. Moreover, the combination of multiomics detection and AI techniques has contributed to the extraction of key features influencing immunotherapy efficacy from extensive data [258, 259], aiding in the construction of predictive models.

Peripheral components

Liquid biopsy troika, consisting of circulating tumour DNA (ctDNA), circulating tumour cells (CTCs), and extracellular vesicles (EVs), offers several advantages for longitudinal monitoring of disease progression, minimally invasive sampling, and providing a holistic reflection of whole cancers. As a result, they are widely utilised for predicting therapeutic efficacy and prognosis and monitoring disease progression in GI cancers [260,261,262,263,264,265,266,267,268,269,270,271,272].

Profiling gene variation using NGS of baseline ctDNA as a biomarker has been universally explored. Gene subsets, including RAS, AKT, PTEN, PI3KCA, BRAF, ERBB2, MET, RAS, BRACA, POLE, POLD-1, TGFB2, RHOA, and PREX2 variations are reportedly related to immunotherapeutic efficacy in GI tract cancers [273,274,275,276,277].

Additionally, peripheral PD-L1is is an important biomarker for immunotherapy, with PD-L1 expression on CTCs or EVs during treatment playing diverse roles in predicting immunotherapy outcomes. The presence of PD-L1 expression on CTCs at baseline indicated a superior response to anti-PD-1/PD-L1 therapy, whereas. post-treatment PD-L1 + CTCs were specifically associated with resistance in GI tract cancers. Furthermore, pre-treatment PD-L1 + CTCs were considered to indicate the presence of druggable tumour cells, whereas post-treatment PD-L1 + CTCs contribute to immune evasion [278,279,280,281,282]. Baseline exosomal PD-L1 was regarded as an indicator of exhausted T cells, and thus unable to react to PD-1 inhibitors, whereas exosomal-PD-L1 released after treatment is thought to be relevant to the elevated activity of T cells [283, 284].

With the advancement of high-throughput detection of trace samples, an increasing number of studies have explored multi-omics features identified by analysing peripheral components as biomarkers for immunotherapy. For instance, Zhang et al. conducted a retrospective analysis of immune-related protein expression in plasma EVs from 112 patients with GC who received ICI therapy. They developed a predictive model (EV score) based on the expression of PD-L1, PD-L2, CD3, and Arg1 on EVs, showing that patients with higher EV-score have superior ICI outcomes [206]. Additionally, biomarkers from plasma proteomics and metabolomics analyses are currently under intense investigation. Components such as peripheral leukaemia inhibitory factor (LIF) [285], proteins associated with the complement cascade pathway, lipid metabolites [286], and glutathione metabolism [287] are closely associated with immunotherapeutic responses.

However, there are still limitations to the universal clinical application of peripheral components. Factors such as the techniques used for sampling, storage protocols, diagnostic standards, and cost should be considered for further exploration.

Epigenetics and gene variation

Epigenetics modification

Epigenetics alterations, including DNA methylation, histone modification, and regulation of non-coding RNA, significantly modulate gene expression. Hence, distinct epigenetic modifications result in various therapeutic efficacies.

Methylation is the most common epigenetic modification, with specific gene methylation levels significantly impacting immune responses. In GI tract cancers, alterations in m6A regulatory factors have been shown to modulate the infiltration of various TIICs, thereby leading to heterogeneity observed in the efficacy of immunotherapy [246, 288,289,290]. Furthermore, researchers have identified molecular subtypes based on epigenetically regulated gene expression profiles in CRC and GC, which exhibit varying sensitivities to immunotherapy [291, 292]. Notably, methylation levels change with treatment; in a single-arm phase II clinical trial involving mCRC, treatment with pembrolizumab plus azacitidine led to a decrease in global methylation, particularly at promoter sites, enhancing universal gene expression [293]. Beyond methylation, the glycosylation and ubiquitination levels of certain genes have also been shown to influence immunotherapy outcomes in GI tract cancers [294, 295]. Numerous studies have explored non-coding RNA, including lncRNA, miRNA and circRNA, as biomarkers for predicting immunotherapy outcomes. These non-coding RNAs may exert their effects through mechanisms such as modulating m6A regulatory factor expression [296], PD-L1 expression [297], immune cell infiltration [298], and cell proliferation [299, 300].

Genetic variation

In advanced GC, CDH1, JAK2, AXIN1, and PTCH1 mutations have been identified as adverse factors for immunotherapy efficacy [301]. Conversely, in EC, ERBB2 mutation has been negatively associated with neoadjuvant immunotherapy response [302]. POLE/POLD1 mutations in patients with CRC exhibited a better response to immunotherapy, even in patients with MSS CRC, highlighting their strong predictive value [303, 304]. Furthermore, approximately 11–25% of patients with CRC harbour KRAS mutations. However, several analyses (such as KEYNOTE-177 and CheckMate-142) demonstrated that pembrolizumab has low efficacy in patients with mCRC harbouring KRAS/NRAS mutations [7, 139]. Nonetheless, findings from the KEYNOTE-164 study showed that the ORR among patients with mCRC receiving pembrolizumab is similar regardless of RAS status (37.0% vs. 42.0%) [305]. Therefore, further large-scale explorations are warranted.

Given the intricate nature and spatial–temporal heterogeneity of GI tract cancers, the dysfunction of a single gene cannot be applied as a general biomarker. Therefore, it is essential to develop prediction models based on subsets of epigenetic alterations or gene variations. For instance, in GI tract cancers, a study analysed the genomic profiles of 227 patients treated with immunotherapy across multiple centres and devised a GIPS model based on the status of six crucial genes (RNF43, CREBBP, CDKN2A, TP53, SPEN, NOTCH3). GIPS can independently serve as an excellent prognostic factor for immunotherapy [306].

Resistance to immunotherapy

Resistance to immunotherapy is one of the main challenges associated with GI tract cancers owing to the insufficient response to immune checkpoint blockade (ICB)‐based immunotherapy. Even in patients who initially benefited from immunotherapy, 46.4% of them developed acquired resistance, most of which occurred within 24 months [307]. In this section, we discuss the recent research progress. First, we discuss emerging advancements in immune, metabolism, microbiota and epigenetics-mediated immunotherapy resistance mechanisms, mainly in GI tract cancers. Second, we review strategies to overcome immunotherapy resistance by combining them with other regimens and aligning efforts with clinical evidence (Fig. 2).

Fig. 2
figure 2

The summary of immunotherapy resistant mechanisms and overcoming strategies in patients with GI tract cancers. This algorithm provides guidance for developing strategies overcoming resistance in patients with GI tract cancers

Mechanisms of resistance

Suppressed activation of T cells

Loss of PD-L1 expression

Considering the reactive anti-tumour immunity of ICIs by blocking the binding of PD-1 and PD-L1, the loss of PD-L1 expression is a contributing factor to ICI resistance. Genetic alterations, such as mutations or loss of function in the JAK/STAT pathway, can diminish PD-L1 expression in cancer cells, leading to both primary and acquired resistance to anti-PD-1 antibodies [307]. Especially in MSI-H cancers, JAK1 frameshifts (loss of function alterations) were observed in 6% (9/158) and 15% (4/27) of MSI-H colorectal and gastric cancers in the TCGA database, respectively, and exhibited decreased expression of IFN-associated genes [308]. In addition to external signalling alterations, internal modifications in CD274 also contributed to the loss of PD-L1 expression.

Among 13 patients with MSS metastatic CRC possessing high-affinity Fcγ receptor 3a (FcγR3a), 3 developed tumour subclones harbouring PD-L1 mutations selectively. These mutations led to the loss of tumour PD-L1 expression following treatment with the PD-L1 antibody avelumab, either through nonsense-mediated RNA decay in the case of PD-L1 K162fs mutation or protein degradation in the case of PD-L1 L88S mutation [309].

Downregulation of IFN-γ signalling

Alterations in the IFN-γ-JAK/STAT signalling pathway at the epigenetic, transcriptional, posttranscriptional, and posttranslational levels are associated with ICI resistance [310]. In CRC, mutations of JAK1/2 and B2M and lack of the IFN-γ receptor [311, 312] have been identified as contributors to ICI resistance. Research on therapeutic strategies that target the genetic, epigenetic, and metabolic elements regulating IFN-γ signalling is underway.

Expression of immune checkpoints beyond PD-1/PD-L1

The compensatory upregulation of alternative checkpoint proteins, such as lymphocyte-activation gene-3 (LAG-3), is a potential mechanism involved in acquired resistance to anti–PD-L1 therapy [313, 314]; LAG-3 is an immunomodulatory receptor that regulates effector T-cell (Teff) homeostasis, proliferation, activation and the suppressive activity of Treg cells [315]. A pioneering human study demonstrated that a combination therapy of favezelimab, an anti-LAG-3 antibody, plus pembrolizumab improved exploratory efficacy in terms of survival and DOR among patients with MSS mCRC [316]. Klapholz et al. found that patients with MSS CRC displayed immune exhaustion signatures (MSS-ImmEx) characterised by the accumulation of Tim-3 + PD-1 + CD8 + TILs, which exhibited a dysfunctional or “exhausted” phenotype [317]. Therefore, dual blockade of Tim-3 and PD-1 may have additive or synergistic anti-tumour effects on patients with GI. Another immune checkpoint, T-cell immunoglobulin and ITIM domain (TIGIT), is associated with CD8 + T-cell exhaustion. Exhausted T cells develop resistance to anti-PD-1 when completing the exhausted transcriptional program [318, 319]. A study involving patients with CRC showed that anti-TIGIT combined with anti-PD-L1 therapy restores the functions of intratumoural CD4 and CD8 T cells, irrespective of microsatellite status, in either primary tumour or liver metastasis [320]. Moreover, an alkaline phosphatase (ALP)-responsive and transformable supramolecular bis-specific cell engager (Supra-BiCE), consisting of both SA-P (a phosphorylated peptide targeting and blocking PD-L1) and SA-T (a phosphorylated peptide targeting and blocking TIGIT), achieved a tumour suppression rate of 98.27% in a CRC model, providing a promising tool for engaging NK and T-cells for cancer immunotherapy [321].

Loss of antigen

Cancer immunoediting occurs during tumour evolution and in the anti-PD-1 therapy process, leading to the downregulation or loss of target antigen, a common mechanism of resistance to immunotherapies [322,323,324]. Likewise, cancers with low TMB and immunogenicity cancers, such as MSS CRC and pancreatic ductal adenocarcinoma (PDAC), tend to be primary refractory to ICIs [325, 326]. Therapeutic combinations with agonistic antibodies against the CD40 receptor (αCD40) are efficacious in preclinical mouse models, which can rescue and generate new T-cell responses against weak affinity or poorly expressed neoantigens or against tumour-associated self-antigens that lack high-affinity T-cell clones owing to central tolerance. Hence, there is therapeutic potential in combining αCD40 with ICI, particularly for treating MSS CRC and other immunologically cold cancers, especially cancers that developed resistance to immunotherapy through antigen loss [325, 326].

Defective antigen presentation

Patients with MSI-H CRC exhibit a favourable response to ICIs due to the TMB-H and the presence of neoantigens [327,328,329,330]. Contrastingly, frequent mutations in the B2M and HLA-ABC genes cause defects in antigen presentation [328, 331, 332], thus impairing the recognition of tumour cells by cytotoxic CD8 + T cells and resulting in acquired resistance to ICIs [333,334,335]. Currently, CAR-T-based therapy that bypasses self-MHC restriction and targets nonrestricted cell surface antigens is the viable immunotherapeutic option for overcoming genome-level defects during antigen processing and presentation pathway components [336].

Suppressive tumour immune environment

Inadequate T-cell infiltration

The critical function of ICIs is to reinvigorate the exhausted T cells. A notable resistance mechanism in the TME is the absence of T cells, indicating a target-missing situation [337]. Several factors, including the presence of abundant cancer-associated fibroblasts (CAF), regulatory T cells (Tregs) and myeloid-derived suppressor cells (MDSCs) [338], expression of immune checkpoints beyond PD-1/PD-L1 (such as TIM-3) [339], and elevated lactate level, can induce T-cell exhaustion [340]. Additionally, specific genetic mutations and alterations in signalling pathways impact T-cell infiltration in GI tract cancers. For instance, mutations in the phosphatase domain of PTEN, which reduce protein expression, have been linked to the reduced levels of intratumoral CD8 + T cells and resistance to anti-PD-1 therapy in dMMR/MSI-H GI tract cancers [341]. Interestingly, the immunosuppressive effect of PTEN loss may not be through the regulation of the PI3K/AKT pathway but rather related to the release of cytokines and chemokines that promote the proliferation and differentiation of MDSC and M2 macrophages and expansion of Tregs and reduce NK cell infiltration [342,343,344]. Furthermore, WNT signalling activation through CTNNB1 mutation, which encodes β-catenin, is associated with a low T-cell infiltration rate [345].

A retrospective analysis of data from the KEYNOTE 177 trial revealed that the highly cytotoxic CD8 infiltration in hypermutated CRC is facilitated by the low activation of the WNT pathway, resulting in an immune hot environment favourable for immune response [346].

Immunosuppressive cells

MDSCs are one of the primary immunosuppressive factions in the TME, as they directly target and release soluble mediators to regulate immune responses [347,348,349], thereby influencing therapeutic efficacy and the prognosis of patients with GI tract cancers [350,351,352,353,354]. Tsutsumi et al. analysed single-cell RNA sequencing (scRNA-seq) data derived from human GC specimens following anti-PD-1-antibody therapy and proposed that tumour-infiltrating MDSCs, particularly those expressing immediate early response 3 (IER3), play a predominant role in the development of the immunosuppressive and ICI-resistant GC tumour immune microenvironment (TIME) [355]. IER3 is involved in protecting cells from apoptosis [356, 357], which suggests that IER3 + M-MDSCs are less susceptible to apoptosis and subsequently differentiate into immunosuppressive macrophages. Although MDSCs represent promising therapeutic targets, developing MDSC-specific treatments remains challenging.

The proportions of circulating and tumour-infiltrating MDSCs could be reduced by chemotherapy [358, 359], targeted therapy [360, 361], all-trans retinoic acid (ATRA) [362], and blockade of chemokine receptors on MDSCs. MDSCs infiltrate into the cancers induced by various cytokines. Under the extreme conditions within cancers, such as low oxygen levels, high oxidative stress, and nutrient deficiency, the MDSCs undergo functional and differentiation changes, transforming into tumour-associated macrophages (TAMs). TAMs play a significant role in promoting the occurrence and progression of GC and are inversely associated with patient prognosis. In particular, CD204-positive TAMs, M2-polarized macrophages, are a critical risk factor for the progression from gastric adenoma to adenocarcinoma [363,364,365,366]. Existing studies have focused on the “reprogramming” of TAMs from “tumour-support cells” to “tumour-killer cells” [367, 368]. For instance, Cao et al. investigated the reprogramming of TAM through V-domain immunoglobulin suppressor of T-cell activation (VISTA), an immune checkpoint associated with immunotherapeutic resistance, across eight independent cohorts involving a total of 1,403 patients with GC. The blockade of VISTA successfully reprogrammed TAMs to a proinflammatory phenotype, enhanced T cell-mediated antitumour immunity, and improved the efficacy of PD-1 inhibitors [369].

Tregs modulate immune homeostasis and, at the same time, inhibit immune responses in cancer patients. Tregs secrete IFN-γ, which enhances the efficacy of ICIs [370] but produces inhibitory cytokines, such as IL-10 and transforming growth factor-β (TGF-β) [371], which are correlated with ICI resistance [372]. The opposite functions of Tregs can be attributed to the heterogeneity of their subsets [373]. C–C motif chemokine receptor 8 (CCR8) is a marker of activated Tregs expressed on Tregs infiltrated in tumour rather than peripheral Tregs [372].

The inhibition of CCR8 expression is a promising strategy in cancer immunotherapy, specifically for enhancing anti-tumour immunity in colon cancer by modulating tumour-resident Tregs. [374]. Furthermore, LM-108, a monoclonal antibody targeting CCR8, is currently being evaluated for its efficacy in treating patients with advanced solid cancers. The treatment groups involved anti-CCR8 monotherapy or in combination with anti-PD-1 therapy (CTR20221680).

Tumour-associated neutrophil (TAN) populations are associated with unfavourable prognosis in patients with cancer [375]. Single-cell transcriptomic analyses have unveiled the striking cellular heterogeneity of neutrophils under pathological conditions and identified their diverse roles in cancer progression. Specifically, pro-tumour TAN-expressing markers, such as CCL3, CCL4, SPP1, and PD-L1, are promising targets for immunotherapy. These TAN subtypes can potentially be targeted either alone or in combination with ICIs to devise more effective cancer treatment strategies [375, 376].

The role of mesenchymal stem cells (MSCs) in cancer remains controversial, as evidence of both pro- and anti-tumour effects has been reported [377, 378]. A potential explanation for these conflicting observations, as suggested by Cascio et al., may be the origin of MSCs and the degree of “cancer education” [379]. In particular, MSCs from local tissues can be epigenetically reprogrammed by the TME into cancer-associated MSCs (CA-MSCs), which drive tumour immune exclusion and resistance to immunotherapy. In GC, MSCs mediate immunosuppression via the CXCR2/HK2 (Hexokinase II)/PD-L1 pathway, blocking GCMSCs-derived IL-8/CXCR2 pathway can reduce PD-L1 expression and lactate production, thereby improving the anti-tumour efficacy of anti-PD-1 immunotherapy; therefore, it may be a promising target for treating advanced GC [380]. MSCs modulate multiple biological processes in cancer, extending beyond immune modulation to immunotherapeutic resistance. Specifically, the exosomes secreted by MSCs can facilitate resistance to immunotherapy [381,382,383,384].

In conclusion, the intimate crosstalk between cancer cells and their surrounding immunosuppressive cells plays a critical role in cancer progression and resistance to therapies. Understanding these interactions will help advance cancer research and treatment. Future studies may ultimately enhance the efficacy of anti-PD-1 therapy by identifying and targeting a particular cell subset or specific inhibitory molecule that broadly neutralises the effect of anti-PD-1 therapy on GI cancer patients [385, 386].

Secreted immunosuppressive factors and extracellular matrix

Immunosuppressive factors secreted by tumour, immune, and stromal cells significantly contribute to the formation of an immunosuppressive TME [387, 388]. Notably, the TGF-β released by cancer cells and CAFs is pivotal to facilitating tumour immune evasion [389], while that released into the TME acts as a chemoattractant factor for fibroblasts to induce the formation of CAFs [390]. Therefore, the repression of TGF-β signalling is critical to enhancing the efficacy of current and forthcoming immunotherapies, although potential adverse effects should be carefully monitored [391]. The composition of the extracellular matrix within the TME also influences resistance [392, 393], as the stiffened extracellular matrices could impede the infiltration of drugs and immune cells into the tumour and induce disorganised neovascular vessels with low drug-delivery efficiency [394]. The interleukin (IL) family plays an important role in the immune cell signalling of GI tract cancers. High levels of IL-6 promote epithelial-mesenchymal transition (EMT), clonogenicity, and immunosuppressive phenotype of EC cells. IL-10, potentially in conjunction with IL-35, can be derived from Tregs, promoting the exhaustion of CD8 + TILs and impeding anti-tumour immunity [395, 396].

Metabolism-mediated resistance

Reprogramming of cellular metabolism is a hallmark of cancers impacting TME, immune landscape and therapy resistance [397, 398]. Microenvironmental ammonia induces T-cell exhaustion in CRC, while the accumulation of lactate suppresses the activity of cytotoxic T lymphocytes [399]. The metabolic processes of tumour cells contribute to an immunosuppressive TME through various mechanisms, including nutrient competition, hypoxia, and acidity. Immunotherapy reshapes the features of metabolism. For instance, anti-PD-1 therapy restores glucose levels in the TME, thereby facilitating T-cell glycolysis and IFN-γ production [400]. Recent advancements in single-cell technologies and analytical algorithms have enhanced the exploration of immune metabolism. Hartmann et al. uncovered the spatial organisation of metabolic programs in human CRC and revealed that T cells expanding within cancers exhibit distinct metabolic profiles compared with excluded T cells at the tumour-immune boundary [401]. Despite these insights, the efficacy of targeted metabolic therapies in enhancing T-cell function and promoting anti-tumour immune responses remains unproven. Unfortunately, the most promising metabolic therapy targeting IDO1 failed in a phase III clinical trial, even when combined with anti-PD-1 therapy [402]. Contrastingly, beyond ICIs, preclinical studies that incorporated metabolic targeting alongside adoptive transfer protocols of autologous T-cells and oncolytic viruses have shown potential, necessitating further exploration [403].

Epigenetics-mediated resistance

Epigenetic alterations in cancer cells affect the TIME. Sundar et al. showed that epigenetic promoter alterations in GI tract cancers mediate immune editing and resistance to immune checkpoint inhibition by generating 5′ truncated protein isoforms missing immunogenic N-terminal peptides. Moreover, a high alternate promoter burden resulted in an immune-depleted TME and contributed to the resistance to ICIs [404]. Xu et al. found that IL-1β-associated nicotinamide nucleotide transhydrogenase (NNT) acetylation leads to iron-sulphur cluster maintenance and immunotherapy primary resistance, both in advanced and locally advanced GC. The blockage of NNT acetylation by IL-1β neutralisation synergises with anti-PD-1 therapy in vivo [405]. Epigenetic modification can also affect the efficacy of immunotherapy by modifying tumour cells [406]. Current research suggests that aberrant alterations in the activities of histone-modifying enzymes [407] and epigenetic modification by histone deacetylase 8 [408, 409] overexpression play a crucial role in resistance to hepatocellular carcinoma immunotherapy. Epigenetic alterations are potential biomarkers for predicting the efficacy of immunotherapy and promising targets for overcoming ICI resistance. The mechanisms driving the resistance remain to be elucidated.

However, there are heterogeneity among the possible mechanisms of resistance to immunotherapy in different GI cancers or in different histological types. In the previous study of our center, H. pylori infection is a beneficial factor for GC immunotherapy by shaping hot tumor microenvironments. However, in dMMR/MSI-H colorectal adenocarcinoma and ESCC patients, H. pylori adversely affects the efficacy of immunotherapy [104]. Besides, EC, especially ESCC, is an extremely high TMB tumour, comparable to lung cancer and melanoma, generating specific neoantigens [410]. But the incidence of EAC is rapidly rising worldwide [411], which are highly heterogeneous and surrounded by a largely immunosuppressive TME, resulting disparate sensibility to immunotherapy [412]. Therefore, GI cancers from different regions, histological types, molecular subtypes as GC TCGA and CRC CMS, microbial infection status may have disparate TME, sensitivity and resistance to ICB.

Interestingly, we consider that unique microenvironment of GI tract cancer, corresponding therapy strategies and treatment resistance of which may due to the particular molecular subtypes when compared to other systems. Most GCs are immunologically ‘cold’, in comparison, EBV (+) GC represents a unique subgroup of GC which is heavily infiltrated by active T/B cells associated with antitumour immunity, making it more sensitive to ICB. In refractory EBV (+) GC tumours after standard chomoimmunotherapy, LAG-3 is upregulated on exhausted T cells, underscores a new promising immunotherapeutic target for EBV (+) GC [120].

AFPGC and HAS are special and rare subtypes of gastric cancer [413]. ScRNA-seq on HAS tumour showed that cytotoxic CD8 + T cells exhibited remarkable heterogeneity in their functional states, with a vast majority of cells following the trend of activation-coupled exhaustion, exemplified by high expression of activation markers GZMA and IFNG and of exhaustion markers PDCD1 and CTLA-4. This suggests that immunotherapy may have a good therapeutic effect on HAS. In a real-world study, anti-PD-1 plus chemotherapy could benefit AFPGC and HAS patients. Because of the small sample size, it is difficult to analyze the efficacy-related predictors such as PD-L1. However, one patient with high expression of PD-L1 exhibited hyperprogressive disease, suggesting the existence of particular resistant mechanism and necessity of further investigation [414].

A negative association between CLDN18.2 expression and the prognosis of anti-PD-1/PD-L1 therapy was reported in a study. This correlation might be due to the unique tumour microenvironment of CLDN18.2-positive GC, specifically the lack of PD-1/PD-L1-positive lymphocytes in CLDN18.2-positive GC limited its chance to benefit from PD-1/PD-L1 inhibitors, while infiltrating neutrophils may also augment this negative therapeutic response. But CLDN18.2-targeted CAR-T cell therapy may be a promising treatment strategy in CLDN18.2-postive patients because of the non-exhausted CD8 + T cells surrounding tumour cells [415].

Higher expression of PD-L1 has been found in trastuzumab-resistant HER-2-positive cells [416], but development of resistance during anti-HER-2 plus anti-PD-1 therapy remains unclear. Based on liquid biopsy, the presence of PD-L1 + CTCs/CECs and their impact on therapy were explored using longitudinal analyses in patients receiving anti-HER-2 plus anti-PD-1 therapy. Study showed that triploid-PD-L1 + CTCs participated in primary and acquired therapeutic resistance before and after treatment. As for PD-L1 + CECs, intratherapeutically detected multiploid PD-L1 + CECs demonstrated a superior clinical response, when triploidy and tetraploidy contributed to acquired resistance [282].

Efforts to overcome resistance to immunotherapy

While efforts have been made to classify the resistance mechanisms of immunotherapy into distinct categories, resistance is a complex and dynamic process with various interrelated mechanisms in the real world [385]. The most promising strategy appears to reverse resistance through strategic combination with other treatments [417] (Table 7). Overcoming resistance to immunotherapy is accomplished by understanding the specific resistance mechanisms rather than relying on arbitrary combinations.

Table 7 Summary of the clinical trials of anti-PD-1/L1 combined with other agents in GI tract cancers

Combination with chemotherapy

Chemotherapy increases the TMB through DNA damage, subsequently enhancing antigen presentation, eliminates immunosuppressive cells, and enhances the function of effector cells, thereby reprograming the TME and augmenting the immune response [418, 419]. The combination of chemotherapy with anti-PD-1/PD-L1 is a standard-of-care option for GI owing to the synergistic effects.

Combination with radiotherapy

Radiotherapy eliminates local lesions, stimulates the systemic antitumor immune response (also known as abscopal effects) [420], and synergises with anti-PD-1/PD-L1 therapy. Furthermore, radiotherapy promotes T-cell infiltration and expands T-cell receptor (TCR) repertoire [421], upregulates PD-L1 on tumour cells targeted by anti-PD-1/PD-L1 [422], increases MHC-I expression on tumour cells, and alleviates resistance to anti-PD-1/PD-L1 [423]. These promising pre-clinical attributes have been substantiated in prospective studies. A phase II trial (NCT03104439) that combined radiation with ipilimumab and nivolumab to treat primary immunotherapy-resistant cancers as MSS CRC and PDAC demonstrated notable clinical benefits and prolonged disease control [424]. Furthermore, as a local therapy and an immunomodulator, radiotherapy can also synergise with immunotherapy in oligometastatic ESCC patients who have either failed previous immunotherapy or acquired immunotherapy resistance. This combination has shown responses even in unirradiated lesions [425]. By reprogramming the TME in cancers with minimal immune infiltration, radiotherapy, combined with immunotherapy, induces a potent mobilisation of both innate and adaptive immunity [426]. However, the optimal combination strategies, including precise timing, optimal dose, fractionation schedule and target sites, remain to be determined [425].

Combination with anti-angiogenesis therapy

Combined antiangiogenic and anti-PD-1/PD-L1 therapy has demonstrated a synergistic effect [427, 428] in enhancing antitumor immunity by inducing vascular normalisation [429]. Following the promising findings from the REGONIVO study, the combination of regorafenib, a VEGFR inhibitor, and PD-1 inhibitors has been considered as a treatment for refractory pMMR/MSS mCRC patients, particularly those with CPS < 1 and low TMB [430]. Nevertheless, patients with liver metastasis could not benefit from this treatment strategy, which necessitates further investigations. A combination of VEGFR inhibitors can potentially extend the applicability of PD-L1/PD-1 inhibitors beyond patients with dMMR/MSI-H mCRC. Moreover, patients with AFPGC, which is characterised by high invasiveness, early metastasis, rapid progression [431], and elevated VEGF-C expression [432], may benefit from targeting the VEGF-C-VEGFR2 pathway. Recent data presented at the 2024 ASCO meeting highlighted the efficacy of apatinib, a VEGFR2 tyrosine kinase inhibitor, combined with anti-PD-1 and chemotherapy in patients with AFPGC (NCT04609176). Encouragingly, the study reported an ORR of 55.6%, a DCR of 86.1%, and an improved prognosis with a 1-year PFS of 42.1% and a 1-year OS of 63%.

Combination with targeted therapy beyond VEGF

Intracellular signal transduction pathways that mediate resistance to immunotherapy are potential targets for combination therapy. TGF-β not only suppresses the immune response but also promotes angiogenesis and activates CAFs. M7824 (also known as Bintrafusp alfa) is an anti-PD-L1/TGF-β receptor II (TGF-βRII) trap agent [433]. Despite the limited efficacy of first-line treatment for post-platinum biliary tract cancer patients [434], M7824 has shown remarkable results in post-chemotherapy EAC [434] and ESCC [435], with ORR of 83% and 100% in patients with immune-excluded phenotype, respectively. Similarly, another dual PD-L1/TGF-β inhibitor, Retlirafusp alfa (SHR-1701), is undergoing evaluation in phase III studies in patients with G/GEJC.

Moreover, DNA damage repair defects may enhance cancer sensitivity to ICI treatment [189]. Therefore, the combination of ICIs and DNA-damaging therapy could theoretically reduce immunotherapy resistance and improve efficacy [436]. Ceralasertib (AZD6738), an ataxia telangiectasia and Rad3-related protein kinase inhibitor, plus durvalumab displayed promising efficacy with an ORR of 22.6% and a DCR of 58.1% in patients with refractory advanced GC. Accompanying translational research during the treatment highlighted the activation of both innate and adaptive immune responses in responders [437].

Combination with other ICIs

Dual blockade of two checkpoints could potently reduce the probability of resistance and has been demonstrated in abundant preclinical studies [438]. Anti-PD-1 plus anti-CTLA-4 has achieved a prolonged mOS of 6.6 months in patients with “cold” cancers, such as pMMR/MSS CRC, but without ORR and significant improvement in PFS. Further subgroup analyses indicated that patients with TMB > 28 and those classified as CMS2 may benefit from the therapy [439]. Cadonilimab (AK104), an anti-PD-1/CTLA-4 bispecific antibody, showed an encouraging tumour response rate with a manageable safety profile in ESCC and HCC [440]. A phase III study of cadonilimab in first-line treatment of G/GEJ adenocarcinoma is underway (NCT05008783). Moreover, a phase I first-in-human study evaluating the activity of FS118, a bispecific antibody targeting LAG-3 and PD-L1, showed clinical benefit in 43 patients (including 5 patients with CRC) resistant to anti-PD-(L)1-based therapy earlier. SD was also observed in patients with previous ICI as their most recent therapy and/or co-expression of LAG-3 and PD-L1. An increase was observed in the counts of CD4 + and CD8 + T cells following treatment in patients with SD (NCT03440437) [441].

Combination with immune modulators

Combining immune modulators with PD-(L)1 targeting agents is a promising strategy for patients with PD-(L)1–refractory disease [442]. In addition to co-inhibitory pathways such as PD-1 and CTLA-4, co-stimulatory pathways, including CD40/CD40L, CD27/CD70, 4-1BB/4-1BBL, GITR/GITRL, and ICOS/ICOSL, are crucial in regulating T-cell function [443]. Agonists targeting co-stimulatory pathways have demonstrated the potential to boost T-cell activity and elicit antitumor immune responses [444]. The TNF receptor superfamily member 9 (CD137 or 4-1BB) is an inducible T-cell costimulatory receptor expressed on activated CD4 + , CD8 + T cells, and NK cells [442]. In a phase I trial involving 12 (19.7%) CRC and 6 (9.8%) PDAC patients, a bispecific antibody targeting PD-L1 and 4-1BB (GEN1046) achieved a DCR of 65.6%. Notably, two patients who had previously progressed on anti-PD-(L)1 therapy achieved PR. IL-17 affects the infiltration and exhaustion of immune cells that contribute to the formation of an immunosuppressive microenvironment [445, 446]; therefore, targeting IL-17 is a promising strategy for overcoming immune suppression and enhancing the sensitivity of anti-PD-1 therapy [447, 448]. The synergistic effect of anti-IL-17 and anti-PD-1 is being verified in several ongoing clinical trials (NCT05061017).

Combination with faecal microbiota transplantation

Findings have proved the efficacy of responder-derived faecal microbiota transplantation (FMT) in reversing resistance to ICIs in patients with melanoma, induced an ORR up to 30% after anti-PD-1 was restarted [449, 450]. Responding patients had an increased frequency of activated dendritic cells, type I interferon signalling and CD8 + T cells in the TME. In GI tract cancers, the attempts of FMT are concentrated in colorectal cancers. In a small sample size, phase II trial (RENMIN-215), responder-derived FMT plus anti-PD-1 and fruquintinib as third-line or above treatment showed inspiring mPFS with 9.6 months in refractory MSS mCRC. Peripheral blood expanded TCRs exhibited the characteristics of antigen-driven responses in responders [451]. There are a few points need to discuss, first, FM is usually derived from the responders, but an additional study found that FMT from untreated healthy donors led to an ORR of 65% in ICB-naive patients with melanoma [452]. Secondly, mechanisms by which FMT improves efficacy or overcome resistance. Based on the elevated DC in melanoma and the expanded TCRs in CRC, we hypothesized that FMT might promote the antigen presentation process. Besides, dosage forms, administration timing et al. warrant further study yet.

New attempts to overcome immunotherapy resistance

Treatment strategy following the true progression of immunotherapy is yet to be elucidated, given the limited evidence currently available. In a study of acquired immunotherapy resistance in GI cancer, chemotherapy was the main post-resistant regimen (23.7%), followed by maintaining the original immunotherapy (12.7%) [307]. In addition to combination strategies to overcome immunotherapy resistance, targeted intervention of the key links of immunotherapy tolerance is another potentially effective strategy. To date, this attempt has mainly included CAR-modified cell therapies, herbal medicines, monomeric drugs, ovs, and other biological agents.

Novel immunotherapy for GI tract cancers

Although immunotherapy of GI has made breakthroughs, the conventional strategies focused on immune cells alone, especially immune checkpoints, have limited benefits. Therefore, the development of novel immunotherapy has explored the whole process of anti-tumour response. First, the transfusion of expanded immune cells has shown great potential in GI tract cancers recently. Second, tumour antigens must be presented to T cells by antigen-presenting cells (APCs) so that they can be recognised by T cells, which makes cancer vaccines an attractive strategy to elicit anti-tumour response. Third, T cells are inhibited by immune checkpoints in cancers. Novel ICIs to rescue the second signal, which is essential to T-cell activation, remains to be identified. Furthermore, oncolytic viruses killing tumour cells in versatile ways, which is far beyond directly lysing tumour cells (Fig. 3).

Fig. 3
figure 3

Novel strategies of GI immunotherapy. We summarized four directions of the development of novel immunotherapy including 1. Transfusion of immune cells, including T cells and innate immune cells, which are edited by gene engineering to recognize tumour antigen such as HER-2, CLDN18.2, and exert potent killing effect;2. Novel immune checkpoint inhibitors can rescue cytotoxic T cells from inhibitory signalling induced by the combination of legend with PD-1, LAG3, TIGIT;3. cancer vaccines, including peptide vaccine and dendritic cells vaccine, can present tumour antigen peptides to activate T cells; 4. Novel tumour virus not only lyses tumour cells, but also provide a versatile platform to kill cancer cells, such as encoding PD-L1 antibody

Transfusion of killer cells

T cells beyond classic cytotoxic T cells

One form of adoptive immunotherapy is the use of viruses to introduce genes encoding novel types of receptors, known as CAR, into the T cells of patients. T-cells expressing CAR have made breakthroughs in the treatment of GI tract cancers.

Qi et al. reported the interim analysis of a phase I clinical trial of Claudin18.2 (CLDN18.2)-targeted CAR-T cells (CT041) in patients with previously treated CLDN18.2-positive digestive system cancers (NCT03874897). The author reported promising efficacy against CLDN18.2-positive digestive cancers, particularly in GC [24]. In another study, CT041 was administered to two patients with metastatic pancreatic cancer. They responded remarkably to CT041. One of the patients achieved partial regression (PR), whereas the other achieved CR of lung metastasis [453].

However, autologous CAR-T cell therapies necessitate an extended period prior to transfusion due to manufacturing time. In contrast, allogeneic donor-derived CAR-T cells, which can be banked for immediate use, address the critical temporal limitation. CYAD-101, an allogeneic CAR-T using a non-gene edited, peptide-based technology to mitigate graft versus host disease (GvHD) is combined with a NKG2D-based CAR. Pre-clinical studies have confirmed that CYAD-101 could maintain CAR-directed anti-tumour activity without inducing GvHD [454]. Clinical grade CYAD-101 cells were produced for the phase I trial (NCT03692429) enrolling fifteen patients with refractory mCRC patients, among whom, two patients achieved PR, nine patients achieved stable disease (SD), and the mPFS was 3.9 months [455]. Based on the encouraging results, KEYNOTE-B79, a phase Ib clinical tiral (NCT04991948) conducted in MSS/pMMR mCRC patients with CYAD-101 is under investigation.

Another challenge faced by CAR-T cell therapy is the limited number of safely targetable cancer antigens on the cell surface. Several tumour antigens, including neoantigens and shared antigens, are derived from intracellular proteins. Therefore, the development of TCR-T cell targeting intracellular antigens may be a prospective treatment option. KRAS is one of the most frequently mutated proto-oncogenes in human cancers. The prevailing oncogenic mutations observed in KRAS involve single amino acid substitutions at codon 12, specifically G12D and G12V. These mutations are highly prevalent, accounting for approximately 60% to 70% of PDAC and 20% to 30% of CRC. Neoantigens related to KRAS mutation can elicit a strong immune response, making it a potential target for TCR. Wang et al. employed HLA-peptide prediction algorithms and determined the potential of HLA-A*11:01 to present mutated variants of KRAS [456]. The researchers extracted murine T cells and subsequently isolated TCR specific to the mutated KRAS variants G12V and G12D. Peripheral blood lymphocytes (PBL), upon transduction with these specific TCR, exhibited a remarkable ability to discern multiple HLA-A*11:01( +) tumour cell lines harbouring the KRAS G12D and G12V. Adoptive transfusion of these transduced PBLs showed potent killing effects of G12D-mutated pancreatic cancer in vivo, which facilitated relevant clinical trials.

A phase I/II study (NCT03745326) involving the administration of PBLs transduced with a murine TCR recognising the G12D variant of mutated RAS in HLA-A*11:01 patients is currently recruiting at National Cancer Institute. Similarly, Chen et al. evaluated the preliminary efficacy and safety of HLA-A*11:01 KRAS G12V, G12D, and G12C mutant antigen-specific TCR-T cells in the treatment of patients with advanced pancreatic cancer, lung cancer, and colorectal cancer (ChiCTR2200057171).

Innate immune cells

Innate immune cells are responsible for nonspecific immune responses, whereas the adaptive anti-tumour response is T-cell-dependent. Although cytotoxic T-cells are the most important killer cells in fighting against cancer, innate immune cells play an irreplaceable role in anti-tumour processes. Mesothelin (MSLN) is an immunotherapeutic target in GC. Cao et al. constructed anti-MSLN CAR NK cells which could specifically kill MSLN-positive GC cells in vitro and in vivo, showing excellent potential for clinic application [457].

Recently, rapid progress has been made in adoptive cell therapy using macrophages as effector cells owing to their phagocytotic, antigen presentation, and high penetration capabilities. Dong et al. generated a novel CAR-Macrophage (CAR-M) based on genetically modified human peritoneal macrophages (PMs) expressing a HER-2-FcεR1γ-CAR (HF-CAR). The researchers observed that HF-CAR-PMs specifically targeted the HER-2-expressing GC cells and that intraperitoneal administration of HF-CAR-PMs significantly facilitated HER-2-positive tumour regression in a peritoneal carcinomatosis (PC) mouse model and prolonged the OS [458]. However, the clinical trial (NCT06224738) has not yet been conducted.

Cancer vaccines

T cells exert killing effects following induction by APCs carrying tumour antigen peptides, which is the first signal of T-cell activation. Cacner vaccines include cancer peptide vaccines and cell vaccines. Peptide-based cancer vaccines typically consist of 20–30 amino acids containing specific epitopes from highly immunogenic antigens to induce the desired immune response by stimulating specific immune responses against cancer cells. Dendritic cells (DCs) are APCs with a unique ability to induce primary immune response. DCs play an important role in adaptive immunity. Neoantigen-based DC vaccines present tumour neoantigens to naïve cytotoxic T cells and stimulate potent immune response. Taken together, therapeutic vaccines hold promise for providing long-term clinical benefits to patients with cancer.

Peptide-based cancer vaccines

EBV latent proteins are expressed in multiple EBV-associated cancers, play a significant role in carcinogenesis and thus represent vital therapeutic targets for these malignancies. Zhao et.al developed mRNA-based therapeutic vaccines designed to express the T-cell-epitope-rich domain of truncated EBV latent proteins. These vaccines effectively activated both cellular and humoral immunity in tumour-bearing mice, leading to suppressed tumour progression [459]. Peng et. al conducted a clinical trial (NCT05714748) to investigate the efficacy of EBV mRNA vaccine against tumours. What’s more, wGc-043 as a mRNA cancer vaccine, recently received FDA approval for clinical trials which was the first granted EBV related mRNA therapeutic cancer vaccine, representing a milestone advancement in the research of future cancer treatments.

An ongoing phase III study (NCT03639714) is assessing the safety, tolerability, and recommended phase II dose of an individualised heterologous chimpanzee adenovirus (ChAd68) and self-amplifying mRNA-based neoantigen vaccine in combination with nivolumab and ipilimumab in patients with advanced metastatic solid cancers. The individualised vaccine regimens were well tolerated, with no dose-limiting toxicities. In addition, the vaccine induces long-lasting neoantigen-specific CD8 + T-cell responses. Despite limitations due to the small study size, the observed increase in OS in MSS CRC warrants further exploration in large-scale randomized studies [460].

However, challenges remain before the vaccine can be widely applied in clinical practice owing to their high level of individualisation. In an ongoing phase I/II study (NCT03953235), Rappaport et al. constructed a therapeutic vaccine encoding 20 shared neoantigens derived from selected common oncogenic driver mutations. The vaccine was administered in combination with the ICIs ipilimumab and nivolumab to patients with advanced/metastatic solid cancers expressing one of the human leukocyte antigen-matched tumour mutations included in the vaccine. Almost all patients (18/19) harboured KRAS mutations. Unfortunately, the ORR was 0%, and the mPFS and OS were 1.9 and 7.9 months, respectively. A notable preference was observed for TP53 neoantigens encoded in the vaccine, suggesting an unidentified immunodominance hierarchy which might influence the efficacy of multi-epitope-shared neoantigen vaccines. Therefore, a more effective vaccination specifically targeting KRAS-derived neoantigens is currently under development and assesment in a subset of patients in a phase II trial [461].

Neoantigen-based dendritic cell vaccines

Tumour antigens include tumour-associated antigens (TAAs) and tumour-specific antigens (TSAs). As they are expressed in both tumour and normal cells, TAAs rarely provoke evident cellular immune responses. In contrast, TSAs exclusive to cancer cells can trigger a robust immune response. Tumour antigens must be captured by DCs and cross-presented to CD8 + T cells to initiate their activation. Subsequently, antigens must be directly presented to tumour cells for recognition by cytotoxic T-cells. Therefore, neoantigen-based DC vaccines hold the potential to induce robust anti-tumour immune responses, with the efficacy and safety confirmed.

Guo et al. documented a case in which a patient with metastatic GC who received a personalised neoantigen-loaded monocyte-derived DC (Neo-MoDC) vaccine followed by combination therapy of the Neo-MoDC and an ICI. The patient developed T-cell responses against neoantigens after receiving the Neo-MoDC vaccine alone. Subsequent combination therapy triggered a stronger immune response and mediated the complete regression of all cancers for over 25 months. Peripheral blood mononuclear cells recognised most of the vaccine neoantigens. The frequency of neoantigen-specific T-cell clones significantly increased post-vaccination (NCT03185429) [462].

Whole-cell vaccines

The whole-cell cancer vaccine, which incorporates a comprehensive array of TSAs, shows great potential in preventing tumour development, progression, and recurrence. Its primary function is to provoke the immune system into identifying and eradicating tumour cells instead of cytotoxic effects on the tumour cells. Researchers have developed a whole-cell tumour vaccine named GVAX. This allogeneic vaccine is modified to generate granulocyte–macrophage colony-stimulating factor (GM-CSF), crucial for triggering immune responses. A phase II study (NCT02981524) evaluated the combination of GVAX, cyclophosphamide, and pembrolizumab in patients with advanced pMMR CRC. Results indicated that the combination of GVAX/cyclophosphamide and pembrolizumab is well tolerated and induces carcinoembryonic antigen (CEA) decrease, without radiographic changes. In this small cohort, PFS and OS rates appeared favourable compared to controls. Notably, CEA responses were absent in the anti-PD-1 monotherapy group, suggesting GVAX's potential to enhance the anti-tumour immune response.

Chen et al. employed the CRISPR-Cas9 system to disable the interferon-β (IFN-β) specific receptor in live tumour cells, subsequently engineering them to produce IFN-β. This strategy effectively inhibited tumour growth and angiogenesis. Furthermore, these engineered tumour cells were manipulated to express GM-CSF to modulate immune responses. This bifunctional tumour cell vaccine directly induced caspase-mediated tumour cell apoptosis while activating and sustaining long-term immunity. Nonetheless, the safety of using living tumour cell vaccines remains under verification, necessitating further evidence and precautions against potential secondary tumours [463].

Tumour cell-derived cancer nano vaccines

Tumour cell-derived cancer nano vaccines introduce tumour cell-derived components as functional units that endow the nano vaccine systems with antitumor function. Liang et. al generated an endoplasmic reticulum stress inducer α-mangostin (αM) into tumour cells and harvested biologically self-assembled tumour cell-derived cancer nano vaccines (αM-Exos) based on the biological process of tumour cell exocytosing nanoparticles through exosomes. Following subcutaneous injection, αM-Exos efficiently migrated to lymphonodes and was expeditiously endocytosed by DCs, delivering tumour antigens and adjuvants to DCs synchronously, which then powerfully triggered antitumor immune responses and established long-term immune memory [464].

ICIs beyond PD-1/PD-L1

T-cell immunoreceptors with Ig and ITIM domains

T-cell immunoreceptors with Ig and ITIM domains (TIGIT) are immunosuppressive receptors expressed on immune cells, predominantly found on the surface of T- and NK cells. TIGIT is significantly overexpressed in tumour-infiltrating lymphocytes across various malignancies [465, 466]. The ligands CD155, CD112, and CD113 are associated with TIGIT, with CD155 and CD112 being notably overexpressed in many tumour types. Zhu et al. observed that the TAMs from CRC showed robustly higher expression of CD155 than the macrophages from adjacent normal tissues. TAM-specific CD155 contributes to M2-phenotype transition, immunosuppression, and progression [467]. CD155/TIGIT signalling regulates CD8 + T-cell metabolism, promoting the progression of GC [468]. TIGIT exerts its immunosuppressive function by competitively binding to CD155, which is also a ligand of CD226, a co-stimulatory receptor on T cells [469]. Specifically, the novel ligand Nectin-4 (PRR4 and PVRL4) exclusively interacts with TIGIT, thereby inactivating NK cells [470].

Domvanalimab, an Fc-silenced IgG1 monoclonal antibody targeting TIGIT, blocks the interaction between CD155 and TIGIT, so that CD155 in turn binds to the CD226 protein and rescues immune activation signalling. An ongoing global multi-arm EDGE-Gastric study (NCT05329766) investigates the safety and efficacy of the combination of domvanalimab (D) and zimberelimab (Z) in patients with locally advanced unresectable or metastatic G/GEJC. As the primary results revealed at 2023 ASCO, the ORR and 6-month PFS rate in the ITT patients was 59% and 77% respectively. Notably, patients with high-PD-L1 expression(TAP ≥ 5%) exhibited superior response (ORR:80% vs.46%;6 m-PFS rate in ITT:93% vs.68%). The regimen was well-tolerated, with a comparable AEs profile to that of anti-PD-1 + FOLFOX. In conclusion, the D + Z + FOLFOX regimen showed encouraging ORR and 6 months-PFS rates, particularly in patients with high PD-L1 expression(TAP ≥ 5%). STAR-221,a randomised phase III first-line clinical trial comparing D + Z + chemotherapy with Nivolumab + chemotherapy is currently in progress [471].

B7-H3

B7-H3, a notable target for cancer owing to its restricted expression in normal cells, is typically overexpressed in multiple solid cancers [472,473,474,475,476]. In addition, B7-H3 participates in the metabolism, migration, invasion, and endothelial-to-mesenchymal transition of cancer cells [477,478,479,480,481]. Deregulated B7-H3 expression is closely associated with worse outcomes of GI cancers [482]. In addition, emerging evidence points to an immune-evasive phenotype due to B7-H3 overexpression, such as inhibition of CD4 + and CD8 + T-cell activation and proliferation, reduction in IL-2 and IFN-γ production [483, 484], and promotion of tumour immune evasion [485].

Zekri et al. generated B7-H3xCD3 bispecific antibody (bsAb) that showed superior tumour cell elimination, enhanced T-cell activation, proliferation, and memory formation in vitro and in vivo [486]. Unfortunately, the clinical trial of B7-H3xCD3 (NCT02628535) was terminated.

LAG-3

Recently, Kelly et al. reported a phase Ib trial (NCT03044613) that evaluated the efficacy of neoadjuvant nivolumab (Arm A, n = 16) or nivolumab-relatlimab (Arm B, n = 16) in combination with chemoradiotherapy CRT in 32 patients with resectable stage II/stage III GEJC. The results showed overall 2-year RFS and OS rates of 72.5% and 82.6%, respectively. Baseline PD-L1 and LAG-3 expression were positively associated with pathological responses. These findings provide valuable insights into the safety profile and promising efficacy of combining PD-1 and LAG-3 inhibitors in neoadjuvant immune therapy for GEJC [487, 488].

MICA/B

NKG2D is an activation receptor expressed on natural killer (NK) cells, natural killer T (NKT) cells, γδT cells, and naïve CD8 + T cells that plays a vital role in the killing of tumour cells. Eight different NKG2D ligands (NKG2DL), including the MICA/MICB (MICA/B) and ULBP1–6 proteins, have been described in humans [489]. NKG2D/NKG2DL axis is important for tumour immunity, as NKG2DL are upregulated by DNA damage and cGAS-STING signalling, which are rarely observed in healthy cells [490, 491]. Evidence suggests that NKG2DL is frequently detected in malignant cancers. Generally, upregulation of NKG2DLs results in the tagging of stressed cells for elimination by cytotoxic lymphocytes. However, proteolytic shedding of MICA/B on tumour cells results in immune escape [492, 493]. Evidence shows that deficient signal of NKG2D leads to an increased susceptibility to spontaneous tumour development and progression in mice models [494].

As illustrated above, MICA/B is a suitable target for cancer therapy. Due to its complicated mechanisms in immune evasion, MICA/B monoclonal antibody (mAb) is another strategy, besides cell therapy, targeting the NKG2D/NKG2DL axis. Treating human cancer cell lines with MICA/B mAb substantially increases the surface density of these NKG2DLs and induces their killing by human NK cells [495]. Capuano et al. illustrated that the CD16 receptor on NK cells could further enhance the therapeutic activity of MICA/B mAb by inducing NK-cell activation through both NKG2D and CD16 receptors [496], while Courau et al. observed that MICA/B mAb enhanced the destruction of CRC tumour spheroid by increasing NK-cell infiltration and activation. NKG2A expression was increased after anti-MICA/B treatment, and the combination of anti-MICA/B and anti-NKG2A was synergistic [497]. A phase I dose-escalation study (NCT05117476) investigating the safety and efficacy of CLN-619 (anti-MICA/B antibody) alone and in combination with pembrolizumab in patients with advanced solid cancers is currently recruiting.

Novel immune modulators

CD8 + T cells require a third signal, along with the first signal and co-stimulating signal, to generate an active response and avoid death and/or tolerance induction. IL-12 and Type I IFN (IFNα/β) are the predominant contributors to the third signal in various responses. Priming CD8 + T-cells in the absence of IL-12 renders them unresponsive to the same antigen [498]. Curtsinger et al. illustrated that the third signal regulates the CD8 + T cells by promoting chromatin remodelling to maintain the transcription of numerous genes needed for differentiation and effector functions [499].

Razak et al. administered the anti-colony-stimulating factor 1 receptor (anti-CSF1R) monoclonal antibody AMG 820 in combination with pembrolizumab to pMMR patients with refractory in CRC and PDAC. The primary endpoints were the incidence of dose-limiting toxicities and AEs and the ORR per immune-related response evaluation criteria in solid cancers at the recommended combination dose. Although pharmacodynamic effects were observed, the anti-tumour activity was insufficient for further evaluation of this combination in larger patient populations (NCT02713529) [500].

Novel oncolytic viruses

Oncolytic viruses (Ovs) are natural or modified viruses that effectively and selectively infect and lyse cancer cells. Natural viruses have limited tumour specificity. With the advancement of gene engineering, modified Ovs can attack cancer cells with high selectivity and versatility. Ovs directly eliminate cancer cells through their cytocidal effects and activate the immune system directly by stimulating immune cells and indirectly by releasing tumour antigens from dead cancer cells. A phase II clinical trial illustrated that the oncolytic H-1 parvovirus significantly activated the immune system with excellent tolerability in patients with metastatic PDAC [501].

Additionally, novel Ovs reshape the TME through anti-angiogenesis, metabolic reprogramming, and decomposition of the extracellular matrix. Novel Ovs have been adopted as delivery systems for other anti-tumour drugs. LOAd703 is an oncolytic adenovirus loaded with a cytomegalovirus-driven transgene cassette encoding CD40L and 4-1BBL (manufactured by Baylor College of Medicine, Houston, TX, USA on behalf of Lokon Pharma). Preclinical models have illustrated that LOAd703-mediated oncolysis is cancer-selective and that LOAd703 can infect adjacent immune and stromal cells, leading to the expression of CD40L and 4-1BBL and the secretion of chemokines [502]. A non-randomised, single-centre, phase I/II study conducted by the same team combining LOAd703 with chemotherapy in patients with advanced PDAC (LOKON001) concluded that this regimen is feasible and safe. Arm 2 of this trial, which combines LOAd703, chemotherapy, and ICIs, is ongoing [503]. VG161 is the first recombinant oncolytic herpes simplex virus type 1 that carries multiple synergistic anti-tumour immunomodulatory factors (IL12, IL15/15RA, and PD-L1-blocking peptide). VG161 can systematically activate acquired and innate immunity in PDAC models and remodel the TIME, indicating strong anti-tumour potential. The anti-tumour effects and safety of VG161 require further investigation prior to clinical application [504].

Although novel immunotherapy has demonstrated substantial outcomes in preclinical studies, as indicated in Table 8, it is hindered by prolonged research duration and high resource consumption. Furthermore, despite some drugs advancing to early clinical trials, challenges such as high toxicity, poor efficacy, and elevated failure rates persist. Consequently, most novel immunotherapies remain confined to the preclinical and early clinical trial phases. Moreover, the majority of current studies predominantly designed for solid cancers, including lung, breast, and prostate cancer, with few novel methods developed based on the characteristics of GI cancers. Therefore, developing efficient novel immunotherapy strategies with a high clinical translation rate, tailored to GI features, within a truncated timeframe, represents a critical challenge that must be addressed in future research.

Table 8 Summary of under-investigation novel immunotherapies involving GI tract cancers

Challenges and Future directions

The immunotherapy of GI tract cancers has made monumental progress in recent years and the efficacy is encouraging and promising. However, opportunities are always accompanied by challenges in realising precise and individual immunotherapy for GI tract cancers.

Despite numerous explorations on novel immune related targets in GI tract cancers, there are still few druggable targets available. Digestive tract cancers including EC, G/GEJC and CRC, are hollow organ cancers, characterized by significant genetic and molecular spatiotemporal heterogeneity and complexity, thus identification of universal targets is extremely limited. Mutations of targets and activation of alternative signalling pathways commonly occur in GI tract cancers which remarkably impair the efficacy. Besides, TME in GI tract cancers tends to be highly suppressive making it challenging to find targets that reliably evoke anti-tumour immunity. In addition, the absence of organ-specific and tumour-specific antigens resulting in on-target, off-tumour toxicities, further limiting the clinical application of targets. Therefore, to improve the success rates of targets transformation, integrating multi-omics massive data for target mining and employing preclinical research models that more accurately mirror the overall characteristics of the human TME for validation, represent the pivotal research directions for future exploration of new targets.

Progress has been made in identifying immunotherapeutic markers for GI tract cancers, with PD-L1 and MSI/dMMR serving as primary biomarkers for EC, G/GEJC and CRC. While substantial potential beneficiaries are still unidentified, there is still a long journey ahead in selecting patients precisely. Relying on a single biomarker fails to stratify patients accurately and precisely. In recent years, research on GI immunotherapy biomarkers has expanded across multiple dimensions with attempts to construct predictive models for outcomes and AEs that integrate multi-omics data using AI techniques. Moreover, current research utilizing baseline specimens does not adequately capture the transformative effects of immunotherapy on the TME. Therefore, incorporating specimens collected post-therapy for dynamic detection represents a trend in future biomarker exploration necessitating an increase in specimen collection frequency. Develop minimally invasive sampling methods and multi-omics detection using trace specimens techniques, such as nuclear marker-labelled PET-CT, liquid biopsy, and gut microbiota analysis, that reflect the TME from multi-dimensions are urgently anticipated.

Additionally, the integration of clinical trials and translational research with high efficiency and quality is an important project worthy of further exploration. As shown in Fig. 4, “Dynamic-Recycle-Closed loop” research model is recommended. A biomarker-driven clinical trial involves patients with biomarker-positive expression, therefore, optimising efficacy and reducing the waste of medical resources. Moreover, the results from the Rationale 305 trial, which focused on OS in the PD-L1-positive population (tumour area positivity ≥ 5) as the primary endpoint, suggest that tislelizumab combined with chemotherapy leads to statistically and clinically significant OS improvements, with the final results highly anticipated [87]. Developing new treatments based on resistance mechanisms, integrated analysis of datasets, and multi-omics information of specimens will contribute to the rapid mutual transformation between clinical and laboratory data. Patients will be more precisely distinguished and ultimately enrolled in corresponding clinical trials. All research in the model is closely connected, time-transformed, and mutually optimised to form a closed loop.

Fig. 4
figure 4

Dynamic-Recycle-Closed loop research model. We designed a novel research pattern especially for the precise and individual immunotherapy in GI tract cancers. Three parts “biomarker-driven clinical trials”, “exploration of new treatments driven by resistance mechanisms” and “diversified and full-cycle in formation analysis for target mining” are mutually linked, and formed a recycle closed loop. The final goal is achieving “precise and efficient immunotherapy”

In the era of precision and individual immunotherapy, the comprehensive management mode for patients with GI tract cancers is undergoing revolutionary changes. Traditional and classic treatment modalities and management are facing challenges. In the context of significantly increased rates of successful transformation surgery for patients with locally advanced GI tract cancers undergoing neoadjuvant immunotherapy, issues that still need to be explored include: (1) whether it is possible for patients with rectal cancer to achieve clinical CR and further adopting “Watch and Wait” treatment strategy to avoid surgery-related adverse; (2) for patients with EC or G/GEJC benefit to neoadjuvant immunotherapy, it is necessary to evaluate whether extensive lymph nodes dissection is still required and whether the scope of surgery can be minimized to preserve organ function. Additionally, given the notable improvement in long-term survival rates in patients with advanced GI tract cancers receiving immunotherapy, it is crucial to integrating multidisciplinary wisdom for future development. Establish joint decision-making platforms for physicians and patients based on Chat GPT 4.0 and utilize intelligent techniques to realize long-term AE management and follow-up of survival. Moreover, leveraging technologies such as Digital Twins in the intelligent design of clinical trials advocates for conducting prospective multi-cohort clinical studies more efficiently, cost-effectively, and with minimal patient involvement while maintaining the reliability of outcomes, which represents a promising direction for future development.

Conclusion

Collectively, immunotherapy has pioneered a new chapter in the treatment of GI tract cancers. However, significant challenges, including limited treatment options, undefined beneficiary groups, complex drug resistance mechanisms, and low success rates of new drug research and development, remain in achieving precise immunotherapy. We anticipate breakthroughs in basic, translational, and early clinical research on new drugs.

Availability of data and materials

No datasets were generated or analysed during the current study.

Abbreviations

AC:

Anal carcinoma

ADCs:

Antibody-drug conjugates

ASCC:

Anal squamous cell cancer

GI:

Gastrointestinal

PD-1:

Programmed death protein 1

ICIs:

Immune checkpoint inhibitors

CRC:

Colorectal cancer

CAR-T:

Chimeric antigen receptor T-cell

ESCC:

Oesophageal squamous cell carcinoma

GERD:

Gastroesophageal reflux disease

HNSCC:

Head and neck squamous cell carcinoma

OS:

Overall survival

DFS:

Disease-free survival

pCR:

Pathologic complete response

PFS:

Progression-free survival

FGFR:

Fibroblast growth factor receptor

EGFR:

Epidermal growth factor receptor inhibitors

CPS:

Combined positive score

EFS:

Event-free survival

RFS:

Recurrence-free survival

DOR:

Duration of response

CR:

Complete response

MPR:

Major pathologic response

ORR:

Objective response rate

MDSCs:

Myeloid-derived suppressor cells

ATRA:

All-trans retinoic acid

TAMs:

Tumour-associated macrophages

VISTA:

V-domain immunoglobulin suppressor of T-cell activation

TGF-β:

Transforming growth factor-β

CCR8:

C-C motif chemokine receptor 8

MSCs:

Mesenchymal stem cells

IL:

Interleukin

EMT:

Epithelial-mesenchymal transition

PDAC:

Pancreatic ductal adenocarcinoma

NNT:

Nicotinamide nucleotide transhydrogenase

NK:

Natural killer

TCR:

T-cell receptor

PMs:

Peritoneal macrophages

PC:

Peritoneal carcinomatosis

APC:

Antigen-presenting cells

TIGIT:

T-cell immunoreceptors with Ig and ITIM domains

CKs:

Chemokines

Ovs:

Oncolytic viruses

AI:

Artificial Intelligence

BRAF:

B-Raf proto-oncogene

CDH1:

Cadherin 1

CDKN2A:

Cyclin-dependent kinase inhibitor 2A

CRC:

Colorectal Cancer

CREBBP:

CREB-binding protein

CT:

Chemotherapy

CTC:

Circulating Tumour Cell

CTLA-4:

Cytotoxic T-Lymphocyte-Associated protein 4

CRT:

Chemoradiotherapy

CXCL:

Chemokine (C-X-C motif) ligand

DC:

Dendritic Cell

DCR:

Disease Control Rate

DNA:

Deoxyribonucleic Acid

dMMR:

Deficient Mismatch Repair

EC:

Oesophageal Cancer

EAC:

Oesophageal Adenocarcinoma

EBV:

Epstein-Barr Virus

EBVaGC:

Epstein-Barr Virus-associated Gastric Cancer

EV:

Extracellular Vesicle

FOXP3:

Forkhead box P3

GC:

Gastric Cancer

GEJC:

Gastric and gastroesophageal junction cancer

GIPS:

Genomic Immunotherapy Prognostic Score

H.pylori:

Helicobacter pylori

ICIs:

Immune Checkpoint Inhibitors

IFN:

Interferon

IL:

Interleukin

irOS:

Immune-related Overall Survival

JAK2:

Janus Kinase 2

KRAS:

Kirsten Rat Sarcoma viral oncogene homolog

LIF:

Leukaemia Inhibitory Factor

LA-ESCC:

Resectable locally advanced ESCC

lncRNA:

Long non-coding RNA

m6A:

N6-methyladenosine

mCRC:

Metastatic Colorectal Cancer

MET:

MET proto-oncogene

mOS:

Median Overall Survival

mPFS:

Median Progression-Free Survival

MSS:

Microsatellite Stable

MSI:

Microsatellite Instability

MSI-H:

Microsatellite Instability-High

nCRT:

Neoadjuvant chemoradiotherapy

NICHE:

Neoadjuvant Immunotherapy for Colon Cancer Patients with Mismatch Repair Deficiency

NGS:

Next-Generation Sequencing

NII:

Neuroinflammatory Infiltration

ORR:

Objective Response Rate

PD-L1:

Programmed Death-Ligand 1

PD-L2:

Programmed Death-Ligand 2

PDGF-B:

Platelet-Derived Growth Factor-B

PI3KCA:

Phosphatidylinositol-4,5-bisphosphate 3-kinase catalytic subunit alpha

PNI:

Perineural Invasion

POLE:

DNA Polymerase Epsilon

POLD-1:

DNA Polymerase Delta 1

PTCH1:

Patched 1

PTEN:

Phosphatase and Tensin Homolog

RAS:

Rat Sarcoma viral oncogene homolog

RCTs:

Randomized controlled trials

RHOA:

Ras Homolog Family Member A

RNF43:

Ring Finger Protein 43

SPEN:

Spen Family Transcriptional Repressor

TP:

Taxane plus platinum

TCGA:

The Cancer Genome Atlas

TIL:

Tumour-Infiltrating Lymphocyte

TIME:

Tumour Immune Microenvironment

TIIC:

Tumour-Infiltrating Immune Cell

TLS:

Tertiary Lymphoid Structure

TME:

Tumour Microenvironment

TMB:

Tumour Mutational Burden

TMB-H:

Tumour Mutational Burden-High

TNF:

Tumour Necrosis Factor

VEGF:

Vascular Endothelial Growth Factor

References

  1. Bray F, et al. Global cancer statistics 2022: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J Clin. 2024;74:229–63. https://doi.org/10.3322/caac.21834.

    Article  PubMed  Google Scholar 

  2. Huang J, et al. Updated epidemiology of gastrointestinal cancers in East Asia. Nat Rev Gastroenterol Hepatol. 2023;20:271–87. https://doi.org/10.1038/s41575-022-00726-3.

    Article  PubMed  Google Scholar 

  3. Dunn GP, Old LJ, Schreiber RD. The three Es of cancer immunoediting. Annu Rev Immunol. 2004;22:329–60. https://doi.org/10.1146/annurev.immunol.22.012703.104803.

    Article  CAS  PubMed  Google Scholar 

  4. Chen DS, Mellman I. Oncology meets immunology: the cancer-immunity cycle. Immunity. 2013;39:1–10. https://doi.org/10.1016/j.immuni.2013.07.012.

    Article  CAS  PubMed  Google Scholar 

  5. Janjigian YY, et al. First-line nivolumab plus chemotherapy versus chemotherapy alone for advanced gastric, gastro-oesophageal junction, and oesophageal adenocarcinoma (CheckMate 649): a randomised, open-label, phase 3 trial. Lancet. 2021;398:27–40. https://doi.org/10.1016/s0140-6736(21)00797-2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Xu J, et al. Sintilimab plus chemotherapy for unresectable gastric or gastroesophageal junction cancer: the ORIENT-16 randomized clinical trial. JAMA. 2023;330:2064–74. https://doi.org/10.1001/jama.2023.19918.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Diaz LA, et al. Pembrolizumab versus chemotherapy for microsatellite instability-high or mismatch repair-deficient metastatic colorectal cancer (KEYNOTE-177): final analysis of a randomised, open-label, phase 3 study. Lancet Oncol. 2022;23:659–70. https://doi.org/10.1016/s1470-2045(22)00197-8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Sun JM, et al. Pembrolizumab plus chemotherapy versus chemotherapy alone for first-line treatment of advanced oesophageal cancer (KEYNOTE-590): a randomised, placebo-controlled, phase 3 study. Lancet. 2021;398:759–71. https://doi.org/10.1016/s0140-6736(21)01234-4.

    Article  CAS  PubMed  Google Scholar 

  9. Janjigian YY, et al. The KEYNOTE-811 trial of dual PD-1 and HER2 blockade in HER2-positive gastric cancer. Nature. 2021;600:727-+. https://doi.org/10.1038/s41586-021-04161-3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Shah MA, et al. First-line pembrolizumab (pembro) plus chemotherapy (chemo) for advanced esophageal cancer: 5-year outcomes from the phase 3 KEYNOTE-590 study. J Clin Oncol. 2024;42:250–250. https://doi.org/10.1200/JCO.2024.42.3_suppl.250.

    Article  Google Scholar 

  11. Andre T, et al. Neoadjuvant Nivolumab Plus Ipilimumab and Adjuvant Nivolumab in Localized Deficient Mismatch Repair/Microsatellite Instability-High Gastric or Esophagogastric Junction Adenocarcinoma: The GERCOR NEONIPIGA Phase II Study. J Clin Oncol. 2023;41:255-+. https://doi.org/10.1200/jco.22.00686.

    Article  CAS  PubMed  Google Scholar 

  12. Pietrantonio F, et al. INFINITY: A multicentre, single-arm, multi-cohort, phase II trial of tremelimumab and durvalumab as neoadjuvant treatment of patients with microsatellite instability-high (MSI) resectable gastric or gastroesophageal junction adenocarcinoma (GAC/GEJAC). J Clin Oncol. 2023;41:358–358.

    Article  Google Scholar 

  13. Raufi AG, et al. Abstract CT009: Phase II trial of perioperative pembrolizumab plus capecitabine and oxaliplatin followed by adjuvant pembrolizumab for resectable gastric and gastroesophageal junction (GC/GEJ) adenocarcinoma. Cancer Res. 2022;82:CT009. https://doi.org/10.1158/1538-7445.Am2022-ct009.

    Article  Google Scholar 

  14. Yuan S, et al. Perioperative PD-1 antibody toripalimab plus SOX or XELOX chemotherapy versus SOX or XELOX alone for locally advanced gastric or gastro-oesophageal junction cancer: Results from a prospective, randomized, open-label, phase II trial. J Clin Oncol. 2023;41:4001.

    Article  Google Scholar 

  15. Al-Batran S-E, et al. Surgical and pathological outcome, and pathological regression, in patients receiving perioperative atezolizumab in combination with FLOT chemotherapy versus FLOT alone for resectable esophagogastric adenocarcinoma: Interim results from DANTE, a randomized, multicenter, phase IIb trial of the FLOT-AIO German Gastric Cancer Group and Swiss SAKK. J Clin Oncol. 2022;40:4003.

    Article  Google Scholar 

  16. Shitara K, et al. Pembrolizumab plus chemotherapy vs chemotherapy as neoadjuvant and adjuvant therapy in locally-advanced gastric and gastroesophageal junction cancer: The phase III KEYNOTE-585 study. Ann Oncol. 2023;34:S1316–S1316. https://doi.org/10.1016/j.annonc.2023.10.075.

    Article  Google Scholar 

  17. Janjigian YY, et al. MATTERHORN: phase III study of durvalumab plus FLOT chemotherapy in resectable gastric/gastroesophageal junction cancer. Future Oncol. 2022;18:2465–73. https://doi.org/10.2217/fon-2022-0093.

    Article  CAS  PubMed  Google Scholar 

  18. Tintelnot J, et al. Pembrolizumab and trastuzumab in combination with FLOT in the perioperative treatment of HER2-positive, localized esophagogastric adenocarcinoma-a phase II trial of the AIO study group (AIO STO 0321). Front Oncol. 2023. https://doi.org/10.3389/fonc.2023.1272175.

    Article  PubMed  PubMed Central  Google Scholar 

  19. Chalabi M, et al. Neoadjuvant immunotherapy leads to pathological responses in MMR-proficient and MMR-deficient early-stage colon cancers. Nat Med. 2020;26:566-+. https://doi.org/10.1038/s41591-020-0805-8.

    Article  CAS  PubMed  Google Scholar 

  20. Chalabi M, et al. Neoadjuvant immune checkpoint inhibition in locally advanced MMR-deficient colon cancer: The NICHE-2 study. Ann Oncol. 2022;33:S1389–S1389. https://doi.org/10.1016/j.annonc.2022.08.016.

    Article  Google Scholar 

  21. Hu H, et al. Neoadjuvant PD-1 blockade with toripalimab, with or without celecoxib, in mismatch repair-deficient or microsatellite instability-high, locally advanced, colorectal cancer (P1CC): a single-centre, parallel-group, noncomparative, randomised, phase 2 trial. Lancet Gastroenterol Hepatol. 2022;7:38–48. https://doi.org/10.1016/s2468-1253(21)00348-4.

    Article  PubMed  Google Scholar 

  22. Ludford K, et al. Neoadjuvant pembrolizumab in localized microsatellite instability high/deficient mismatch repair solid tumors. J Clin Oncol. 2023;41:2181-+. https://doi.org/10.1200/jco.22.01351.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Chen G, et al. Neoadjuvant PD-1 blockade with sintilimab in mismatch-repair deficient, locally advanced rectal cancer: an open-label, single-centre phase 2 study. Lancet Gastroenterol Hepatol. 2023;8:422–31. https://doi.org/10.1016/s2468-1253(22)00439-3.

    Article  PubMed  Google Scholar 

  24. Qi CS, et al. Claudin18.2-specific CAR T cells in gastrointestinal cancers: phase 1 trial interim results. Nat Med. 2022;28:1189-+. https://doi.org/10.1038/s41591-022-01800-8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. National Health Commission Of The People's Republic Of, C. Chinese guidelines for diagnosis and treatment of esophageal carcinoma 2018 (English version). Chin J Cancer Res 31, 223–258, https://doi.org/10.21147/j.issn.1000-9604.2019.02.01 (2019).

  26. Luo H, et al. Effect of camrelizumab vs placebo added to chemotherapy on survival and progression-free survival in patients with advanced or metastatic esophageal squamous cell carcinoma: the ESCORT-1st randomized clinical trial. JAMA. 2021;326:916–25. https://doi.org/10.1001/jama.2021.12836.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Matz M, et al. Worldwide trends in esophageal cancer survival, by sub-site, morphology, and sex: an analysis of 696,974 adults diagnosed in 60 countries during 2000–2014 (CONCORD-3). Cancer Commun (Lond). 2023;43:963–80. https://doi.org/10.1002/cac2.12457.

    Article  PubMed  PubMed Central  Google Scholar 

  28. Morgan E, et al. The global landscape of esophageal squamous cell carcinoma and esophageal adenocarcinoma incidence and mortality in 2020 and projections to 2040: new estimates from GLOBOCAN 2020. Gastroenterology. 2022;163:649-658.e642. https://doi.org/10.1053/j.gastro.2022.05.054.

    Article  PubMed  Google Scholar 

  29. Integrated genomic characterization of oesophageal carcinoma. Nature. 2017;541:169–75. https://doi.org/10.1038/nature20805.

    Article  CAS  Google Scholar 

  30. Shah MA, et al. Improving outcomes in patients with oesophageal cancer. Nat Rev Clin Oncol. 2023;20:390–407. https://doi.org/10.1038/s41571-023-00757-y.

    Article  CAS  PubMed  Google Scholar 

  31. Kato K, et al. Nivolumab versus chemotherapy in patients with advanced oesophageal squamous cell carcinoma refractory or intolerant to previous chemotherapy (ATTRACTION-3): a multicentre, randomised, open-label, phase 3 trial. Lancet Oncol. 2019;20:1506–17. https://doi.org/10.1016/s1470-2045(19)30626-6.

    Article  CAS  PubMed  Google Scholar 

  32. Kojima T, et al. Randomized phase III KEYNOTE-181 study of pembrolizumab versus chemotherapy in advanced esophageal cancer. J Clin Oncol. 2020;38:4138–48. https://doi.org/10.1200/jco.20.01888.

    Article  CAS  PubMed  Google Scholar 

  33. Huang J, et al. Camrelizumab versus investigator’s choice of chemotherapy as second-line therapy for advanced or metastatic oesophageal squamous cell carcinoma (ESCORT): a multicentre, randomised, open-label, phase 3 study. Lancet Oncol. 2020;21:832–42. https://doi.org/10.1016/s1470-2045(20)30110-8.

    Article  CAS  PubMed  Google Scholar 

  34. Xu J, et al. Clinical and biomarker analyses of sintilimab versus chemotherapy as second-line therapy for advanced or metastatic esophageal squamous cell carcinoma: a randomized, open-label phase 2 study (ORIENT-2). Nat Commun. 2022;13:857. https://doi.org/10.1038/s41467-022-28408-3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Shen L, et al. Tislelizumab versus chemotherapy as second-line treatment for advanced or metastatic esophageal squamous cell carcinoma (RATIONALE-302): a randomized phase III study. J Clin Oncol. 2022;40:3065–76. https://doi.org/10.1200/jco.21.01926.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Doki Y, et al. Nivolumab combination therapy in advanced esophageal squamous-cell carcinoma. N Engl J Med. 2022;386:449–62. https://doi.org/10.1056/NEJMoa2111380.

    Article  CAS  PubMed  Google Scholar 

  37. Lu Z, et al. Sintilimab versus placebo in combination with chemotherapy as first line treatment for locally advanced or metastatic oesophageal squamous cell carcinoma (ORIENT-15): multicentre, randomised, double blind, phase 3 trial. BMJ. 2022;377: e068714. https://doi.org/10.1136/bmj-2021-068714.

    Article  PubMed  PubMed Central  Google Scholar 

  38. Wang ZX, et al. Toripalimab plus chemotherapy in treatment-naïve, advanced esophageal squamous cell carcinoma (JUPITER-06): a multi-center phase 3 trial. Cancer Cell. 2022;40:277-288.e273. https://doi.org/10.1016/j.ccell.2022.02.007.

    Article  CAS  PubMed  Google Scholar 

  39. Xu J, et al. Tislelizumab plus chemotherapy versus placebo plus chemotherapy as first-line treatment for advanced or metastatic oesophageal squamous cell carcinoma (RATIONALE-306): a global, randomised, placebo-controlled, phase 3 study. Lancet Oncol. 2023;24:483–95. https://doi.org/10.1016/S1470-2045(23)00108-0.

    Article  CAS  PubMed  Google Scholar 

  40. Song Y, et al. First-line serplulimab or placebo plus chemotherapy in PD-L1-positive esophageal squamous cell carcinoma: a randomized, double-blind phase 3 trial. Nat Med. 2023;29:473–82. https://doi.org/10.1038/s41591-022-02179-2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Li J, et al. First-line sugemalimab with chemotherapy for advanced esophageal squamous cell carcinoma: a randomized phase 3 study. Nat Med. 2024;30:740–8. https://doi.org/10.1038/s41591-024-02797-y.

    Article  CAS  PubMed  Google Scholar 

  42. Hsu C-H, et al. SKYSCRAPER-08: A phase III, randomized, double-blind, placebo-controlled study of first-line (1L) tiragolumab (tira) + atezolizumab (atezo) and chemotherapy (CT) in patients (pts) with esophageal squamous cell carcinoma (ESCC). J Clin Oncol. 2024;42:245–245. https://doi.org/10.1200/JCO.2024.42.3_suppl.245.

    Article  Google Scholar 

  43. Kelly RJ, et al. Adjuvant nivolumab in resected esophageal or gastroesophageal junction cancer. N Engl J Med. 2021;384:1191–203. https://doi.org/10.1056/NEJMoa2032125.

    Article  CAS  PubMed  Google Scholar 

  44. Li Y, et al. Chemotherapy plus camrelizumab versus chemotherapy alone as neoadjuvant treatment for resectable esophageal squamous cell carcinoma (ESCORT-NEO): A multi-center, randomized phase III trial. J Clin Oncol. 2024;42:LBA244. https://doi.org/10.1200/JCO.2024.42.3_suppl.LBA244.

    Article  Google Scholar 

  45. Rogers JE, et al. Nivolumab combination therapy as first-line treatments for unresectable, advanced or metastatic esophageal squamous cell carcinoma. Expert Rev Anticancer Ther. 2023;23:565–71. https://doi.org/10.1080/14737140.2023.2207826.

    Article  CAS  PubMed  Google Scholar 

  46. Li J, et al. O-4 GEMSTONE-304: a phase 3 study of sugemalimab plus chemotherapy versus chemotherapy as first-line treatment of patients with unresectable locally advanced, recurrent or metastatic esophageal squamous cell carcinoma (ESCC). Ann Oncol. 2023;34:S181–2. https://doi.org/10.1016/j.annonc.2023.04.019.

    Article  Google Scholar 

  47. Lu Z, et al. Effectiveness and safety of camrelizumab in advanced esophageal cancer: a prospective multicenter observational cohort studies (ESCORT-RWS). J Clin Oncol. 2023;41:4049–4049. https://doi.org/10.1200/JCO.2023.41.16_suppl.4049.

    Article  Google Scholar 

  48. Liu J, et al. Multicenter, single-arm, phase II trial of camrelizumab and chemotherapy as neoadjuvant treatment for locally advanced esophageal squamous cell carcinoma. J Immunother Cancer. 2022;10: e004291. https://doi.org/10.1136/jitc-2021-004291.

    Article  PubMed  PubMed Central  Google Scholar 

  49. Yan X, et al. Tislelizumab combined with chemotherapy as neoadjuvant therapy for surgically resectable esophageal cancer: a prospective, single-arm, phase II study (TD-NICE). Int J Surg. 2022;103: 106680. https://doi.org/10.1016/j.ijsu.2022.106680.

    Article  PubMed  Google Scholar 

  50. Shang X, et al. LBA3 Safety and efficacy of pembrolizumab combined with paclitaxel and cisplatin as a neoadjuvant treatment for locally advanced resectable (stage III) esophageal squamous cell carcinoma (Keystone-001): Interim analysis of a prospective, single-arm, single-center, phase II trial. Ann Oncol. 2021;32:S1428–9. https://doi.org/10.1016/j.annonc.2021.10.218.

    Article  Google Scholar 

  51. Xing W, et al. The sequence of chemotherapy and toripalimab might influence the efficacy of neoadjuvant chemoimmunotherapy in locally advanced esophageal squamous cell cancer-a phase II study. Front Immunol. 2021;12: 772450. https://doi.org/10.3389/fimmu.2021.772450.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Ma J, et al. Camrelizumab combined with paclitaxel and nedaplatin as neoadjuvant therapy for locally advanced esophageal squamous cell carcinoma (ESPRIT): a phase II, single-arm, exploratory research. J Clin Oncol. 2021;39:e16033–e16033. https://doi.org/10.1200/JCO.2021.39.15_suppl.e16033.

    Article  Google Scholar 

  53. Shen D, et al. The safety and efficacy of neoadjuvant PD-1 inhibitor with chemotherapy for locally advanced esophageal squamous cell carcinoma. J Gastrointest Oncol. 2021;12:1–10. https://doi.org/10.21037/jgo-20-599.

    Article  PubMed  PubMed Central  Google Scholar 

  54. Liu J, et al. Neoadjuvant camrelizumab plus chemotherapy for resectable, locally advanced esophageal squamous cell carcinoma (NIC-ESCC2019): A multicenter, phase 2 study. Int J Cancer. 2022;151:128–37. https://doi.org/10.1002/ijc.33976.

    Article  CAS  PubMed  Google Scholar 

  55. Yang P, et al. Neoadjuvant camrelizumab plus chemotherapy in treating locally advanced esophageal squamous cell carcinoma patients: a pilot study. World J Surg Oncol. 2021;19:333. https://doi.org/10.1186/s12957-021-02446-5.

    Article  PubMed  PubMed Central  Google Scholar 

  56. Gu Y, et al. 175P A study of neoadjuvant sintilimab combined with triplet chemotherapy of lipo-paclitaxel, cisplatin, and S-1 for resectable esophageal squamous cell carcinoma (ESCC). Ann Oncol. 2020;31:S1307–8. https://doi.org/10.1016/j.annonc.2020.10.196.

    Article  Google Scholar 

  57. Wang Z. Neoadjuvant camrelizumab combined with chemotherapy and apatinib for locally advanced thoracic esophageal squamous cell carcinoma (ESCC): A single-arm, open-label, phase Ib study. J Clin Oncol. 2021;39:4047–4047. https://doi.org/10.1200/JCO.2021.39.15_suppl.4047.

    Article  Google Scholar 

  58. Zhang Z, et al. Neoadjuvant sintilimab plus chemotherapy for locally advanced esophageal squamous cell carcinoma: a single-arm, single-center, phase 2 trial (ESONICT-1). Ann Transl Med. 2021;9:1623. https://doi.org/10.21037/atm-21-5381.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Li C, et al. Preoperative pembrolizumab combined with chemoradiotherapy for oesophageal squamous cell carcinoma (PALACE-1). Eur J Cancer. 2021;144:232–41. https://doi.org/10.1016/j.ejca.2020.11.039.

    Article  CAS  PubMed  Google Scholar 

  60. Yang Y, et al. Two-year outcomes of clinical N2–3 esophageal squamous cell carcinoma after neoadjuvant chemotherapy and immunotherapy from the phase 2 NICE study. J Thorac Cardiovasc Surg. 2024;167:838-847.e831. https://doi.org/10.1016/j.jtcvs.2023.08.056.

    Article  PubMed  Google Scholar 

  61. van Hagen P, et al. Preoperative chemoradiotherapy for esophageal or junctional cancer. N Engl J Med. 2012;366:2074–84. https://doi.org/10.1056/NEJMoa1112088.

    Article  PubMed  Google Scholar 

  62. Yang H, et al. Neoadjuvant chemoradiotherapy followed by surgery versus surgery alone for locally advanced squamous cell carcinoma of the esophagus (NEOCRTEC5010): a phase III multicenter, randomized, Open-Label Clinical Trial. J Clin Oncol. 2018;36:2796–803. https://doi.org/10.1200/jco.2018.79.1483.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Chen R, et al. A phase II clinical trial of toripalimab combined with neoadjuvant chemoradiotherapy in locally advanced esophageal squamous cell carcinoma (NEOCRTEC1901). EClinicalMedicine. 2023;62: 102118. https://doi.org/10.1016/j.eclinm.2023.102118.

    Article  PubMed  PubMed Central  Google Scholar 

  64. Tang H, et al. Neoadjuvant chemoradiotherapy versus neoadjuvant chemotherapy followed by minimally invasive esophagectomy for locally advanced esophageal squamous cell carcinoma: a prospective multicenter randomized clinical trial. Ann Oncol. 2023;34:163–72. https://doi.org/10.1016/j.annonc.2022.10.508.

    Article  CAS  PubMed  Google Scholar 

  65. van den Ende T, et al. Neoadjuvant chemoradiotherapy combined with atezolizumab for resectable esophageal adenocarcinoma: a single-arm phase II feasibility trial (PERFECT). Clin Cancer Res. 2021;27:3351–9. https://doi.org/10.1158/1078-0432.Ccr-20-4443.

    Article  PubMed  Google Scholar 

  66. Zhu Y, et al. Toripalimab combined with definitive chemoradiotherapy in locally advanced oesophageal squamous cell carcinoma (EC-CRT-001): a single-arm, phase 2 trial. Lancet Oncol. 2023;24:371–82. https://doi.org/10.1016/s1470-2045(23)00060-8.

    Article  CAS  PubMed  Google Scholar 

  67. Jiang N, et al. SCALE-1: safety and efficacy of short course neoadjuvant chemo-radiotherapy plus toripalimab for locally advanced resectable squamous cell carcinoma of esophagus. J Clin Oncol. 2022;40:4063–4063. https://doi.org/10.1200/JCO.2022.40.16_suppl.4063.

    Article  Google Scholar 

  68. Li Y, et al. Clinical efficacy of combination therapy of an immune checkpoint inhibitor with taxane plus platinum versus an immune checkpoint inhibitor with fluorouracil plus platinum in the first-line treatment of patients with locally advanced, metastatic, or recurrent esophageal squamous cell carcinoma. Front Oncol. 2022;12:1015302. https://doi.org/10.3389/fonc.2022.1015302.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Galluzzi L, Humeau J, Buqué A, Zitvogel L, Kroemer G. Immunostimulation with chemotherapy in the era of immune checkpoint inhibitors. Nat Rev Clin Oncol. 2020;17:725–41. https://doi.org/10.1038/s41571-020-0413-z.

    Article  PubMed  Google Scholar 

  70. Zhang Q, Cao G, Fan Z. Two cycles versus four cycles of neoadjuvant camrelizumab plus chemotherapy in patients with locally advanced esophageal squamous cell carcinoma (ESCC): A prospective, multicenter and randomized study. J Clin Oncol. 2023;41:350–350. https://doi.org/10.1200/JCO.2023.41.4_suppl.350.

    Article  Google Scholar 

  71. Yin J, et al. Neoadjuvant adebrelimab in locally advanced resectable esophageal squamous cell carcinoma: a phase 1b trial. Nat Med. 2023;29:2068–78. https://doi.org/10.1038/s41591-023-02469-3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Kim TW, et al. Anti-TIGIT antibody tiragolumab alone or with atezolizumab in patients with advanced solid tumors: a phase 1a/1b nonrandomized controlled trial. JAMA Oncol. 2023;9:1574–82. https://doi.org/10.1001/jamaoncol.2023.3867.

    Article  PubMed  PubMed Central  Google Scholar 

  73. Wang F, et al. 1020MO AdvanTIG-203: Phase II randomized, multicenter study of ociperlimab (OCI) + tislelizumab (TIS) in patients (pts) with unresectable, locally advanced, recurrent/metastatic esophageal squamous cell carcinoma (ESCC) and programmed cell death-ligand 1 (PD-L1) positivity. Ann Oncol. 2023;34:S621. https://doi.org/10.1016/j.annonc.2023.09.2159.

    Article  Google Scholar 

  74. Sun J-M, et al. MORPHEUS-EC: A phase Ib/II open-label, randomized study of first-line tiragolumab (tira) + atezolizumab (atezo) + chemotherapy (CT) in patients (pts) with esophageal cancer (EC). J Clin Oncol. 2024;42:324–324. https://doi.org/10.1200/JCO.2024.42.3_suppl.324.

    Article  Google Scholar 

  75. Liu Q, et al. Systemic therapy with or without local intervention for oligometastatic oesophageal squamous cell carcinoma (ESO-Shanghai 13): an open-label, randomised, phase 2 trial. Lancet Gastroenterol Hepatol. 2024;9:45–55. https://doi.org/10.1016/s2468-1253(23)00316-3.

    Article  PubMed  Google Scholar 

  76. Chen B, et al. 1576P Comparison of efficacy and safety of first-line systemic therapy combined with or without radiotherapy for stage IVb esophageal squamous cell carcinoma: a propensity score matching analysis. Ann Oncol. 2023;34:S879. https://doi.org/10.1016/j.annonc.2023.09.1488.

    Article  Google Scholar 

  77. Wagner AD, et al. Chemotherapy for advanced gastric cancer. Cochrane Database Syst Rev. 2017;8:CD004064. https://doi.org/10.1002/14651858.CD004064.pub4.

    Article  PubMed  Google Scholar 

  78. Fuchs CS, et al. Safety and efficacy of pembrolizumab monotherapy in patients with previously treated advanced gastric and gastroesophageal junction cancer: phase 2 clinical KEYNOTE-059 trial. JAMA Oncol. 2018;4:e180013–e180013. https://doi.org/10.1001/jamaoncol.2018.0013.

    Article  PubMed  PubMed Central  Google Scholar 

  79. Kang YK, et al. Nivolumab in patients with advanced gastric or gastro-oesophageal junction cancer refractory to, or intolerant of, at least two previous chemotherapy regimens (ONO-4538-12, ATTRACTION-2): a randomised, double-blind, placebo-controlled, phase 3 trial. Lancet. 2017;390:2461–71. https://doi.org/10.1016/s0140-6736(17)31827-5.

    Article  CAS  PubMed  Google Scholar 

  80. Shitara K, et al. Pembrolizumab versus paclitaxel for previously treated, advanced gastric or gastro-oesophageal junction cancer (KEYNOTE-061): a randomised, open-label, controlled, phase 3 trial. Lancet. 2018;392:123–33. https://doi.org/10.1016/s0140-6736(18)31257-1.

    Article  CAS  PubMed  Google Scholar 

  81. Kang YK, et al. Nivolumab plus chemotherapy versus placebo plus chemotherapy in patients with HER2-negative, untreated, unresectable advanced or recurrent gastric or gastro-oesophageal junction cancer (ATTRACTION-4): a randomised, multicentre, double-blind, placebo-controlled, phase 3 trial. Lancet Oncol. 2022;23:234–47. https://doi.org/10.1016/s1470-2045(21)00692-6.

    Article  CAS  PubMed  Google Scholar 

  82. Rha SY, et al. Pembrolizumab plus chemotherapy versus placebo plus chemotherapy for HER2-negative advanced gastric cancer (KEYNOTE-859): a multicentre, randomised, double-blind, phase 3 trial. Lancet Oncol. 2023;24:1181–95. https://doi.org/10.1016/S1470-2045(23)00515-6.

    Article  CAS  PubMed  Google Scholar 

  83. Shitara K, et al. Nivolumab plus chemotherapy or ipilimumab in gastro-oesophageal cancer. Nature. 2022;603:942–8. https://doi.org/10.1038/s41586-022-04508-4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Shitara K, et al. Nivolumab (NIVO) + chemotherapy (chemo) vs chemo as first-line (1L) treatment for advanced gastric cancer/gastroesophageal junction cancer/esophageal adenocarcinoma (GC/GEJC/EAC): 4 year (yr) follow-up of CheckMate 649. J Clin Oncol. 2024;42:306–306. https://doi.org/10.1200/JCO.2024.42.3_suppl.306.

    Article  Google Scholar 

  85. Zhang X, et al. LBA79 GEMSTONE-303: Prespecified progression-free survival (PFS) and overall survival (OS) final analyses of a phase III study of sugemalimab plus chemotherapy vs placebo plus chemotherapy in treatment-naïve advanced gastric or gastroesophageal junction (G/GEJ) adenocarcinoma. Ann Oncol. 2023;34:S1319. https://doi.org/10.1016/j.annonc.2023.10.080.

    Article  Google Scholar 

  86. Xu RH, et al. 139MO Tislelizumab (TIS) plus chemotherapy (Chemo) vs placebo (PBO) plus chemo as first-line (1L) treatment of advanced gastric or gastroesophageal junction adenocarcinoma (GC/GEJC): Final analysis results of the RATIONALE-305 study. Ann Oncol. 2023;34:S1526–7. https://doi.org/10.1016/j.annonc.2023.10.274.

    Article  Google Scholar 

  87. Moehler MH, et al. Rationale 305: Phase 3 study of tislelizumab plus chemotherapy vs placebo plus chemotherapy as first-line treatment (1L) of advanced gastric or gastroesophageal junction adenocarcinoma (GC/GEJC). J Clin Oncol. 2023;41:286–286. https://doi.org/10.1200/JCO.2023.41.4_suppl.286.

    Article  Google Scholar 

  88. Janjigian YY, et al. Pembrolizumab plus trastuzumab and chemotherapy for HER2-positive gastric or gastro-oesophageal junction adenocarcinoma: interim analyses from the phase 3 KEYNOTE-811 randomised placebo-controlled trial. Lancet. 2023;402:2197–208. https://doi.org/10.1016/s0140-6736(23)02033-0.

    Article  CAS  PubMed  Google Scholar 

  89. Shitara K, et al. Neoadjuvant and adjuvant pembrolizumab plus chemotherapy in locally advanced gastric or gastro-oesophageal cancer (KEYNOTE-585): an interim analysis of the multicentre, double-blind, randomised phase 3 study. Lancet Oncol. 2024;25:212–24. https://doi.org/10.1016/s1470-2045(23)00541-7.

    Article  CAS  PubMed  Google Scholar 

  90. Shitara K, et al. LBA74 Pembrolizumab plus chemotherapy vs chemotherapy as neoadjuvant and adjuvant therapy in locally-advanced gastric and gastroesophageal junction cancer: The phase III KEYNOTE-585 study. Ann Oncol. 2023;34:S1316. https://doi.org/10.1016/j.annonc.2023.10.075.

    Article  Google Scholar 

  91. Janjigian YY, et al. LBA73 Pathological complete response (pCR) to durvalumab plus 5-fluorouracil, leucovorin, oxaliplatin and docetaxel (FLOT) in resectable gastric and gastroesophageal junction cancer (GC/GEJC): Interim results of the global, phase III MATTERHORN study. Ann Oncol. 2023;34:S1315–6. https://doi.org/10.1016/j.annonc.2023.10.074.

    Article  Google Scholar 

  92. Terashima M, et al. ATTRACTION-5: A phase 3 study of nivolumab plus chemotherapy as postoperative adjuvant treatment for pathological stage III (pStage III) gastric or gastroesophageal junction (G/GEJ) cancer. J Clin Oncol. 2023;41:4000–4000. https://doi.org/10.1200/JCO.2023.41.16_suppl.4000.

    Article  Google Scholar 

  93. André T, et al. Neoadjuvant nivolumab plus ipilimumab and adjuvant nivolumab in localized deficient mismatch repair/microsatellite instability-high gastric or esophagogastric junction adenocarcinoma: the GERCOR NEONIPIGA phase II study. J Clin Oncol. 2023;41:255–65. https://doi.org/10.1200/jco.22.00686.

    Article  PubMed  Google Scholar 

  94. Pietrantonio F, et al. INFINITY: A multicentre, single-arm, multi-cohort, phase II trial of tremelimumab and durvalumab as neoadjuvant treatment of patients with microsatellite instability-high (MSI) resectable gastric or gastroesophageal junction adenocarcinoma (GAC/GEJAC). J Clin Oncol. 2023;41:358–358. https://doi.org/10.1200/JCO.2023.41.4_suppl.358.

    Article  Google Scholar 

  95. Fuchs CS, et al. Safety and efficacy of pembrolizumab monotherapy in patients with previously treated advanced gastric and gastroesophageal junction cancer: phase 2 clinical KEYNOTE-059 trial. JAMA Oncol. 2018;4: e180013. https://doi.org/10.1001/jamaoncol.2018.0013.

    Article  PubMed  PubMed Central  Google Scholar 

  96. Shitara K, et al. Efficacy and safety of pembrolizumab or pembrolizumab plus chemotherapy vs chemotherapy alone for patients with first-line, advanced gastric cancer: the KEYNOTE-062 Phase 3 randomized clinical trial. JAMA Oncol. 2020;6:1571–80. https://doi.org/10.1001/jamaoncol.2020.3370.

    Article  PubMed  Google Scholar 

  97. Cruz-Correa M, et al. Tislelizumab (TIS) plus chemotherapy (Chemo) vs placebo (PBO) plus chemo as first-line (1L) treatment of advanced gastric or gastroesophageal junction adenocarcinoma (GC/GEJC): Health-related quality of life (HRQoL) outcomes in the RATIONALE-305 study. J Clin Oncol. 2024;42:290–290. https://doi.org/10.1200/JCO.2024.42.3_suppl.290.

    Article  Google Scholar 

  98. Shen L, et al. First-line (1L) nivolumab (NIVO) plus chemotherapy (chemo) vs chemo in patients (pts) with advanced gastric cancer, gastroesophageal junction cancer, and esophageal adenocarcinoma (GC/GEJC/EAC): CheckMate 649 Chinese subgroup analysis with 3-year follow-up. J Clin Oncol. 2023;41:353–353. https://doi.org/10.1200/JCO.2023.41.4_suppl.353.

    Article  Google Scholar 

  99. Janjigian YY, et al. First-Line nivolumab plus chemotherapy for advanced gastric, gastroesophageal junction, and esophageal adenocarcinoma: 3-year follow-up of the phase iii checkmate 649 trial. J Clin Oncol. 2024;42:2012–20. https://doi.org/10.1200/JCO.23.01601.

    Article  CAS  PubMed  Google Scholar 

  100. Liu Z, Zhang D, Chen S. Unveiling the gastric microbiota: implications for gastric carcinogenesis, immune responses, and clinical prospects. J Exp Clin Cancer Res. 2024;43:118. https://doi.org/10.1186/s13046-024-03034-7.

    Article  PubMed  PubMed Central  Google Scholar 

  101. Zavros Y, Merchant JL. The immune microenvironment in gastric adenocarcinoma. Nat Rev Gastroenterol Hepatol. 2022;19:451–67. https://doi.org/10.1038/s41575-022-00591-0.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Zhu Y, Zhu F, Ba H, Chen J, Bian X. Helicobacter pylori infection and PD-L1 expression in gastric cancer: a meta-analysis. Eur J Clin Invest. 2023;53: e13880. https://doi.org/10.1111/eci.13880.

    Article  CAS  PubMed  Google Scholar 

  103. Liu T, et al. First-line nivolumab plus chemotherapy vs chemotherapy in patients with advanced gastric, gastroesophageal junction and esophageal adenocarcinoma: CheckMate 649 Chinese subgroup analysis. Int J Cancer. 2023;152:749–60. https://doi.org/10.1002/ijc.34296.

    Article  CAS  PubMed  Google Scholar 

  104. Jia K, et al. Helicobacter pylori and immunotherapy for gastrointestinal cancer. Innovation. 2024;5:100561. https://doi.org/10.1016/j.xinn.2023.100561.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Bai Y, et al. Efficacy and predictive biomarkers of immunotherapy in Epstein-Barr virus-associated gastric cancer. J Immunother Cancer. 2022. https://doi.org/10.1136/jitc-2021-004080.

    Article  PubMed  PubMed Central  Google Scholar 

  106. Bang YJ, et al. Trastuzumab in combination with chemotherapy versus chemotherapy alone for treatment of HER2-positive advanced gastric or gastro-oesophageal junction cancer (ToGA): a phase 3, open-label, randomised controlled trial. Lancet. 2010;376:687–97. https://doi.org/10.1016/S0140-6736(10)61121-X.

    Article  CAS  PubMed  Google Scholar 

  107. Oki E, et al. Protein expression of programmed death 1 ligand 1 and HER2 in gastric carcinoma. Oncology. 2017;93:387–94. https://doi.org/10.1159/000479231.

    Article  CAS  PubMed  Google Scholar 

  108. Wang Y, et al. Disitamab vedotin (RC48) plus toripalimab for HER2-expressing advanced gastric or gastroesophageal junction and other solid tumours: a multicentre, open label, dose escalation and expansion phase 1 trial. EClinicalMedicine. 2024;68: 102415. https://doi.org/10.1016/j.eclinm.2023.102415.

    Article  PubMed  PubMed Central  Google Scholar 

  109. Shitara K, et al. Zolbetuximab plus mFOLFOX6 in patients with CLDN18.2-positive, HER2-negative, untreated, locally advanced unresectable or metastatic gastric or gastro-oesophageal junction adenocarcinoma (SPOTLIGHT): a multicentre, randomised, double-blind, phase 3 trial. Lancet. 2023;401:1655–68. https://doi.org/10.1016/S0140-6736(23)00620-7.

    Article  CAS  PubMed  Google Scholar 

  110. Shah MA, et al. Zolbetuximab plus CAPOX in CLDN18.2-positive gastric or gastroesophageal junction adenocarcinoma: the randomized, phase 3 GLOW trial. Nat Med. 2023;29:2133–41. https://doi.org/10.1038/s41591-023-02465-7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Nishibata T, et al. Effect of anti-claudin 18.2 monoclonal antibody zolbetuximab alone or combined with chemotherapy or programmed cell death-1 blockade in syngeneic and xenograft gastric cancer models. J Pharmacol Sci. 2024;155:84–93. https://doi.org/10.1016/j.jphs.2024.04.004.

    Article  CAS  PubMed  Google Scholar 

  112. Wang Y, et al. 132P A phase I clinical trial of QLS31905 in advanced solid tumors. Immuno-Oncol Technol. 2023;20: 100604. https://doi.org/10.1016/j.iotech.2023.100604.

    Article  Google Scholar 

  113. Wang FH, et al. The Chinese Society of Clinical Oncology (CSCO): Clinical guidelines for the diagnosis and treatment of gastric cancer, 2023. Cancer Commun (Lond). 2024;44:127–72. https://doi.org/10.1002/cac2.12516.

    Article  CAS  PubMed  Google Scholar 

  114. Chao J, et al. Assessment of pembrolizumab therapy for the treatment of microsatellite instability-high gastric or gastroesophageal junction cancer among patients in the KEYNOTE-059, KEYNOTE-061, and KEYNOTE-062 clinical trials. JAMA Oncol. 2021;7:895–902. https://doi.org/10.1001/jamaoncol.2021.0275.

    Article  PubMed  PubMed Central  Google Scholar 

  115. Muro K, et al. 1513MO A phase II study of nivolumab plus low dose ipilimumab as first -line therapy in patients with advanced gastric or esophago-gastric junction MSI-H tumor: First results of the NO LIMIT study (WJOG13320G/CA209-7W7). Ann Oncol. 2023;34:S852–3. https://doi.org/10.1016/j.annonc.2023.09.1426.

    Article  Google Scholar 

  116. Li J, et al. Subcutaneous envafolimab monotherapy in patients with advanced defective mismatch repair/microsatellite instability high solid tumors. J Hematol Oncol. 2021;14:95. https://doi.org/10.1186/s13045-021-01095-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Li J, et al. Tislelizumab in previously treated, locally advanced unresectable/metastatic microsatellite instability-high/mismatch repair-deficient solid tumors. Chin J Cancer Res. 2024;36:257–69. https://doi.org/10.21147/j.issn.1000-9604.2024.03.03.

    Article  PubMed  PubMed Central  Google Scholar 

  118. Qin S, et al. Serplulimab, a novel anti-PD-1 antibody, in patients with microsatellite instability-high solid tumours: an open-label, single-arm, multicentre, phase II trial. Br J Cancer. 2022;127:2241–8. https://doi.org/10.1038/s41416-022-02001-3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Wu X, et al. 601P Pembrolizumab in patients of Chinese descent with microsatellite instability-high/mismatch repair deficient advanced solid tumors: KEYNOTE-158. Ann Oncol. 2023;34:S1707. https://doi.org/10.1016/j.annonc.2023.10.316.

    Article  Google Scholar 

  120. Qiu MZ, et al. Dynamic single-cell mapping unveils Epstein-Barr virus-imprinted T-cell exhaustion and on-treatment response. Signal Transduct Target Ther. 2023;8:370. https://doi.org/10.1038/s41392-023-01622-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Kim ST, et al. Comprehensive molecular characterization of clinical responses to PD-1 inhibition in metastatic gastric cancer. Nat Med. 2018;24:1449–58. https://doi.org/10.1038/s41591-018-0101-z.

    Article  CAS  PubMed  Google Scholar 

  122. Sun YT, et al. PD-1 antibody camrelizumab for Epstein-Barr virus-positive metastatic gastric cancer: a single-arm, open-label, phase 2 trial. Am J Cancer Res. 2021;11:5006–15.

    CAS  PubMed  PubMed Central  Google Scholar 

  123. Li XD, et al. Characteristic analysis of alpha-fetoprotein-producing gastric carcinoma in China. World J Surg Oncol. 2013;11:246. https://doi.org/10.1186/1477-7819-11-246.

    Article  PubMed  PubMed Central  Google Scholar 

  124. Wang Y, et al. Camrelizumab plus apatinib and SOX as first-line treatment in patients with alpha-fetoprotein–producing gastric or gastroesophageal junction adenocarcinoma: A single-arm, multi-center, phase 2 trial. J Clin Oncol. 2024;42:351–351. https://doi.org/10.1200/JCO.2024.42.3_suppl.351.

    Article  Google Scholar 

  125. Ji, J. et al. Cadonilimab plus chemotherapy versus chemotherapy as first-line treatment for unresectable locally advanced or metastatic gastric or gastroesophageal junction (G_GEJ) adenocarcinoma (COMPASSION-15)_ A randomized, double-blind, phase 3 trial. AACR annual meeting Session CTPL02 (2024).

  126. Moehler M, et al. Phase III Trial of Avelumab Maintenance After First-Line Induction Chemotherapy Versus Continuation of Chemotherapy in Patients With Gastric Cancers: Results From JAVELIN Gastric 100. J Clin Oncol. 2021;39:966–77. https://doi.org/10.1200/JCO.20.00892.

    Article  CAS  PubMed  Google Scholar 

  127. Yuan SQ, et al. Perioperative toripalimab and chemotherapy in locally advanced gastric or gastro-esophageal junction cancer: a randomized phase 2 trial. Nat Med. 2024;30:552–9. https://doi.org/10.1038/s41591-023-02721-w.

    Article  CAS  PubMed  Google Scholar 

  128. Ding X, et al. PERSIST: A multicenter, randomized phase II trial of perioperative oxaliplatin and S-1 (SOX) with or without sintilimab in resectable locally advanced gastric/gastroesophageal junction cancer (GC/GEJC). J Clin Oncol. 2023;41:364–364. https://doi.org/10.1200/JCO.2023.41.4_suppl.364.

    Article  Google Scholar 

  129. Lorenzen S, et al. Perioperative atezolizumab plus fluorouracil, leucovorin, oxaliplatin, and docetaxel for resectable esophagogastric cancer: interim results from the randomized, multicenter, phase II/III DANTE/IKF-s633 trial. J Clin Oncol. 2024;42:410–20. https://doi.org/10.1200/JCO.23.00975.

    Article  CAS  PubMed  Google Scholar 

  130. Raimondi A, et al. TremelImumab and Durvalumab Combination for the Non-OperatIve Management (NOM) of Microsatellite InstabiliTY (MSI)-High Resectable Gastric or Gastroesophageal Junction Cancer: The Multicentre, Single-Arm, Multi-Cohort, Phase II INFINITY Study. Cancers. 2021. https://doi.org/10.3390/cancers13112839.

    Article  PubMed  PubMed Central  Google Scholar 

  131. Guidoboni M, et al. Microsatellite instability and high content of activated cytotoxic lymphocytes identify colon cancer patients with a favorable prognosis. Am J Pathol. 2001;159:297–304. https://doi.org/10.1016/S0002-9440(10)61695-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Mulet-Margalef N, et al. Challenges and therapeutic opportunities in the dMMR/MSI-H colorectal cancer landscape. Cancers. 2023. https://doi.org/10.3390/cancers15041022.

    Article  PubMed  PubMed Central  Google Scholar 

  133. Bonneville R, et al. Landscape of microsatellite instability across 39 cancer types. JCO Precis Oncol. 2017. https://doi.org/10.1200/PO.17.00073.

    Article  PubMed  PubMed Central  Google Scholar 

  134. Ribic CM, et al. Tumor microsatellite-instability status as a predictor of benefit from fluorouracil-based adjuvant chemotherapy for colon cancer. N Engl J Med. 2003;349:247–57. https://doi.org/10.1056/NEJMoa022289.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Sargent DJ, et al. Defective mismatch repair as a predictive marker for lack of efficacy of fluorouracil-based adjuvant therapy in colon cancer. J Clin Oncol. 2010;28:3219–26. https://doi.org/10.1200/jco.2009.27.1825.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. André T, et al. Pembrolizumab in microsatellite-instability–high advanced colorectal cancer. N Engl J Med. 2020;383:2207–18. https://doi.org/10.1056/NEJMoa2017699.

    Article  PubMed  Google Scholar 

  137. Cohen R, et al. Microsatellite Instability in patients with stage III colon cancer receiving fluoropyrimidine with or without oxaliplatin: an ACCENT pooled analysis of 12 adjuvant trials. J Clin Oncol. 2021;39:642–51. https://doi.org/10.1200/jco.20.01600.

    Article  CAS  PubMed  Google Scholar 

  138. Asaoka Y, Ijichi H, Koike K. PD-1 Blockade in tumors with mismatch-repair deficiency. N Engl J Med. 2015;373:1979–1979. https://doi.org/10.1056/NEJMc1510353.

    Article  PubMed  Google Scholar 

  139. Overman MJ, et al. Nivolumab in patients with metastatic DNA mismatch repair-deficient or microsatellite instability-high colorectal cancer (CheckMate 142): an open-label, multicentre, phase 2 study. Lancet Oncol. 2017;18:1182–91. https://doi.org/10.1016/s1470-2045(17)30422-9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Le DT, et al. Pembrolizumab for previously treated, microsatellite instability-high/mismatch repair-deficient advanced colorectal cancer: final analysis of KEYNOTE-164. Eur J Cancer. 2023;186:185–95. https://doi.org/10.1016/j.ejca.2023.02.016.

    Article  CAS  PubMed  Google Scholar 

  141. Shiu KK, et al. LBA32 Pembrolizumab versus chemotherapy in microsatellite instability-high (MSI-H)/mismatch repair-deficient (dMMR) metastatic colorectal cancer (mCRC): 5-year follow-up of the randomized phase III KEYNOTE-177 study. Ann Oncol. 2023;34:S1271–2. https://doi.org/10.1016/j.annonc.2023.10.024.

    Article  Google Scholar 

  142. Andre T, et al. Nivolumab (NIVO) plus ipilimumab (IPI) vs chemotherapy (chemo) as first-line (1L) treatment for microsatellite instability-high/mismatch repair-deficient (MSI-H/dMMR) metastatic colorectal cancer (mCRC): First results of the CheckMate 8HW study. J Clin Oncol. 2024;42:LBA768. https://doi.org/10.1200/JCO.2024.42.3_suppl.LBA768.

    Article  Google Scholar 

  143. Overman MJ, et al. Nivolumab (NIVO) ± ipilimumab (IPI) in patients (pts) with microsatellite instability-high/mismatch repair-deficient (MSI-H/dMMR) metastatic colorectal cancer (mCRC): Five-year follow-up from CheckMate 142. J Clin Oncol. 2022;40:3510–3510. https://doi.org/10.1200/JCO.2022.40.16_suppl.3510.

    Article  Google Scholar 

  144. Lenz H-J, et al. First-line (1L) nivolumab (NIVO) + ipilimumab (IPI) in patients (pts) with microsatellite instability-high/mismatch repair deficient (MSI-H/dMMR) metastatic colorectal cancer (mCRC): 64-month (mo) follow-up from CheckMate 142. J Clin Oncol. 2023;41:3550–3550. https://doi.org/10.1200/JCO.2023.41.16_suppl.3550.

    Article  Google Scholar 

  145. Chen M, et al. PD-1/PD-L1 inhibitor plus chemotherapy versus PD-1/PD-L1 inhibitor in microsatellite instability gastrointestinal cancers: a multicenter retrospective study. JCO Precis Oncol. 2023;7: e2200463. https://doi.org/10.1200/po.22.00463.

    Article  PubMed  PubMed Central  Google Scholar 

  146. Overman M, et al. Colorectal cancer metastatic dMMR immuno-therapy (COMMIT) study: A randomized phase III study of atezolizumab (atezo) monotherapy versus mFOLFOX6/bevacizumab/atezo in the first-line treatment of patients (pts) with deficient DNA mismatch repair (dMMR) or microsatellite instability-high (MSI-H) metastatic colorectal cancer (mCRC)—NRG-GI004/SWOG-S1610. J Clin Oncol. 2024;42:TPS231. https://doi.org/10.1200/JCO.2024.42.3_suppl.TPS231.

    Article  Google Scholar 

  147. Wu Z, et al. PD-1 blockade plus COX inhibitors in dMMR metastatic colorectal cancer: Clinical, genomic, and immunologic analyses from the PCOX trial. Med. 2024. https://doi.org/10.1016/j.medj.2024.05.002.

    Article  PubMed  Google Scholar 

  148. Chen M, et al. The optimal therapy after progression on immune checkpoint inhibitors in MSI metastatic gastrointestinal cancer patients: a multicenter retrospective cohort study. Cancers. 2022. https://doi.org/10.3390/cancers14205158.

    Article  PubMed  PubMed Central  Google Scholar 

  149. Chalabi M, et al. Neoadjuvant immunotherapy leads to pathological responses in MMR-proficient and MMR-deficient early-stage colon cancers. Nat Med. 2020;26:566–76. https://doi.org/10.1038/s41591-020-0805-8.

    Article  Google Scholar 

  150. Chalabi M, et al. LBA7 Neoadjuvant immune checkpoint inhibition in locally advanced MMR-deficient colon cancer: The NICHE-2 study. Ann Oncol. 2022;33:S1389. https://doi.org/10.1016/j.annonc.2022.08.016.

    Article  Google Scholar 

  151. Cercek A, et al. PD-1 blockade in mismatch repair-deficient, locally advanced rectal cancer. N Engl J Med. 2022;386:2363–76. https://doi.org/10.1056/NEJMoa2201445.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Hu H, et al. Neoadjuvant PD-1 blockade with toripalimab, with or without celecoxib, in mismatch repair-deficient or microsatellite instability-high, locally advanced, colorectal cancer (PICC): a single-centre, parallel-group, non-comparative, randomised, phase 2 trial. Lancet Gastroenterol Hepatol. 2022;7:38–48. https://doi.org/10.1016/S2468-1253(21)00348-4.

    Article  PubMed  Google Scholar 

  153. Sinicrope FA, et al. Randomized trial of standard chemotherapy alone or combined with atezolizumab as adjuvant therapy for patients with stage III colon cancer and deficient mismatch repair (ATOMIC, Alliance A021502). J Clin Oncol. 2019;37:e15169–e15169. https://doi.org/10.1200/JCO.2019.37.15_suppl.e15169.

    Article  Google Scholar 

  154. Lau D, et al. POLEM: Avelumab plus fluoropyrimidine-based chemotherapy as adjuvant treatment for stage III dMMR or POLE exonuclease domain mutant colon cancer—a phase III randomized study. J Clin Oncol. 2019;37:TPS3615. https://doi.org/10.1200/JCO.2019.37.15_suppl.TPS3615.

    Article  Google Scholar 

  155. Bando H, Ohtsu A, Yoshino T. Therapeutic landscape and future direction of metastatic colorectal cancer. Nat Rev Gastroenterol Hepatol. 2023;20:306–22. https://doi.org/10.1038/s41575-022-00736-1.

    Article  CAS  PubMed  Google Scholar 

  156. Zhao W, et al. Colorectal cancer immunotherapy-Recent progress and future directions. Cancer Lett. 2022;545: 215816. https://doi.org/10.1016/j.canlet.2022.215816.

    Article  CAS  PubMed  Google Scholar 

  157. Kawazoe A, et al. Lenvatinib plus pembrolizumab versus standard of care for previously treated metastatic colorectal cancer: final analysis of the randomized, open-label, phase III LEAP-017 study. J Clin Oncol. 2024. https://doi.org/10.1200/jco.23.02736.

    Article  PubMed  Google Scholar 

  158. Fakih M, et al. Regorafenib, ipilimumab, and nivolumab for patients with microsatellite stable colorectal cancer and disease progression with prior chemotherapy: a phase 1 nonrandomized clinical trial. JAMA Oncol. 2023;9:627–34. https://doi.org/10.1001/jamaoncol.2022.7845.

    Article  PubMed  PubMed Central  Google Scholar 

  159. Segal NH, et al. Phase II single-arm study of durvalumab and tremelimumab with concurrent radiotherapy in patients with mismatch repair-proficient metastatic colorectal cancer. Clin Cancer Res. 2021;27:2200–8. https://doi.org/10.1158/1078-0432.Ccr-20-2474.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Thibaudin M, et al. First-line durvalumab and tremelimumab with chemotherapy in RAS-mutated metastatic colorectal cancer: a phase 1b/2 trial. Nat Med. 2023;29:2087–98. https://doi.org/10.1038/s41591-023-02497-z.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Wang F, et al. Combined anti-PD-1, HDAC inhibitor and anti-VEGF for MSS/pMMR colorectal cancer: a randomized phase 2 trial. Nat Med. 2024;30:1035–43. https://doi.org/10.1038/s41591-024-02813-1.

    Article  CAS  PubMed  Google Scholar 

  162. Tian J, et al. Combined PD-1, BRAF and MEK inhibition in BRAFV600E colorectal cancer: a phase 2 trial. Nat Med. 2023;29:458–66. https://doi.org/10.1038/s41591-022-02181-8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Morano F, et al. Temozolomide followed by combination with low-dose ipilimumab and nivolumab in patients with microsatellite-stable, O(6)-methylguanine-DNA methyltransferase-silenced metastatic colorectal cancer: the MAYA trial. J Clin Oncol. 2022;40:1562–73. https://doi.org/10.1200/jco.21.02583.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Boccaccino A, et al. An immune-related gene expression profile to predict the efficacy of adding atezolizumab to first-line FOLFOXIRI plus bevacizumab in metastatic colorectal cancer: A translational analysis of the phase II randomized AtezoTRIBE study. J Clin Oncol. 2022;40:3581–3581. https://doi.org/10.1200/JCO.2022.40.16_suppl.3581.

    Article  Google Scholar 

  165. Lenz HJ, et al. Modified FOLFOX6 plus bevacizumab with and without nivolumab for first-line treatment of metastatic colorectal cancer: phase 2 results from the CheckMate 9X8 randomized clinical trial. J Immunother Cancer. 2024. https://doi.org/10.1136/jitc-2023-008409.

    Article  PubMed  PubMed Central  Google Scholar 

  166. Gondal TA, et al. Anal cancer: the past, present and future. Curr Oncol. 2023;30:3232–50. https://doi.org/10.3390/curroncol30030246.

    Article  PubMed  PubMed Central  Google Scholar 

  167. Kim R, et al. Carboplatin and paclitaxel treatment is effective in advanced anal cancer. Oncology. 2014;87:125–32. https://doi.org/10.1159/000361051.

    Article  CAS  PubMed  Google Scholar 

  168. Kim S, et al. Docetaxel, cisplatin, and fluorouracil chemotherapy for metastatic or unresectable locally recurrent anal squamous cell carcinoma (Epitopes-HPV02): a multicentre, single-arm, phase 2 study. Lancet Oncol. 2018;19:1094–106. https://doi.org/10.1016/s1470-2045(18)30321-8.

    Article  CAS  PubMed  Google Scholar 

  169. Rao S, et al. International rare cancers initiative multicenter randomized phase II trial of cisplatin and fluorouracil versus carboplatin and paclitaxel in advanced anal cancer: interaact. J Clin Oncol. 2020;38:2510–8. https://doi.org/10.1200/jco.19.03266.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Morris VK, et al. Nivolumab for previously treated unresectable metastatic anal cancer (NCI9673): a multicentre, single-arm, phase 2 study. Lancet Oncol. 2017;18:446–53. https://doi.org/10.1016/s1470-2045(17)30104-3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Lonardi S, et al. Randomized phase II trial of avelumab alone or in combination with cetuximab for patients with previously treated, locally advanced, or metastatic squamous cell anal carcinoma: the CARACAS study. J Immunother Cancer. 2021. https://doi.org/10.1136/jitc-2021-002996.

    Article  PubMed  PubMed Central  Google Scholar 

  172. Marabelle A, et al. Pembrolizumab for previously treated advanced anal squamous cell carcinoma: results from the non-randomised, multicohort, multicentre, phase 2 KEYNOTE-158 study. Lancet Gastroenterol Hepatol. 2022;7:446–54. https://doi.org/10.1016/s2468-1253(21)00382-4.

    Article  PubMed  Google Scholar 

  173. Rao S, et al. A phase II study of retifanlimab (INCMGA00012) in patients with squamous carcinoma of the anal canal who have progressed following platinum-based chemotherapy (POD1UM-202). ESMO Open. 2022;7: 100529. https://doi.org/10.1016/j.esmoop.2022.100529.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. Morris V, et al. O-12 NCI9673 (Part B): A multi-institutional ETCTN randomized phase II study of nivolumab with or without ipilimumab in refractory, metastatic squamous cell carcinoma of the anal canal. Ann Oncol. 2023;34:S185–6. https://doi.org/10.1016/j.annonc.2023.04.027.

    Article  Google Scholar 

  175. Morris V, et al. 403MO Atezolizumab in combination with bevacizumab for patients with unresectable/metastatic anal cancer. Ann Oncol. 2020;31:S412. https://doi.org/10.1016/j.annonc.2020.08.514.

    Article  Google Scholar 

  176. Kim S, et al. Atezolizumab plus modified docetaxel, cisplatin, and fluorouracil as first-line treatment for advanced anal cancer (SCARCE C17–02 PRODIGE 60): a randomised, non-comparative, phase 2 study. Lancet Oncol. 2024;25:518–28. https://doi.org/10.1016/s1470-2045(24)00081-0.

    Article  CAS  PubMed  Google Scholar 

  177. Roth MT, et al. A randomized phase III study of immune checkpoint inhibition with chemotherapy in treatment-naive metastatic anal cancer patients: A trial of the ECOG-ACRIN cancer research group (EA2176). J Clin Oncol. 2021;39:TPS3614. https://doi.org/10.1200/JCO.2021.39.15_suppl.TPS3614.

    Article  Google Scholar 

  178. Rao S, Jones M, Bowman J, Tian C, Spano JP. POD1UM-303/InterAACT 2: A phase III, global, randomized, double-blind study of retifanlimab or placebo plus carboplatin-paclitaxel in patients with locally advanced or metastatic squamous cell anal carcinoma. Front Oncol. 2022;12: 935383. https://doi.org/10.3389/fonc.2022.935383.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Yamashita K, Iwatsuki M, Ajani JA, Baba H. Programmed death ligand-1 expression in gastrointestinal cancer: clinical significance and future challenges. Ann Gastroenterol Surg. 2020;4:369–78. https://doi.org/10.1002/ags3.12348.

    Article  PubMed  PubMed Central  Google Scholar 

  180. Janjigian YY, et al. Nivolumab (NIVO) plus chemotherapy (chemo) vs chemo as first-line (1L) treatment for advanced gastric cancer/gastroesophageal junction cancer/esophageal adenocarcinoma (GC/GEJC/EAC): 3-year follow-up from CheckMate 649. J Clin Oncol. 2023;41:291–291.

    Article  Google Scholar 

  181. Kojima T, et al. Randomized phase III KEYNOTE-181 study of pembrolizumab versus chemotherapy in advanced esophageal cancer. J Clin Oncol. 2020. https://doi.org/10.1200/jco.20.01888.

    Article  PubMed  Google Scholar 

  182. Janjigian YY, et al. First-Line Nivolumab Plus Chemotherapy for Advanced Gastric, Gastroesophageal Junction, and Esophageal Adenocarcinoma: 3-Year Follow-Up of the Phase III CheckMate 649 Trial. J Clin Oncol Off J Am Soc Clin Oncol. 2024. https://doi.org/10.1200/jco.23.01601.

    Article  Google Scholar 

  183. Xie T, et al. Appropriate PD-L1 cutoff value for gastric cancer immunotherapy: a systematic review and meta-analysis. Front Oncol. 2021. https://doi.org/10.3389/fonc.2021.646355.

    Article  PubMed  PubMed Central  Google Scholar 

  184. Zhao JJ, et al. Low programmed death-ligand 1-expressing subgroup outcomes of first-line immune checkpoint inhibitors in gastric or esophageal adenocarcinoma. J Clin Oncol. 2022;40:392-+. https://doi.org/10.1200/jco.21.01862.

    Article  CAS  PubMed  Google Scholar 

  185. Mansfield AS, et al. Heterogeneity of programmed cell death ligand 1 expression in multifocal lung cancer. Clin Cancer Res. 2016;22:2177–82. https://doi.org/10.1158/1078-0432.ccr-15-2246.

    Article  CAS  PubMed  Google Scholar 

  186. Lim SH, et al. Changes in tumour expression of programmed death-ligand 1 after neoadjuvant concurrent chemoradiotherapy in patients with squamous oesophageal cancer. Eur J Cancer. 2016;52:1–9. https://doi.org/10.1016/j.ejca.2015.09.019.

    Article  CAS  PubMed  Google Scholar 

  187. Zhang C, et al. Immediate and substantial evolution of T-cell repertoire in peripheral blood and tumor microenvironment of patients with esophageal squamous cell carcinoma treated with preoperative chemotherapy. Carcinogenesis. 2018;39:1389–98. https://doi.org/10.1093/carcin/bgy116.

    Article  CAS  PubMed  Google Scholar 

  188. Yeong J, et al. Choice of PD-L1 immunohistochemistry assay influences clinical eligibility for gastric cancer immunotherapy. Gastric Cancer. 2022;25:741–50. https://doi.org/10.1007/s10120-022-01301-0.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Lemery S, Keegan P, Pazdur R. First FDA approval agnostic of cancer site - when a biomarker defines the indication. N Engl J Med. 2017;377:1409–12. https://doi.org/10.1056/NEJMp1709968.

    Article  PubMed  Google Scholar 

  190. Llosa NJ, et al. The vigorous immune microenvironment of microsatellite instable colon cancer is balanced by multiple counter-inhibitory checkpoints. Cancer Discov. 2015;5:43–51. https://doi.org/10.1158/2159-8290.cd-14-0863.

    Article  CAS  PubMed  Google Scholar 

  191. Marabelle A, et al. Efficacy of pembrolizumab in patients with noncolorectal high microsatellite instability/mismatch repair-deficient cancer: results from the phase II KEYNOTE-158 Study. J Clin Oncol. 2020;38:1-+. https://doi.org/10.1200/jco.19.02105.

    Article  CAS  PubMed  Google Scholar 

  192. O’Neil BH, et al. Safety and antitumor activity of the anti-PD-1 antibody pembrolizumab in patients with advanced colorectal carcinoma. PLoS ONE. 2017. https://doi.org/10.1371/journal.pone.0189848.

    Article  PubMed  PubMed Central  Google Scholar 

  193. Le DT, et al. Programmed death-1 blockade in mismatch repair deficient colorectal cancer. J Clin Oncol. 2016. https://doi.org/10.1200/JCO.2016.34.15_suppl.103.

    Article  PubMed  PubMed Central  Google Scholar 

  194. Andre T, et al. Pembrolizumab in microsatellite-instability-high advanced colorectal cancer. N Engl J Med. 2020;383:2207–18. https://doi.org/10.1056/NEJMoa2017699.

    Article  CAS  PubMed  Google Scholar 

  195. Shiu KK, et al. Pembrolizumab versus chemotherapy in microsatellite instability-high (MSI-H)/mismatch repair-deficient (dMMR) metastatic colorectal cancer (mCRC): 5-year follow-up of the randomized phase III KEYNOTE-177 study. Ann Oncol. 2023;34:S1271–2. https://doi.org/10.1016/j.annonc.2023.10.024.

    Article  Google Scholar 

  196. Verschoor YL, et al. LBA31 Neoadjuvant nivolumab plus relatlimab (anti-LAG3) in locally advanced MMR-deficient colon cancers: the NICHE-3 study. Ann Oncol. 2023;34:S1270. https://doi.org/10.1016/j.annonc.2023.10.023.

    Article  Google Scholar 

  197. Kishore C, Bhadra P. Current advancements and future perspectives of immunotherapy in colorectal cancer research. Eur J Pharmacol. 2021. https://doi.org/10.1016/j.ejphar.2020.173819.

    Article  PubMed  Google Scholar 

  198. Salem ME, et al. Comparative molecular analyses of esophageal squamous cell carcinoma, esophageal adenocarcinoma, and gastric adenocarcinoma. Oncologist. 2018;23:1319–27. https://doi.org/10.1634/theoncologist.2018-0143.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  199. Bartley AN, et al. Mismatch repair and microsatellite instability testing for immune checkpoint inhibitor therapy guideline from the college of american pathologists in collaboration with the association for molecular pathology and fight colorectal cancer. Arch Pathol Lab Med. 2022;146:1194–210. https://doi.org/10.5858/arpa.2021-0632-CP.

    Article  CAS  PubMed  Google Scholar 

  200. Middha S, et al. Reliable pan-cancer microsatellite instability assessment by using targeted next-generation sequencing data. Jco Precis Oncol. 2017. https://doi.org/10.1200/po.17.00084.

    Article  PubMed  PubMed Central  Google Scholar 

  201. Saeed OAM, et al. Evaluating mismatch repair deficiency for solid tumor immunotherapy eligibility: immunohistochemistry versus microsatellite molecular testing. Hum Pathol. 2021;115:10–8. https://doi.org/10.1016/j.humpath.2021.05.009.

    Article  CAS  PubMed  Google Scholar 

  202. Chen M-L, et al. Comparison of microsatellite status detection methods in colorectal carcinoma. Int J Clin Exp Pathol. 2018;11:1431–8.

    PubMed  PubMed Central  Google Scholar 

  203. Cohen R, et al. Assessment of local clinical practice for testing of mismatch repair deficiency in metastatic colorectal cancer: the need for new diagnostic guidelines prior to immunotherapy. Ann Oncol. 2018;2:9.

    Google Scholar 

  204. Boyle TA, et al. Summary of microsatellite instability test results from laboratories participating in proficiency surveys proficiency survey results from 2005 to 2012. Arch Pathol Lab Med. 2014;138:363–70. https://doi.org/10.5858/arpa.2013-0159-CP.

    Article  PubMed  Google Scholar 

  205. Wang J, et al. Mutational analysis of microsatellite-stable gastrointestinal cancer with high tumour mutational burden: a retrospective cohort study. Lancet Oncol. 2023;24:151–61. https://doi.org/10.1016/s1470-2045(22)00783-5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  206. Zhou KI, et al. Spatial and temporal heterogeneity of PD-L1 expression and tumor mutational burden in gastroesophageal adenocarcinoma at baseline diagnosis and after chemotherapy. Clin Cancer Res. 2020;26:6453–63. https://doi.org/10.1158/1078-0432.ccr-20-2085.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  207. Wang F, et al. Safety, efficacy and tumor mutational burden as a biomarker of overall survival benefit in chemo-refractory gastric cancer treated with toripalimab, a PD-1 antibody in phase Ib/II clinical trial NCT02915432. Ann Oncol Off J Eur Soc Med Oncol. 2019;30:1479–86. https://doi.org/10.1093/annonc/mdz197.

    Article  CAS  Google Scholar 

  208. Shitara, K. et al. The association of tissue tumor mutational burden (tTMB) using the Foundation Medicine genomic platform with efficacy of pembrolizumab versus paclitaxel in patients (pts) with gastric cancer (GC) from KEYNOTE-061. J Clin Oncol 38 (2020).

  209. McGrail DJ, et al. High tumor mutation burden fails to predict immune checkpoint blockade response across all cancer types. Ann Oncol. 2021;32:661–72. https://doi.org/10.1016/j.annonc.2021.02.006.

    Article  CAS  PubMed  Google Scholar 

  210. Bass AJ, et al. Comprehensive molecular characterization of gastric adenocarcinoma. Nature. 2014;513:202–9. https://doi.org/10.1038/nature13480.

    Article  CAS  Google Scholar 

  211. Gullo I, et al. The transcriptomic landscape of gastric cancer: insights into epstein-barr virus infected and microsatellite unstable tumors. Int J Mol Sci. 2018. https://doi.org/10.3390/ijms19072079.

    Article  PubMed  PubMed Central  Google Scholar 

  212. Qiu M-Z, et al. Dynamic single-cell mapping unveils Epstein?Barr virus-imprinted T-cell exhaustion and on-treatment response. Signal Transd Targeted Ther. 2023. https://doi.org/10.1038/s41392-023-01622-1.

    Article  Google Scholar 

  213. Kim ST, et al. Comprehensive molecular characterization of clinical responses to PD-1 inhibition in metastatic gastric cancer. Nat Med. 2018;24:1449-+. https://doi.org/10.1038/s41591-018-0101-z.

    Article  CAS  PubMed  Google Scholar 

  214. Xie T, et al. Positive status of epstein-barr virus as a biomarker for gastric cancer immunotherapy: a prospective observational study. J Immunother. 2020;43:139–44. https://doi.org/10.1097/cji.0000000000000316.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  215. Usui Y, et al. Helicobacter pylori, Homologous-Recombination Genes, and Gastric Cancer. N Engl J Med. 2023;388:1181–90. https://doi.org/10.1056/NEJMoa2211807.

    Article  CAS  PubMed  Google Scholar 

  216. Oster P, et al. Helicobacter pylori infection has a detrimental impact on the efficacy of cancer immunotherapies. Gut. 2022;71:457–66. https://doi.org/10.1136/gutjnl-2020-323392.

    Article  CAS  PubMed  Google Scholar 

  217. Ralser A, et al. Helicobacter pylori promotes colorectal carcinogenesis by deregulating intestinal immunity and inducing a mucus-degrading microbiota signature. Gut. 2023;72:1258–70. https://doi.org/10.1136/gutjnl-2022-328075.

    Article  CAS  PubMed  Google Scholar 

  218. Gong X, et al. Helicobacter pylori infection reduces the efficacy of cancer immunotherapy: a systematic review and meta-analysis. Helicobacter. 2023. https://doi.org/10.1111/hel.13011.

    Article  PubMed  Google Scholar 

  219. Shi Y, Zheng H, Wang M, Ding S. Influence of Helicobacter pylori infection on PD-1/PD-L1 blockade therapy needs more attention. Helicobacter. 2022. https://doi.org/10.1111/hel.12878.

    Article  PubMed  Google Scholar 

  220. Magahis PT, et al. Impact of Helicobacter pylori infection status on outcomes among patients with advanced gastric cancer treated with immune checkpoint inhibitors. J Immunother Cancer. 2023. https://doi.org/10.1136/jitc-2023-007699.

    Article  PubMed  PubMed Central  Google Scholar 

  221. Jia W, Rajani C, Xu H, Zheng X. Gut microbiota alterations are distinct for primary colorectal cancer and hepatocellular carcinoma. Protein Cell. 2021;12:374–93. https://doi.org/10.1007/s13238-020-00748-0.

    Article  PubMed  Google Scholar 

  222. Stewart OA, Wu F, Chen Y. The role of gastric microbiota in gastric cancer. Gut Microbes. 2020;11:1220–30. https://doi.org/10.1080/19490976.2020.1762520.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  223. Nasrollahzadeh D, et al. Variations of gastric corpus microbiota are associated with early esophageal squamous cell carcinoma and squamous dysplasia. Sci Rep. 2015;5:8820. https://doi.org/10.1038/srep08820.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  224. Gopalakrishnan V, Helmink BA, Spencer CN, Reuben A, Wargo JA. The influence of the gut microbiome on cancer, immunity, and cancer immunotherapy. Cancer Cell. 2018;33:570–80. https://doi.org/10.1016/j.ccell.2018.03.015.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  225. Spencer SP, Fragiadakis GK, Sonnenburg JL. Pursuing human-relevant gut microbiota-immune interactions. Immunity. 2019;51:225–39. https://doi.org/10.1016/j.immuni.2019.08.002.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  226. Lu Y, et al. Gut microbiota influence immunotherapy responses: mechanisms and therapeutic strategies. J Hematol Oncol. 2022;15:47. https://doi.org/10.1186/s13045-022-01273-9.

    Article  PubMed  PubMed Central  Google Scholar 

  227. Zhang SL, et al. Pectin supplement significantly enhanced the anti-PD-1 efficacy in tumor-bearing mice humanized with gut microbiota from patients with colorectal cancer. Theranostics. 2021;11:4155–70. https://doi.org/10.7150/thno.54476.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  228. Jia D, et al. Microbial metabolite enhances immunotherapy efficacy by modulating T cell stemness in pan-cancer. Cell. 2024;187:1651-1665.e1621. https://doi.org/10.1016/j.cell.2024.02.022.

    Article  CAS  PubMed  Google Scholar 

  229. Gunjur A, et al. A gut microbial signature for combination immune checkpoint blockade across cancer types. Nat Med. 2024;30:797–809. https://doi.org/10.1038/s41591-024-02823-z.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  230. Peng Z, et al. The gut microbiome is associated with clinical response to anti-PD-1/PD-L1 immunotherapy in gastrointestinal cancer. Cancer Immunol Res. 2020;8:1251–61. https://doi.org/10.1158/2326-6066.Cir-19-1014.

    Article  PubMed  Google Scholar 

  231. Han Z, et al. The gut microbiome affects response of treatments in HER2-negative advanced gastric cancer. Clin Transl Med. 2023;13: e1312. https://doi.org/10.1002/ctm2.1312.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  232. Cheng S, et al. Multi-omics of the gut microbial ecosystem in patients with microsatellite-instability-high gastrointestinal cancer resistant to immunotherapy. Cell Rep Med. 2024;5: 101355. https://doi.org/10.1016/j.xcrm.2023.101355.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  233. Verschoor YL, et al. Neoadjuvant atezolizumab plus chemotherapy in gastric and gastroesophageal junction adenocarcinoma: the phase 2 PANDA trial. Nat Med. 2024. https://doi.org/10.1038/s41591-023-02758-x.

    Article  PubMed  PubMed Central  Google Scholar 

  234. Yu P, Fu YX. Tumor-infiltrating T lymphocytes: friends or foes? Lab Invest. 2006;86:231–45. https://doi.org/10.1038/labinvest.3700389.

    Article  CAS  PubMed  Google Scholar 

  235. Sakaguchi S, Yamaguchi T, Nomura T, Ono M. Regulatory T cells and immune tolerance. Cell. 2008;133:775–87. https://doi.org/10.1016/j.cell.2008.05.009.

    Article  CAS  PubMed  Google Scholar 

  236. Chen X, et al. First-line camrelizumab (a PD-1 inhibitor) plus apatinib (an VEGFR-2 inhibitor) and chemotherapy for advanced gastric cancer (SPACE): a phase 1 study. Signal Transduct Target Ther. 2024;9:73–73. https://doi.org/10.1038/s41392-024-01773-9.

    Article  PubMed  PubMed Central  Google Scholar 

  237. Masuda K, et al. Multiplexed single-cell analysis reveals prognostic and nonprognostic T cell types in human colorectal cancer. Jci Insight. 2022. https://doi.org/10.1172/jci.insight.154646.

    Article  PubMed  PubMed Central  Google Scholar 

  238. Thibaudin M, et al. First-line durvalumab and tremelimumab with chemotherapy in RAS-mutated metastatic colorectal cancer: a phase 1b/2 trial. Nat Med. 2023;29:2087-+. https://doi.org/10.1038/s41591-023-02497-z.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  239. Salgado R, et al. The evaluation of tumor-infiltrating lymphocytes (TILs) in breast cancer: recommendations by an International TILs Working Group 2014. Ann Oncol. 2015;26:259–71. https://doi.org/10.1093/annonc/mdu450.

    Article  CAS  PubMed  Google Scholar 

  240. Pilipow K, Darwich A, Losurdo A. T-cell-based breast cancer immunotherapy. Semin Cancer Biol. 2021;72:90–101. https://doi.org/10.1016/j.semcancer.2020.05.019.

    Article  CAS  PubMed  Google Scholar 

  241. Ayers M, et al. IFN-γ-related mRNA profile predicts clinical response to PD-1 blockade. J Clin Investig. 2017;127:2930–40. https://doi.org/10.1172/jci91190.

    Article  PubMed  PubMed Central  Google Scholar 

  242. Doi T, et al. Safety and antitumor activity of the anti-programmed death-1 antibody pembrolizumab in patients with advanced esophageal carcinoma. J Clin Oncol. 2018;36:61-+. https://doi.org/10.1200/jco.2017.74.9846.

    Article  CAS  PubMed  Google Scholar 

  243. Zou D, Song A, Yong W. Prognostic role of IL-8 in cancer patients treated with immune checkpoint inhibitors: a system review and meta-analysis. Front Oncol. 2023. https://doi.org/10.3389/fonc.2023.1176574.

    Article  PubMed  PubMed Central  Google Scholar 

  244. Wang F, et al. Combined anti-PD-1, HDAC inhibitor and anti-VEGF for MSS/pMMR colorectal cancer: a randomized phase 2 trial. Nat Med. 2024. https://doi.org/10.1038/s41591-024-02813-1.

    Article  PubMed  PubMed Central  Google Scholar 

  245. Bagaev A, et al. Conserved pan-cancer microenvironment subtypes predict response to immunotherapy. Cancer Cell. 2021;39:845-+. https://doi.org/10.1016/j.ccell.2021.04.014.

    Article  CAS  PubMed  Google Scholar 

  246. Zeng D, et al. Tumor microenvironment evaluation promotes precise checkpoint immunotherapy of advanced gastric cancer. J Immunother Cancer. 2021. https://doi.org/10.1136/jitc-2021-002467.

    Article  PubMed  PubMed Central  Google Scholar 

  247. Sheng W, Li X, Li J, Mi Y, Li F. Evaluating prognostic value and relevant gene signatures of tumor microenvironment characterization in esophageal carcinoma. J Gastrointest Oncol. 2021;12:1228-+. https://doi.org/10.2103/jgo-21-371.

    Article  PubMed  PubMed Central  Google Scholar 

  248. Wang J-B, et al. Tumor immunophenotyping-derived signature identifies prognosis and neoadjuvant immunotherapeutic responsiveness in gastric cancer. Adv Sci. 2023. https://doi.org/10.1002/advs.202207417.

    Article  Google Scholar 

  249. Haas M, et al. Stromal regulatory T-cells are associated with a favourable prognosis in gastric cancer of the cardia. Bmc Gastroenterol. 2009. https://doi.org/10.1186/1471-230x-9-65.

    Article  PubMed  PubMed Central  Google Scholar 

  250. Lee JS, Won HS, Sun DS, Hong JH, Ko YH. Prognostic role of tumor-infiltrating lymphocytes in gastric cancer: a systematic review and meta-analysis. Medicine. 2018. https://doi.org/10.1097/md.0000000000011769.

    Article  PubMed  PubMed Central  Google Scholar 

  251. Li F, et al. CD4/CD8+T cells, DC subsets, Foxp3, and IDO expression are predictive indictors of gastric cancer prognosis. Cancer Med. 2019;8:7330–44. https://doi.org/10.1002/cam4.2596.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  252. Sadeghi Rad H, et al. The evolving landscape of predictive biomarkers in immuno-oncology with a focus on spatial technologies. Clin Transl Immunol. 2020. https://doi.org/10.1002/cti2.1215.

    Article  Google Scholar 

  253. Sautes-Fridman C, Petitprez F, Calderaro J, Fridman WH. Tertiary lymphoid structures in the era of cancer immunotherapy. Nat Rev Cancer. 2019;19:307–25. https://doi.org/10.1038/s41568-019-0144-6.

    Article  CAS  PubMed  Google Scholar 

  254. Zhan Z, et al. High endothelial venules proportion in tertiary lymphoid structure is a prognostic marker and correlated with anti-tumor immune microenvironment in colorectal cancer. Ann Med. 2023;55:114–26. https://doi.org/10.1080/07853890.2022.2153911.

    Article  CAS  PubMed  Google Scholar 

  255. Saberzadeh-Ardestani B, et al. Immune marker spatial distribution and clinical outcome after PD-1 blockade in mismatch repair-deficient. Adv Colorectal Carcinomas Clin Cancer Res. 2023;29:4268–77. https://doi.org/10.1158/1078-0432.ccr-23-1109.

    Article  CAS  Google Scholar 

  256. Mpekris F, et al. Combining microenvironment normalization strategies to improve cancer immunotherapy. Proc Natl Acad Sci USA. 2020;117:3728–37. https://doi.org/10.1073/pnas.1919764117.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  257. Li X, et al. Predicting response to immunotherapy in gastric cancer via assessing perineural invasion-mediated inflammation in tumor microenvironment. J Exp Clin Cancer Res. 2023. https://doi.org/10.1186/s13046-023-02730-0.

    Article  PubMed  PubMed Central  Google Scholar 

  258. Fassler DJ, et al. Deep learning-based image analysis methods for brightfield-acquired multiplex immunohistochemistry images. Diagn Pathol. 2020. https://doi.org/10.1186/s13000-020-01003-0.

    Article  PubMed  PubMed Central  Google Scholar 

  259. Che G, et al. Circumventing drug resistance in gastric cancer: a spatial multi-omics exploration of chemo and immuno-therapeutic response dynamics. Drug Resist Updat. 2024;74: 101080. https://doi.org/10.1016/j.drup.2024.101080.

    Article  CAS  PubMed  Google Scholar 

  260. Nakamura Y, et al. Clinical utility of circulating tumor DNA sequencing in advanced gastrointestinal cancer: SCRUM-Japan GI-SCREEN and GOZILA studies. Nat Med. 2020. https://doi.org/10.1038/s41591-020-1063-5.

    Article  PubMed  Google Scholar 

  261. Wang H, et al. Minimal residual disease guided radical chemoradiotherapy combined with immunotherapy after neoadjuvant immunochemotherapy followed by adjuvant immunotherapy for esophageal squamous cell cancer (ECMRD-001): a study protocol for a prospective cohort study. Front Immunol. 2024. https://doi.org/10.3389/fimmu.2023.1330928.

    Article  PubMed  PubMed Central  Google Scholar 

  262. Maron SB, et al. Determinants of survival with combined HER2 and PD-1 blockade in metastatic esophagogastric cancer. Clin Cancer Res. 2023;29:3633–40. https://doi.org/10.1158/1078-0432.ccr-22-3769.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  263. Tie J, Cohen JD, Wang Y. Circulating tumor DNA analyses as markers of recurrence risk and benefit of adjuvant therapy for stage III colon cancer (vol 66, pg 584, 2019). JAMA Oncol. 2019;5:1811–1811. https://doi.org/10.1001/jamaoncol.2019.5667.

    Article  Google Scholar 

  264. Nakamura Y, et al. Circulating tumor DNA-guided treatment with pertuzumab plus trastuzumab for HER2-amplified metastatic colorectal cancer: a phase 2 trial. Nat Med. 2021;27:1899-+. https://doi.org/10.1038/s41591-021-01553-w.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  265. Sartore-Bianchi A, et al. Circulating tumor DNA to guide rechallenge with panitumumab in metastatic colorectal cancer: the phase 2 CHRONOS trial. Nat Med. 2022;28:1612-+. https://doi.org/10.1038/s41591-022-01886-0.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  266. Cremolini C, et al. Rechallenge for patients with RAS and BRAF wild-type metastatic colorectal cancer with acquired resistance to first-line cetuximab and irinotecan a phase 2 single-arm clinical trial. JAMA Oncol. 2019;5:343–50. https://doi.org/10.1001/jamaoncol.2018.5080.

    Article  PubMed  Google Scholar 

  267. Jung S-H, et al. Monitoring the outcomes of systemic chemotherapy including immune checkpoint inhibitor for HER2-positive metastatic gastric cancer by liquid biopsy. Yonsei Med J. 2023;64:531–40. https://doi.org/10.3349/ymj.2023.0096.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  268. Hyung J, et al. Safety and efficacy of trastuzumab biosimilar plus irinotecan or gemcitabine in patients with previously treated HER2 (ERBB2)-positive non-breast/non-gastric solid tumors: a phase II basket trial with circulating tumor DNA analysis. Esmo Open. 2023. https://doi.org/10.1016/j.esmoop.2023.101583.

    Article  PubMed  PubMed Central  Google Scholar 

  269. Zhuang W, et al. Discovery and validation of methylation signatures in circulating cell-free DNA for early detection of esophageal cancer: A case-control study. J Clin Oncol. 2021. https://doi.org/10.1200/JCO.2021.39.15_suppl.e16027.

    Article  Google Scholar 

  270. Smyth EC, et al. EGFR amplification and outcome in a randomised phase III trial of chemotherapy alone or chemotherapy plus panitumumab for advanced gastro-oesophageal cancers. Gut. 2021;70:1632–41. https://doi.org/10.1136/gutjnl-2020-322658.

    Article  CAS  PubMed  Google Scholar 

  271. Kim ST, et al. Impact of genomic alterations on lapatinib treatment outcome and cell-free genomic landscape during HER2 therapy in HER2+gastric cancer patients. Ann Oncol. 2018;29:1037–48. https://doi.org/10.1093/annonc/mdy034.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  272. Li Y, et al. Evolutionary expression of HER2 conferred by chromosome aneuploidy on circulating gastric cancer cells contributes to developing targeted and chemotherapeutic resistance. Clin Cancer Res. 2018;24:5261–71. https://doi.org/10.1158/1078-0432.ccr-18-1205.

    Article  CAS  PubMed  Google Scholar 

  273. Kato S, et al. Genomic assessment of blood-derived circulating tumor DNA in patients with colorectal cancers: correlation with tissue sequencing, therapeutic response, and survival. JCO Precis Oncol. 2019. https://doi.org/10.1200/po.18.00158.

    Article  PubMed  PubMed Central  Google Scholar 

  274. Jin Y, et al. The predicting role of circulating tumor DNA landscape in gastric cancer patients treated with immune checkpoint inhibitors. Mol Cancer. 2020;19:154. https://doi.org/10.1186/s12943-020-01274-7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  275. Bikhchandani M, et al. POLE-mutant colon cancer treated with PD-1 blockade showing clearance of circulating tumor dna and prolonged disease-free interval. Genes. 2023. https://doi.org/10.3390/genes14051054.

    Article  PubMed  PubMed Central  Google Scholar 

  276. Zhang C, et al. Clinical implications of plasma ctDNA features and dynamics in gastric cancer treated with HER2-targeted therapies. Clin Transl Med. 2020;10: e254. https://doi.org/10.1002/ctm2.254.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  277. Al Zoughbi W, et al. Validation of a circulating tumor DNA-based next-generation sequencing assay in a cohort of patients with solid tumors: a proposed solution for decentralized plasma testing. Oncologist. 2021;26:e1971–81. https://doi.org/10.1002/onco.13905.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  278. Cheng B, et al. Enumeration and characterization of circulating tumor cells and its application in advanced gastric cancer. Onco Targets Ther. 2019;12:7887–96. https://doi.org/10.2147/ott.S223222.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  279. Yue C, et al. Dynamic change of PD-L1 expression on circulating tumor cells in advanced solid tumor patients undergoing PD-1 blockade therapy. Oncoimmunology. 2018;7: e1438111. https://doi.org/10.1080/2162402x.2018.1438111.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  280. Kennedy LC, et al. Liquid Biopsy assessment of circulating tumor cell PD-L1 and IRF-1 expression in patients with advanced solid tumors receiving immune checkpoint inhibitor. Target Oncol. 2022;17:329–41. https://doi.org/10.1007/s11523-022-00891-0.

    Article  PubMed  PubMed Central  Google Scholar 

  281. Ouyang Y, et al. Prognostic significance of programmed cell death-ligand 1 expression on circulating tumor cells in various cancers: a systematic review and meta-analysis. Cancer Med. 2021;10:7021–39. https://doi.org/10.1002/cam4.4236.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  282. Chong X, et al. Tracking circulating PD-L1-positive cells to monitor the outcome of patients with gastric cancer receiving anti-HER2 plus anti-PD-1 therapy. Hum Cell. 2024;37:258–70. https://doi.org/10.1007/s13577-023-00990-8.

    Article  CAS  PubMed  Google Scholar 

  283. Xie F, Xu M, Lu J, Mao L, Wang S. The role of exosomal PD-L1 in tumor progression and immunotherapy. Mol Cancer. 2019;18:146. https://doi.org/10.1186/s12943-019-1074-3.

    Article  PubMed  PubMed Central  Google Scholar 

  284. Zhang C, et al. Plasma extracellular vesicle derived protein profile predicting and monitoring immunotherapeutic outcomes of gastric cancer. J Extracellular Vesicles. 2022. https://doi.org/10.1002/jev2.12209.

    Article  Google Scholar 

  285. Loriot Y, et al. Plasma proteomics identifies leukemia inhibitory factor (LIF) as a novel predictive biomarker of immune-checkpoint blockade resistance. Ann Oncol. 2021;32:1381–90. https://doi.org/10.1016/j.annonc.2021.08.1748.

    Article  CAS  PubMed  Google Scholar 

  286. Li Z-C, et al. Proteomic and metabolomic features in patients with HCC responding to lenvatinib and anti-PD1 therapy. Cell Rep. 2024. https://doi.org/10.1016/j.celrep.2024.113877.

    Article  PubMed  PubMed Central  Google Scholar 

  287. Chen Y, et al. Metabolomic machine learning predictor for diagnosis and prognosis of gastric cancer. Nat Commun. 2024;15:1657. https://doi.org/10.1038/s41467-024-46043-y.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  288. Peng W, et al. Upregulated METTL3 promotes metastasis of colorectal Cancer via miR-1246/SPRED2/MAPK signaling pathway. J Exp Clin Cancer Res. 2019. https://doi.org/10.1186/s13046-019-1408-4.

    Article  PubMed  PubMed Central  Google Scholar 

  289. Zhao H, et al. m6A regulators is differently expressed and correlated with immune response of esophageal cancer. Front Cell Dev Biol. 2021. https://doi.org/10.3389/fcell.2021.650023.

    Article  PubMed  PubMed Central  Google Scholar 

  290. Zhang B, et al. m6A regulator-mediated methylation modification patterns and tumor microenvironment infiltration characterization in gastric cancer. Mol Cancer. 2020. https://doi.org/10.1186/s12943-020-01170-0.

    Article  PubMed  PubMed Central  Google Scholar 

  291. Wang X, Liu J, Wang D, Feng M, Wu X. Epigenetically regulated gene expression profiles reveal four molecular subtypes with prognostic and therapeutic implications in colorectal cancer. Brief Bioinform. 2021. https://doi.org/10.1093/bib/bbaa309.

    Article  PubMed  PubMed Central  Google Scholar 

  292. Weng S, et al. Epigenetically regulated gene expression profiles decipher four molecular subtypes with prognostic and therapeutic implications in gastric cancer. Clin Epigenetics. 2023. https://doi.org/10.1186/s13148-023-01478-w.

    Article  PubMed  PubMed Central  Google Scholar 

  293. Kuang C, et al. Pembrolizumab plus azacitidine in patients with chemotherapy refractory metastatic colorectal cancer: a single-arm phase 2 trial and correlative biomarker analysis. Clin Epigenetics. 2022. https://doi.org/10.1186/s13148-021-01226-y.

    Article  PubMed  PubMed Central  Google Scholar 

  294. Li F, et al. Expression patterns of glycosylation regulators define tumor microenvironment and immunotherapy in gastric cancer. Front Cell Dev Biol. 2022. https://doi.org/10.3389/fcell.2022.811075.

    Article  PubMed  PubMed Central  Google Scholar 

  295. Wang Z, et al. Abrogation of USP7 is an alternative strategy to downregulate PD-L1 and sensitize gastric cancer cells to T cells killing. Acta Pharmaceutica Sinica B. 2021;11:694–707. https://doi.org/10.1016/j.apsb.2020.11.005.

    Article  CAS  PubMed  Google Scholar 

  296. Ni W, et al. Long noncoding RNA GAS5 inhibits progression of colorectal cancer by interacting with and triggering YAP phosphorylation and degradation and is negatively regulated by the m6A reader YTHDF3. Mol Cancer. 2019. https://doi.org/10.1186/s12943-019-1079-y.

    Article  PubMed  PubMed Central  Google Scholar 

  297. Ni W, et al. Targeting cholesterol biosynthesis promotes anti-tumor immunity by inhibiting long noncoding RNA SNHG29-mediated YAP activation. Mol Ther. 2021;29:2995–3010. https://doi.org/10.1016/j.ymthe.2021.05.012.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  298. Yu H, et al. Long noncoding RNA MIR4435-2HG suppresses colorectal cancer initiation and progression by reprogramming neutrophils. Cancer Immunol Res. 2022;10:1095–110. https://doi.org/10.1158/2326-6066.cir-21-1011.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  299. Xia C, et al. Insulin-like growth factor 2 mRNA-binding protein 2-stabilized long non-coding RNA Taurine up-regulated gene 1 (TUG1) promotes cisplatin-resistance of colorectal cancer via modulating autophagy. Bioengineered. 2022;13:2450–69. https://doi.org/10.1080/21655979.2021.2012918.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  300. Ding N, et al. A tumor-suppressive molecular axis EP300/circRERE/ miR-6837–3p/MAVS activates type I IFN pathway and antitumor immunity to suppress colorectal cancer. Clin Cancer Res. 2023. https://doi.org/10.1158/1078-0432.ccr-22-3836.

    Article  PubMed  PubMed Central  Google Scholar 

  301. Kwon M, et al. Determinants of response and intrinsic resistance to PD-1 blockade in microsatellite instability-high gastric cancer. Cancer Discov. 2021;11:2168–85. https://doi.org/10.1158/2159-8290.cd-21-0219.

    Article  CAS  PubMed  Google Scholar 

  302. Chen X, et al. Neoadjuvant sintilimab and chemotherapy in patients with potentially resectable esophageal squamous cell carcinoma (KEEP-G 03): an open-label, single-arm, phase 2 trial. J Immunother Cancer. 2023. https://doi.org/10.1136/jitc-2022-005830.

    Article  PubMed  PubMed Central  Google Scholar 

  303. Wang F, et al. Evaluation of POLE and POLD1 mutations as biomarkers for immunotherapy outcomes across multiple cancer types. JAMA Oncol. 2019;5:1504–6. https://doi.org/10.1001/jamaoncol.2019.2963.

    Article  PubMed  PubMed Central  Google Scholar 

  304. Gong J, Wang C, Lee PP, Chu P, Fakih M. Response to PD-1 blockade in microsatellite stable metastatic colorectal cancer harboring a POLE mutation. J Natl Comp Cancer Net. 2017;15:142–7. https://doi.org/10.6004/jnccn.2017.0016.

    Article  Google Scholar 

  305. Le DT, et al. Phase II open-label study of pembrolizumab in treatment-refractory, microsatellite instability-high/mismatch repair-deficient metastatic colorectal cancer: KEYNOTE-164. J Clin Oncol. 2020;38:11. https://doi.org/10.1200/jco.19.02107.

    Article  CAS  PubMed  Google Scholar 

  306. Jiao X, et al. A genomic mutation signature predicts the clinical outcomes of immunotherapy and characterizes immunophenotypes in gastrointestinal cancer. Npj Precis Oncol. 2021;5:36. https://doi.org/10.1038/s41698-021-00172-5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  307. Zhuo N, et al. Characteristics and prognosis of acquired resistance to immune checkpoint inhibitors in gastrointestinal cancer. JAMA Netw Open. 2022;5: e224637. https://doi.org/10.1001/jamanetworkopen.2022.4637.

    Article  PubMed  PubMed Central  Google Scholar 

  308. Albacker LA, et al. Loss of function JAK1 mutations occur at high frequency in cancers with microsatellite instability and are suggestive of immune evasion. PLoS ONE. 2017;12: e0176181. https://doi.org/10.1371/journal.pone.0176181.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  309. Stein A, et al. PD-L1 targeting and subclonal immune escape mediated by PD-L1 mutations in metastatic colorectal cancer. J Immunother Cancer. 2021. https://doi.org/10.1136/jitc-2021-002844.

    Article  PubMed  PubMed Central  Google Scholar 

  310. Du W, Frankel TL, Green M, Zou W. IFNgamma signaling integrity in colorectal cancer immunity and immunotherapy. Cell Mol Immunol. 2022;19:23–32. https://doi.org/10.1038/s41423-021-00735-3.

    Article  CAS  PubMed  Google Scholar 

  311. Du W, et al. Loss of optineurin drives cancer immune evasion via palmitoylation-dependent IFNGR1 lysosomal sorting and degradation. Cancer Discov. 2021;11:1826–43. https://doi.org/10.1158/2159-8290.CD-20-1571.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  312. Lv C, Yuan D, Cao Y. Downregulation of interferon-gamma receptor expression endows resistance to anti-programmed death protein 1 therapy in colorectal cancer. J Pharmacol Exp Ther. 2021;376:21–8. https://doi.org/10.1124/jpet.120.000284.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  313. Koyama S, et al. Adaptive resistance to therapeutic PD-1 blockade is associated with upregulation of alternative immune checkpoints. Nat Commun. 2016;7:10501. https://doi.org/10.1038/ncomms10501.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  314. Woo SR, et al. Immune inhibitory molecules LAG-3 and PD-1 synergistically regulate T-cell function to promote tumoral immune escape. Cancer Res. 2012;72:917–27. https://doi.org/10.1158/0008-5472.CAN-11-1620.

    Article  CAS  PubMed  Google Scholar 

  315. Maruhashi T, Sugiura D, Okazaki IM, Okazaki T. LAG-3: from molecular functions to clinical applications. J Immunother Cancer. 2020. https://doi.org/10.1136/jitc-2020-001014.

    Article  PubMed  PubMed Central  Google Scholar 

  316. Garralda E, et al. A first-in-human study of the anti-LAG-3 antibody favezelimab plus pembrolizumab in previously treated, advanced microsatellite stable colorectal cancer. ESMO Open. 2022;7: 100639. https://doi.org/10.1016/j.esmoop.2022.100639.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  317. Klapholz M, Drage MG, Srivastava A, Anderson AC. Presence of Tim3(+) and PD-1(+) CD8(+) T cells identifies microsatellite stable colorectal carcinomas with immune exhaustion and distinct clinicopathological features. J Pathol. 2022;257:186–97. https://doi.org/10.1002/path.5877.

    Article  CAS  PubMed  Google Scholar 

  318. Kong Y, et al. T-Cell Immunoglobulin and ITIM Domain (TIGIT) Associates with CD8+ T-Cell Exhaustion and Poor Clinical Outcome in AML Patients. Clin Cancer Res. 2016;22:3057–66. https://doi.org/10.1158/1078-0432.CCR-15-2626.

    Article  CAS  PubMed  Google Scholar 

  319. Johnston RJ, et al. The immunoreceptor TIGIT regulates antitumor and antiviral CD8(+) T cell effector function. Cancer Cell. 2014;26:923–37. https://doi.org/10.1016/j.ccell.2014.10.018.

    Article  CAS  PubMed  Google Scholar 

  320. Thibaudin M, et al. Targeting PD-L1 and TIGIT could restore intratumoral CD8 T cell function in human colorectal cancer. Cancer Immunol Immunother. 2022;71:2549–63. https://doi.org/10.1007/s00262-022-03182-9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  321. Chen Y, et al. A transformable supramolecular bispecific cell engager for augmenting natural killer and T cell-based cancer immunotherapy. Adv Mater. 2024;36: e2306736. https://doi.org/10.1002/adma.202306736.

    Article  CAS  PubMed  Google Scholar 

  322. Rosenthal R, et al. Neoantigen-directed immune escape in lung cancer evolution. Nature. 2019;567:479–85. https://doi.org/10.1038/s41586-019-1032-7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  323. Anagnostou V, et al. Evolution of neoantigen landscape during immune checkpoint blockade in non-small cell lung cancer. Cancer Discov. 2017;7:264–76. https://doi.org/10.1158/2159-8290.CD-16-0828.

    Article  CAS  PubMed  Google Scholar 

  324. Hirschhorn D, et al. T cell immunotherapies engage neutrophils to eliminate tumor antigen escape variants. Cell. 2023;186:1432–47. https://doi.org/10.1016/j.cell.2023.03.007.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  325. Kabacaoglu D, Ciecielski KJ, Ruess DA, Algul H. Immune checkpoint inhibition for pancreatic ductal adenocarcinoma: current limitations and future options. Front Immunol. 2018;9:1878. https://doi.org/10.3389/fimmu.2018.01878.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  326. Westcott PMK, et al. Low neoantigen expression and poor T-cell priming underlie early immune escape in colorectal cancer. Nat Cancer. 2021;2:1071–85. https://doi.org/10.1038/s43018-021-00247-z.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  327. Schumacher TN, Schreiber RD. Neoantigens in cancer immunotherapy. Science. 2015;348:69–74. https://doi.org/10.1126/science.aaa4971.

    Article  CAS  PubMed  Google Scholar 

  328. Grasso CS, et al. Genetic mechanisms of immune evasion in colorectal cancer. Cancer Discov. 2018;8:730–49. https://doi.org/10.1158/2159-8290.CD-17-1327.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  329. Le DT, et al. Mismatch repair deficiency predicts response of solid tumors to PD-1 blockade. Science. 2017;357:409–13. https://doi.org/10.1126/science.aan6733.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  330. Le DT, et al. PD-1 blockade in tumors with mismatch-repair deficiency. N Engl J Med. 2015;372:2509–20. https://doi.org/10.1056/NEJMoa1500596.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  331. Liu F, et al. Prevalence and associations of beta2-microglobulin mutations in MSI-H/dMMR cancers. Oncologist. 2023;28:e136–44. https://doi.org/10.1093/oncolo/oyac268.

    Article  PubMed  PubMed Central  Google Scholar 

  332. Smith ME, Bodmer WF, Bodmer JG. Selective loss of HLA-A, B, C locus products in colorectal adenocarcinoma. Lancet. 1988;1:823–4. https://doi.org/10.1016/s0140-6736(88)91682-0.

    Article  CAS  PubMed  Google Scholar 

  333. Tikidzhieva A, et al. Microsatellite instability and Beta2-Microglobulin mutations as prognostic markers in colon cancer: results of the FOGT-4 trial. Br J Cancer. 2012;106:1239–45. https://doi.org/10.1038/bjc.2012.53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  334. Bernal M, Ruiz-Cabello F, Concha A, Paschen A, Garrido F. Implication of the beta2-microglobulin gene in the generation of tumor escape phenotypes. Cancer Immunol Immunother. 2012;61:1359–71. https://doi.org/10.1007/s00262-012-1321-6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  335. Kloor M, et al. Beta2-microglobulin mutations in microsatellite unstable colorectal tumors. Int J Cancer. 2007;121:454–8. https://doi.org/10.1002/ijc.22691.

    Article  CAS  PubMed  Google Scholar 

  336. Hargadon KM. Genetic dysregulation of immunologic and oncogenic signaling pathways associated with tumor-intrinsic immune resistance: a molecular basis for combination targeted therapy-immunotherapy for cancer. Cell Mol Life Sci. 2023;80:40. https://doi.org/10.1007/s00018-023-04689-9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  337. Kim TK, Herbst RS, Chen L. Defining and understanding adaptive resistance in cancer immunotherapy. Trends Immunol. 2018;39:624–31. https://doi.org/10.1016/j.it.2018.05.001.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  338. Farhood B, Najafi M, Mortezaee K. CD8(+) cytotoxic T lymphocytes in cancer immunotherapy: a review. J Cell Physiol. 2019;234:8509–21. https://doi.org/10.1002/jcp.27782.

    Article  CAS  PubMed  Google Scholar 

  339. Wei SC, et al. Distinct cellular mechanisms underlie anti-CTLA-4 and anti-PD-1 checkpoint blockade. Cell. 2017;170:1120–33. https://doi.org/10.1016/j.cell.2017.07.024.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  340. Fischer K, et al. Inhibitory effect of tumor cell-derived lactic acid on human T cells. Blood. 2007;109:3812–9. https://doi.org/10.1182/blood-2006-07-035972.

    Article  CAS  PubMed  Google Scholar 

  341. Chida K, et al. A Low tumor mutational burden and PTEN mutations are predictors of a negative response to PD-1 blockade in MSI-H/dMMR gastrointestinal tumors. Clin Cancer Res. 2021;27:3714–24. https://doi.org/10.1158/1078-0432.CCR-21-0401.

    Article  CAS  PubMed  Google Scholar 

  342. Vidotto T, et al. Emerging role of PTEN loss in evasion of the immune response to tumours. Br J Cancer. 2020;122:1732–43. https://doi.org/10.1038/s41416-020-0834-6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  343. Piro G, et al. Revising PTEN in the era of immunotherapy: new perspectives for an old story. Cancers. 2019. https://doi.org/10.3390/cancers11101525.

    Article  PubMed  PubMed Central  Google Scholar 

  344. Conciatori F, et al. PTEN Function at the interface between cancer and tumor microenvironment: implications for response to immunotherapy. Int J Mol Sci. 2020. https://doi.org/10.3390/ijms21155337.

    Article  PubMed  PubMed Central  Google Scholar 

  345. Elez E, et al. A comprehensive biomarker analysis of microsatellite unstable/mismatch repair deficient colorectal cancer cohort treated with immunotherapy. Int J Mol Sci. 2022. https://doi.org/10.3390/ijms24010118.

    Article  PubMed  PubMed Central  Google Scholar 

  346. Bortolomeazzi M, et al. Immunogenomics of colorectal cancer response to checkpoint blockade: analysis of the KEYNOTE 177 trial and validation cohorts. Gastroenterology. 2021;161:1179–93. https://doi.org/10.1053/j.gastro.2021.06.064.

    Article  CAS  PubMed  Google Scholar 

  347. Gabrilovich DI, Ostrand-Rosenberg S, Bronte V. Coordinated regulation of myeloid cells by tumours. Nat Rev Immunol. 2012;12:253–68. https://doi.org/10.1038/nri3175.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  348. Gabrilovich DI, Nagaraj S. Myeloid-derived suppressor cells as regulators of the immune system. Nat Rev Immunol. 2009;9:162–74. https://doi.org/10.1038/nri2506.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  349. Marvel D, Gabrilovich DI. Myeloid-derived suppressor cells in the tumor microenvironment: expect the unexpected. J Clin Invest. 2015;125:3356–64. https://doi.org/10.1172/JCI80005.

    Article  PubMed  PubMed Central  Google Scholar 

  350. Zhang S, et al. The role of myeloid-derived suppressor cells in patients with solid tumors: a meta-analysis. PLoS ONE. 2016;11: e0164514. https://doi.org/10.1371/journal.pone.0164514.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  351. Limagne E, et al. Accumulation of MDSC and Th17 Cells in patients with metastatic colorectal cancer predicts the efficacy of a FOLFOX-bevacizumab drug treatment regimen. Cancer Res. 2016;76:5241–52. https://doi.org/10.1158/0008-5472.CAN-15-3164.

    Article  CAS  PubMed  Google Scholar 

  352. Markowitz J, et al. Patients with pancreatic adenocarcinoma exhibit elevated levels of myeloid-derived suppressor cells upon progression of disease. Cancer Immunol Immunother. 2015;64:149–59. https://doi.org/10.1007/s00262-014-1618-8.

    Article  CAS  PubMed  Google Scholar 

  353. Karakasheva TA, et al. CD38-expressing myeloid-derived suppressor cells promote tumor growth in a murine model of esophageal cancer. Cancer Res. 2015;75:4074–85. https://doi.org/10.1158/0008-5472.CAN-14-3639.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  354. Chen MF, Chen PT, Kuan FC, Chen WC. The predictive value of pretreatment neutrophil-to-lymphocyte ratio in esophageal squamous cell carcinoma. Ann Surg Oncol. 2019;26:190–9. https://doi.org/10.1245/s10434-018-6944-1.

    Article  PubMed  Google Scholar 

  355. Tsutsumi C, et al. Tumor-infiltrating monocytic myeloid-derived suppressor cells contribute to the development of an immunosuppressive tumor microenvironment in gastric cancer. Gastric Cancer. 2024;27:248–62. https://doi.org/10.1007/s10120-023-01456-4.

    Article  CAS  PubMed  Google Scholar 

  356. Xiao F, Dai Y, Hu Y, Lu M, Dai Q. Expression profile analysis identifies IER3 to predict overall survival and promote lymph node metastasis in tongue cancer. Cancer Cell Int. 2019;19:307. https://doi.org/10.1186/s12935-019-1028-2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  357. Zhang Y, et al. Impaired apoptosis, extended duration of immune responses, and a lupus-like autoimmune disease in IEX-1-transgenic mice. Proc Natl Acad Sci U S A. 2002;99:878–83. https://doi.org/10.1073/pnas.022326699.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  358. Li N, et al. Subtypes of MDSCs in mechanisms and prognosis of gastric cancer and are inhibited by epirubicin and paclitaxel. Discov Med. 2018;25:99–112.

    PubMed  Google Scholar 

  359. Lin F, et al. The effect of low-dose chemotherapy on the tumor microenvironment and its antitumor activity combined with anti-PD-1 antibody. Immunotherapy. 2022. https://doi.org/10.2217/imt-2021-0018.

    Article  PubMed  Google Scholar 

  360. Chen H, et al. PD-1 inhibitor combined with apatinib modulate the tumor microenvironment and potentiate anti-tumor effect in mice bearing gastric cancer. Int Immunopharmacol. 2021;99: 107929. https://doi.org/10.1016/j.intimp.2021.107929.

    Article  CAS  PubMed  Google Scholar 

  361. Tobin RP, et al. Targeting myeloid-derived suppressor cells using all-trans retinoic acid in melanoma patients treated with Ipilimumab. Int Immunopharmacol. 2018;63:282–91. https://doi.org/10.1016/j.intimp.2018.08.007.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  362. Bauer R, et al. Blockade of myeloid-derived suppressor cell expansion with all-trans retinoic acid increases the efficacy of antiangiogenic therapy. Cancer Res. 2018;78:3220–32. https://doi.org/10.1158/0008-5472.CAN-17-3415.

    Article  CAS  PubMed  Google Scholar 

  363. Taniyama D, et al. Long-term follow-up study of gastric adenoma; tumor-associated macrophages are associated to carcinoma development in gastric adenoma. Gastric Cancer. 2017;20:929–39. https://doi.org/10.1007/s10120-017-0713-x.

    Article  CAS  PubMed  Google Scholar 

  364. Zhou Z, et al. Research progress on tumour-associated macrophages in gastric cancer (Review). Oncol Rep. 2021. https://doi.org/10.3892/or.2021.7986.

    Article  PubMed  PubMed Central  Google Scholar 

  365. Wang Z, et al. Tumor-associated macrophages regulate gastric cancer cell invasion and metastasis through TGFbeta2/NF-kappaB/Kindlin-2 axis. Chin J Cancer Res. 2020;32:72–88. https://doi.org/10.21147/j.issn.1000-9604.2020.01.09.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  366. Zhang AZ, et al. Immune infiltration in gastric cancer microenvironment and its clinical significance. Front Cell Dev Biol. 2021;9: 762029. https://doi.org/10.3389/fcell.2021.762029.

    Article  PubMed  Google Scholar 

  367. Zhao Y, Bai Y, Shen M, Li Y. Therapeutic strategies for gastric cancer targeting immune cells: future directions. Front Immunol. 2022;13: 992762. https://doi.org/10.3389/fimmu.2022.992762.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  368. Wu K, et al. Redefining tumor-associated macrophage subpopulations and functions in the tumor microenvironment. Front Immunol. 2020;11:1731. https://doi.org/10.3389/fimmu.2020.01731.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  369. Cao Y, et al. Blockade of V-domain immunoglobulin suppressor of T-cell activation reprograms tumour-associated macrophages and improves efficacy of PD-1 inhibitor in gastric cancer. Clin Transl Med. 2024;14: e1578. https://doi.org/10.1002/ctm2.1578.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  370. Di Pilato M, et al. Targeting the CBM complex causes T(reg) cells to prime tumours for immune checkpoint therapy. Nature. 2019;570:112–6. https://doi.org/10.1038/s41586-019-1215-2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  371. Toomer KH, Malek TR. Cytokine signaling in the development and homeostasis of regulatory T cells. Cold Spring Harb Perspect Biol. 2018. https://doi.org/10.1101/cshperspect.a028597.

    Article  PubMed  PubMed Central  Google Scholar 

  372. Guo Y, et al. Blockade of TNF-alpha/TNFR2 signalling suppresses colorectal cancer and enhances the efficacy of anti-PD1 immunotherapy by decreasing CCR8+ T regulatory cells. J Mol Cell Biol. 2023. https://doi.org/10.1093/jmcb/mjad067.

    Article  PubMed  PubMed Central  Google Scholar 

  373. Wing JB, Tanaka A, Sakaguchi S. Human FOXP3(+) regulatory T cell heterogeneity and function in autoimmunity and cancer. Immunity. 2019;50:302–16. https://doi.org/10.1016/j.immuni.2019.01.020.

    Article  CAS  PubMed  Google Scholar 

  374. Villarreal DO, et al. Targeting CCR8 induces protective antitumor immunity and enhances vaccine-induced responses in colon cancer. Cancer Res. 2018;78:5340–8. https://doi.org/10.1158/0008-5472.CAN-18-1119.

    Article  CAS  PubMed  Google Scholar 

  375. Xue R, et al. Liver tumour immune microenvironment subtypes and neutrophil heterogeneity. Nature. 2022;612:141–7. https://doi.org/10.1038/s41586-022-05400-x.

    Article  CAS  PubMed  Google Scholar 

  376. Hu J, et al. Tumor microenvironment remodeling after neoadjuvant immunotherapy in non-small cell lung cancer revealed by single-cell RNA sequencing. Genome Med. 2023;15:14. https://doi.org/10.1186/s13073-023-01164-9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  377. Chandler C, Liu T, Buckanovich R, Coffman LG. The double edge sword of fibrosis in cancer. Transl Res. 2019;209:55–67. https://doi.org/10.1016/j.trsl.2019.02.006.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  378. Gomes CM. The dual role of mesenchymal stem cells in tumor progression. Stem Cell Res Ther. 2013;4:42. https://doi.org/10.1186/scrt189.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  379. Cascio S, et al. Cancer-associated MSC drive tumor immune exclusion and resistance to immunotherapy, which can be overcome by Hedgehog inhibition. Sci Adv. 2021;7:eabi5790. https://doi.org/10.1126/sciadv.abi5790.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  380. Huang C, et al. Gastric cancer mesenchymal stem cells via the CXCR2/HK2/PD-L1 pathway mediate immunosuppression. Gastric Cancer. 2023;26:691–707. https://doi.org/10.1007/s10120-023-01405-1.

    Article  CAS  PubMed  Google Scholar 

  381. Lin Z, et al. Mesenchymal stem cell-derived exosomes in cancer therapy resistance: recent advances and therapeutic potential. Mol Cancer. 2022;21:179. https://doi.org/10.1186/s12943-022-01650-5.

    Article  PubMed  PubMed Central  Google Scholar 

  382. Weng Z, et al. Therapeutic roles of mesenchymal stem cell-derived extracellular vesicles in cancer. J Hematol Oncol. 2021;14:136. https://doi.org/10.1186/s13045-021-01141-y.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  383. Lyu T, et al. Exosomes from BM-MSCs promote acute myeloid leukemia cell proliferation, invasion and chemoresistance via upregulation of S100A4. Exp Hematol Oncol. 2021;10:24. https://doi.org/10.1186/s40164-021-00220-7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  384. Bliss SA, et al. Mesenchymal stem cell-derived exosomes stimulate cycling quiescence and early breast cancer dormancy in bone marrow. Cancer Res. 2016;76:5832–44. https://doi.org/10.1158/0008-5472.CAN-16-1092.

    Article  CAS  PubMed  Google Scholar 

  385. Vesely MD, Zhang T, Chen L. Resistance mechanisms to anti-PD cancer immunotherapy. Annu Rev Immunol. 2022;40:45–74. https://doi.org/10.1146/annurev-immunol-070621-030155.

    Article  CAS  PubMed  Google Scholar 

  386. Tie Y, Tang F, Wei YQ, Wei XW. Immunosuppressive cells in cancer: mechanisms and potential therapeutic targets. J Hematol Oncol. 2022;15:61. https://doi.org/10.1186/s13045-022-01282-8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  387. Schalper KA, et al. Elevated serum interleukin-8 is associated with enhanced intratumor neutrophils and reduced clinical benefit of immune-checkpoint inhibitors. Nat Med. 2020;26:688–92. https://doi.org/10.1038/s41591-020-0856-x.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  388. Cerezo-Wallis D, et al. Midkine rewires the melanoma microenvironment toward a tolerogenic and immune-resistant state. Nat Med. 2020;26:1865–77. https://doi.org/10.1038/s41591-020-1073-3.

    Article  CAS  PubMed  Google Scholar 

  389. Batlle E, Massague J. Transforming Growth factor-beta signaling in immunity and cancer. Immunity. 2019;50:924–40. https://doi.org/10.1016/j.immuni.2019.03.024.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  390. Lebrun JJ. The dual role of TGFbeta in human cancer: from tumor suppression to cancer metastasis. ISRN Mol Biol. 2012;2012: 381428. https://doi.org/10.5402/2012/381428.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  391. Derynck R, Turley SJ, Akhurst RJ. TGFbeta biology in cancer progression and immunotherapy. Nat Rev Clin Oncol. 2021;18:9–34. https://doi.org/10.1038/s41571-020-0403-1.

    Article  PubMed  Google Scholar 

  392. Gonzalez-Molina J, et al. Chemotherapy as a regulator of extracellular matrix-cell communication: Implications in therapy resistance. Semin Cancer Biol. 2022;86:224–36. https://doi.org/10.1016/j.semcancer.2022.03.012.

    Article  CAS  PubMed  Google Scholar 

  393. Huang J, et al. Extracellular matrix and its therapeutic potential for cancer treatment. Signal Transduct Target Ther. 2021;6:153. https://doi.org/10.1038/s41392-021-00544-0.

    Article  PubMed  PubMed Central  Google Scholar 

  394. Ramjiawan RR, Griffioen AW, Duda DG. Anti-angiogenesis for cancer revisited: Is there a role for combinations with immunotherapy? Angiogenesis. 2017;20:185–204. https://doi.org/10.1007/s10456-017-9552-y.

    Article  PubMed  PubMed Central  Google Scholar 

  395. Correction for Ebbing et al., Stromal-derived interleukin 6 drives epithelial-to-mesenchymal transition and therapy resistance in esophageal adenocarcinoma. Proc Natl Acad Sci U S A 118, https://doi.org/10.1073/pnas.2113728118 (2021).

  396. Fang P, et al. Immunotherapy resistance in esophageal cancer: Possible mechanisms and clinical implications. Front Immunol. 2022;13: 975986. https://doi.org/10.3389/fimmu.2022.975986.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  397. Bell HN, et al. Microenvironmental ammonia enhances T cell exhaustion in colorectal cancer. Cell Metab. 2023;35:134–49. https://doi.org/10.1016/j.cmet.2022.11.013.

    Article  CAS  PubMed  Google Scholar 

  398. Ganguly K, Kimmelman AC. Reprogramming of tissue metabolism during cancer metastasis. Trends Cancer. 2023;9:461–71. https://doi.org/10.1016/j.trecan.2023.02.005.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  399. Wang ZH, Peng WB, Zhang P, Yang XP, Zhou Q. Lactate in the tumour microenvironment: from immune modulation to therapy. EBioMedicine. 2021;73: 103627. https://doi.org/10.1016/j.ebiom.2021.103627.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  400. Chang CH, et al. Metabolic competition in the tumor microenvironment is a driver of cancer progression. Cell. 2015;162:1229–41. https://doi.org/10.1016/j.cell.2015.08.016.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  401. Hartmann FJ, et al. Single-cell metabolic profiling of human cytotoxic T cells. Nat Biotechnol. 2021;39:186–97. https://doi.org/10.1038/s41587-020-0651-8.

    Article  CAS  PubMed  Google Scholar 

  402. Long GV, et al. Epacadostat plus pembrolizumab versus placebo plus pembrolizumab in patients with unresectable or metastatic melanoma (ECHO-301/KEYNOTE-252): a phase 3, randomised, double-blind study. Lancet Oncol. 2019;20:1083–97. https://doi.org/10.1016/S1470-2045(19)30274-8.

    Article  CAS  PubMed  Google Scholar 

  403. Lian X, et al. Immunometabolic rewiring in tumorigenesis and anti-tumor immunotherapy. Mol Cancer. 2022;21:27. https://doi.org/10.1186/s12943-021-01486-5.

    Article  PubMed  PubMed Central  Google Scholar 

  404. Sundar R, et al. Epigenetic promoter alterations in GI tumour immune-editing and resistance to immune checkpoint inhibition. Gut. 2022;71:1277–88. https://doi.org/10.1136/gutjnl-2021-324420.

    Article  CAS  PubMed  Google Scholar 

  405. Han Y, et al. IL-1beta-associated NNT acetylation orchestrates iron-sulfur cluster maintenance and cancer immunotherapy resistance. Mol Cell. 2023;83:1887–902. https://doi.org/10.1016/j.molcel.2023.05.011.

    Article  CAS  PubMed  Google Scholar 

  406. Tao S, Liang S, Zeng T, Yin D. Epigenetic modification-related mechanisms of hepatocellular carcinoma resistance to immune checkpoint inhibition. Front Immunol. 2022;13:1043667. https://doi.org/10.3389/fimmu.2022.1043667.

    Article  CAS  PubMed  Google Scholar 

  407. Kim HJ, Cantor H, Cosmopoulos K. Overcoming immune checkpoint blockade resistance via EZH2 inhibition. Trends Immunol. 2020;41:948–63. https://doi.org/10.1016/j.it.2020.08.010.

    Article  CAS  PubMed  Google Scholar 

  408. Tian Y, et al. Histone deacetylase HDAC8 promotes insulin resistance and beta-catenin activation in NAFLD-associated hepatocellular carcinoma. Cancer Res. 2015;75:4803–16. https://doi.org/10.1158/0008-5472.CAN-14-3786.

    Article  CAS  PubMed  Google Scholar 

  409. Yang W, et al. A selective HDAC8 inhibitor potentiates antitumor immunity and efficacy of immune checkpoint blockade in hepatocellular carcinoma. Sci Transl Med. 2021. https://doi.org/10.1126/scitranslmed.aaz6804.

    Article  PubMed  PubMed Central  Google Scholar 

  410. Dhupar R, et al. Targeting immune checkpoints in esophageal cancer: a high mutational load tumor. Ann Thorac Surg. 2017;103:1340–9. https://doi.org/10.1016/j.athoracsur.2016.12.011.

    Article  PubMed  Google Scholar 

  411. Thrift AP. Global burden and epidemiology of Barrett oesophagus and oesophageal cancer. Nat Rev Gastroenterol Hepatol. 2021;18:432–43. https://doi.org/10.1038/s41575-021-00419-3.

    Article  PubMed  Google Scholar 

  412. Belle CJ, Lonie JM, Brosda S, Barbour AP. Tumour microenvironment influences response to treatment in oesophageal adenocarcinoma. Front Immunol. 2023;14:1330635. https://doi.org/10.3389/fimmu.2023.1330635.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  413. Liu Z, et al. Genomic and transcriptomic profiling of hepatoid adenocarcinoma of the stomach. Oncogene. 2021;40:5705–17. https://doi.org/10.1038/s41388-021-01976-2.

    Article  CAS  PubMed  Google Scholar 

  414. Li W, et al. Effect of immune checkpoint inhibitors plus chemotherapy on advanced gastric cancer patients with elevated serum AFP or hepatoid adenocarcinoma. Cancer Manag Res. 2020;12:11113–9. https://doi.org/10.2147/cmar.S276969.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  415. Jia K, et al. Multiplex immunohistochemistry defines the tumor immune microenvironment and immunotherapeutic outcome in CLDN18.2-positive gastric cancer. BMC Med. 2022;20:223. https://doi.org/10.1186/s12916-022-02421-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  416. Jin M-H, et al. WEE1 inhibition reverses trastuzumab resistance in HER2-positive cancers. Gastric Cancer. 2021;24:1003–20. https://doi.org/10.1007/s10120-021-01176-7.

    Article  CAS  PubMed  Google Scholar 

  417. Pang K, et al. Research progress of therapeutic effects and drug resistance of immunotherapy based on PD-1/PD-L1 blockade. Drug Resist Updat. 2023;66: 100907. https://doi.org/10.1016/j.drup.2022.100907.

    Article  CAS  PubMed  Google Scholar 

  418. Baxter MA, Middleton F, Cagney HP, Petty RD. Resistance to immune checkpoint inhibitors in advanced gastro-oesophageal cancers. Br J Cancer. 2021;125:1068–79. https://doi.org/10.1038/s41416-021-01425-7.

    Article  PubMed  PubMed Central  Google Scholar 

  419. Mathew M, Enzler T, Shu CA, Rizvi NA. Combining chemotherapy with PD-1 blockade in NSCLC. Pharmacol Ther. 2018;186:130–7. https://doi.org/10.1016/j.pharmthera.2018.01.003.

    Article  CAS  PubMed  Google Scholar 

  420. Liu Y, et al. Abscopal effect of radiotherapy combined with immune checkpoint inhibitors. J Hematol Oncol. 2018. https://doi.org/10.1186/s13045-018-0647-8.

    Article  PubMed  PubMed Central  Google Scholar 

  421. Twyman-Saint Victor C, et al. Radiation and dual checkpoint blockade activate non-redundant immune mechanisms in cancer. Nature. 2015;520:373-+. https://doi.org/10.1038/nature14292.

    Article  CAS  PubMed  Google Scholar 

  422. Deng L, et al. Irradiation and anti-PD-L1 treatment synergistically promote antitumor immunity in mice. J Clin Invest. 2014;124:687–95. https://doi.org/10.1172/jci67313.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  423. Wang X, et al. Suppression of type I IFN signaling in tumors mediates resistance to anti-PD-1 treatment that can be overcome by radiotherapy. Can Res. 2017;77:839–50. https://doi.org/10.1158/0008-5472.can-15-3142.

    Article  CAS  Google Scholar 

  424. Parikh AR, et al. Radiation therapy enhances immunotherapy response in microsatellite stable colorectal and pancreatic adenocarcinoma in a phase II trial. Nat Cancer. 2021;2:1124-+. https://doi.org/10.1038/s43018-021-00269-7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  425. Zhao W, et al. Radiotherapy plus camrelizumab and irinotecan for oligometastatic esophageal squamous cell carcinoma patients after first-line immunotherapy plus chemotherapy failure: An open-label, single-arm, phase II trial. Radiother Oncol. 2023. https://doi.org/10.1016/j.radonc.2023.109679.

    Article  PubMed  Google Scholar 

  426. Herrera FG, et al. Low-dose radiotherapy reverses tumor immune desertification and resistance to immunotherapy. Cancer Discov. 2022;12:108–33. https://doi.org/10.1158/2159-8290.cd-21-0003.

    Article  CAS  PubMed  Google Scholar 

  427. Sharma P, Allison JP. Immune checkpoint targeting in cancer therapy: toward combination strategies with curative potential. Cell. 2015;161:205–14. https://doi.org/10.1016/j.cell.2015.03.030.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  428. Allen E, et al. Combined antiangiogenic and anti-PD-L1 therapy stimulates tumor immunity through HEV formation. Sci Transl Med. 2017. https://doi.org/10.1126/scitranslmed.aak9679.

    Article  PubMed  PubMed Central  Google Scholar 

  429. Fukurnura D, Kloepper J, Amoozgar Z, Duda DG, Jain RK. Enhancing cancer immunotherapy using antiangiogenics: opportunities and challenges. Nat Rev Clin Oncol. 2018;15:325–40. https://doi.org/10.1038/nrclinonc.2018.29.

    Article  CAS  Google Scholar 

  430. Fukuoka S, et al. Regorafenib plus nivolumab in patients with advanced gastric or colorectal cancer: an open-label, dose-escalation, and dose-expansion phase Ib Trial (REGONIVO, EPOC1603). J Clin Oncol. 2020;38:2053-+. https://doi.org/10.1200/jco.19.03296.

    Article  CAS  PubMed  Google Scholar 

  431. Li N, et al. Efficacy and safety of apatinib for the treatment of AFP-producing gastric cancer. Transl Oncol. 2021. https://doi.org/10.1016/j.tranon.2020.101004.

    Article  PubMed  PubMed Central  Google Scholar 

  432. Kamei S, et al. Evaluation of VEGF and VEGF-C expression in gastric cancer cells producing α-fetoprotein. J Gastroenterol. 2003;38:540–7. https://doi.org/10.1007/s00535-002-1099-y.

    Article  CAS  PubMed  Google Scholar 

  433. Grenga I, et al. Anti-PD-L1/TGFI3R2 (M7824) fusion protein induces immunogenic modulation of human urothelial carcinoma cell lines, rendering them more susceptible to immune-mediated recognition and lysis. Urol Oncol Semin Original Investig. 2018. https://doi.org/10.1016/j.urolonc.2017.09.027.

    Article  Google Scholar 

  434. Yoo, C. et al. M7824 (MSB0011359C), a bifunctional fusion protein targeting PD-L1 and TGF-β, in Asian patients with pretreated biliary tract cancer: Preliminary results from a phase I trial. Ann Oncol 29 (2018).

  435. Lin C-C, et al. Bintrafusp Alfa, a Bifunctional Fusion Protein Targeting TGFβ and PD-L1, in Patients with Esophageal Squamous Cell Carcinoma: Results from a Phase 1 Cohort in Asia. Target Oncol. 2021;16:447–59. https://doi.org/10.1007/s11523-021-00810-9.

    Article  PubMed  PubMed Central  Google Scholar 

  436. Li A, et al. Prospects for combining immune checkpoint blockade with PARP inhibition. J Hematol Oncol. 2019. https://doi.org/10.1186/s13045-019-0784-8.

    Article  PubMed  PubMed Central  Google Scholar 

  437. Kwon M, et al. Phase II study of ceralasertib (AZD6738) in combination with durvalumab in patients with advanced gastric cancer. J Immunother Cancer. 2022. https://doi.org/10.1136/jitc-2022-005041.

    Article  PubMed  PubMed Central  Google Scholar 

  438. Klein C, Brinkmann U, Reichert JM, Kontermann RE. The present and future of bispecific antibodies for cancer therapy. Nat Rev Drug Discovery. 2024;23:301–19. https://doi.org/10.1038/s41573-024-00896-6.

    Article  CAS  PubMed  Google Scholar 

  439. Chen EX, et al. Effect of combined immune checkpoint inhibition vs best supportive care alone in patients with advanced colorectal cancer the canadian cancer trials group CO.26 study. Jama Oncol. 2020;6:831–8. https://doi.org/10.1001/jamaoncol.2020.0910.

    Article  PubMed  Google Scholar 

  440. Gao X, et al. Safety and antitumour activity of cadonilimab, an anti-PD-1/CTLA-4 bispecific antibody, for patients with advanced solid tumours (COMPASSION-03): a multicentre, open-label, phase 1b/2 trial. Lancet Oncol. 2023;24:1134–46. https://doi.org/10.1016/s1470-2045(23)00411-4.

    Article  CAS  PubMed  Google Scholar 

  441. Yap TA, et al. A Phase 1 First-in-Human Study of FS118, a Tetravalent Bispecific Antibody Targeting LAG-3 and PD-L1 in Patients with Advanced Cancer and PD-L1 Resistance. Clin Cancer Res. 2023;29:888–98. https://doi.org/10.1158/1078-0432.ccr-22-1449.

    Article  CAS  PubMed  Google Scholar 

  442. Chester C, Sanmamed MF, Wang J, Melero I. Immunotherapy targeting 4–1BB: mechanistic rationale, clinical results, and future strategies. Blood. 2018;131:49–57. https://doi.org/10.1182/blood-2017-06-741041.

    Article  CAS  PubMed  Google Scholar 

  443. Marin-Acevedo JA, Kimbrough EO, Manochakian R, Zhao Y, Lou Y. Immunotherapies targeting stimulatory pathways and beyond. J Hematol Oncol. 2021. https://doi.org/10.1186/s13045-021-01085-3.

    Article  PubMed  PubMed Central  Google Scholar 

  444. Pourakbari R, et al. Co-stimulatory agonists: an insight into the immunotherapy of cancer. Excli J. 2021;20:1055–85. https://doi.org/10.17179/excli2021-3522.

    Article  PubMed  PubMed Central  Google Scholar 

  445. Kim B-S, et al. Type 17 immunity promotes the exhaustion of CD8+T cells in cancer. J Immunother Cancer. 2021. https://doi.org/10.1136/jitc-2021-002603.

    Article  PubMed  PubMed Central  Google Scholar 

  446. Garlanda C, Mantovani A. Interleukin-1 in tumor progression, therapy, and prevention. Cancer Cell. 2021;39:1023–7. https://doi.org/10.1016/j.ccell.2021.04.011.

    Article  CAS  PubMed  Google Scholar 

  447. Liu C, et al. Blocking IL-17A enhances tumor response to anti-PD-1 immunotherapy in microsatellite stable colorectal cancer. J Immunother Cancer. 2021. https://doi.org/10.1136/jitc-2020-001895.

    Article  PubMed  PubMed Central  Google Scholar 

  448. Nagaoka K, et al. Deep immunophenotyping at the single-cell level identifies a combination of anti-IL-17 and checkpoint blockade as an effective treatment in a preclinical model of data-guided personalized immunotherapy. J Immunother Cancer. 2020. https://doi.org/10.1136/jitc-2020-001358.

    Article  PubMed  PubMed Central  Google Scholar 

  449. Davar D, et al. Fecal microbiota transplant overcomes resistance to anti-PD-1 therapy in melanoma patients. Science. 2021;371:595–602. https://doi.org/10.1126/science.abf3363.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  450. Baruch EN, et al. Fecal microbiota transplant promotes response in immunotherapy-refractory melanoma patients. Science. 2021;371:602–9. https://doi.org/10.1126/science.abb5920.

    Article  CAS  PubMed  Google Scholar 

  451. Zhao W, et al. Fecal microbiota transplantation plus tislelizumab and fruquintinib in refractory microsatellite stable metastatic colorectal cancer: an open-label, single-arm, phase II trial (RENMIN-215). EClinicalMedicine. 2023;66: 102315. https://doi.org/10.1016/j.eclinm.2023.102315.

    Article  PubMed  PubMed Central  Google Scholar 

  452. Routy B, et al. Author Correction: Fecal microbiota transplantation plus anti-PD-1 immunotherapy in advanced melanoma: a phase I trial. Nat Med. 2024;30:604. https://doi.org/10.1038/s41591-023-02650-8.

    Article  CAS  PubMed  Google Scholar 

  453. Qi C, et al. CT041 CAR T cell therapy for Claudin18.2-positive metastatic pancreatic cancer. J Hematol Oncol. 2023;16:102. https://doi.org/10.1186/s13045-023-01491-9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  454. Michaux A, et al. Clinical grade manufacture of CYAD-101, a NKG2D-based, first in class, non-gene-edited allogeneic CAR T-Cell therapy. J Immunother. 2022;45:150–61. https://doi.org/10.1097/CJI.0000000000000413.

    Article  CAS  PubMed  Google Scholar 

  455. Prenen H, et al. Updated data from alloSHRINK phase I first-in-human study evaluating CYAD-101, an innovative non-gene edited allogeneic CAR-T in mCRC. J Clin Oncol. 2021;39:74–74. https://doi.org/10.1200/JCO.2021.39.3_suppl.74.

    Article  Google Scholar 

  456. Wang QJ, et al. Identification of T-cell Receptors Targeting KRAS-Mutated Human Tumors. Cancer Immunol Res. 2016;4:204–14. https://doi.org/10.1158/2326-6066.CIR-15-0188.

    Article  CAS  PubMed  Google Scholar 

  457. Cao B, et al. Development of mesothelin-specific CAR NK-92 cells for the treatment of gastric cancer. Int J Biol Sci. 2021;17:3850–61. https://doi.org/10.7150/ijbs.64630.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  458. Dong X, et al. Efficacy evaluation of chimeric antigen receptor-modified human peritoneal macrophages in the treatment of gastric cancer. Br J Cancer. 2023;129:551–62. https://doi.org/10.1038/s41416-023-02319-6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  459. Zhao G-X, et al. mRNA-based vaccines targeting the t-cell epitope-rich domain of epstein barr virus latent proteins elicit robust anti-tumor immunity in mice. Adv Sci. 2023;10:2302116. https://doi.org/10.1002/advs.202302116.

    Article  CAS  Google Scholar 

  460. Palmer CD, et al. Individualized, heterologous chimpanzee adenovirus and self-amplifying mRNA neoantigen vaccine for advanced metastatic solid tumors: phase 1 trial interim results. Nat Med. 2022;28:1619–29. https://doi.org/10.1038/s41591-022-01937-6.

    Article  CAS  PubMed  Google Scholar 

  461. Rappaport AR, et al. A shared neoantigen vaccine combined with immune checkpoint blockade for advanced metastatic solid tumors: phase 1 trial interim results. Nat Med. 2024. https://doi.org/10.1038/s41591-024-02851-9.

    Article  PubMed  Google Scholar 

  462. Guo Z, et al. Durable complete response to neoantigen-loaded dendritic-cell vaccine following anti-PD-1 therapy in metastatic gastric cancer. NPJ Precis Oncol. 2022;6:34. https://doi.org/10.1038/s41698-022-00279-3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  463. Chen K-S, et al. Bifunctional cancer cell–based vaccine concomitantly drives direct tumor killing and antitumor immunity. Sci Transl Med. 2023;15:eabo4778. https://doi.org/10.1126/scitranslmed.abo4778.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  464. Liang K, et al. Biologically self-assembled tumor cell-derived cancer nanovaccines as an all-in-one platform for cancer immunotherapy. ACS Nano. 2024;18:6702–17. https://doi.org/10.1021/acsnano.4c01050.

    Article  CAS  PubMed  Google Scholar 

  465. Liu X, et al. PD-1+ TIGIT+ CD8+ T cells are associated with pathogenesis and progression of patients with hepatitis B virus-related hepatocellular carcinoma. Cancer Immunol Immunother. 2019;68:2041–54. https://doi.org/10.1007/s00262-019-02426-5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  466. Weimer P, et al. Tissue-specific expression of TIGIT, PD-1, TIM-3, and CD39 by γδ T Cells in Ovarian Cancer. Cells. 2022. https://doi.org/10.3390/cells11060964.

    Article  PubMed  PubMed Central  Google Scholar 

  467. Zhu X, et al. Tumor-associated macrophage-specific CD155 contributes to M2-phenotype transition, immunosuppression, and tumor progression in colorectal cancer. J Immunother Cancer. 2022. https://doi.org/10.1136/jitc-2021-004219.

    Article  PubMed  PubMed Central  Google Scholar 

  468. He W, et al. CD155T/TIGIT signaling regulates CD8+ T-cell metabolism and promotes tumor progression in human gastric cancer. Can Res. 2017;77:6375–88. https://doi.org/10.1158/0008-5472.CAN-17-0381.

    Article  CAS  Google Scholar 

  469. Chiang EY, Mellman I. TIGIT-CD226-PVR axis: advancing immune checkpoint blockade for cancer immunotherapy. J Immunother Cancer. 2022. https://doi.org/10.1136/jitc-2022-004711.

    Article  PubMed  PubMed Central  Google Scholar 

  470. Reches A, et al. Nectin4 is a novel TIGIT ligand which combines checkpoint inhibition and tumor specificity. J Immunother Cancer. 2020. https://doi.org/10.1136/jitc-2019-000266.

    Article  PubMed  PubMed Central  Google Scholar 

  471. Janjigian YY, et al. EDGE-Gastric Arm A1: Phase 2 study of domvanalimab, zimberelimab, and FOLFOX in first-line (1L) advanced gastroesophageal cancer. J Clin Oncol. 2023;41:433248–433248. https://doi.org/10.1200/JCO.2023.41.36_suppl.433248.

    Article  Google Scholar 

  472. Majzner RG, et al. CAR T cells targeting B7–H3, a pan-cancer antigen, demonstrate potent preclinical activity against pediatric solid tumors and brain tumors. Clin Cancer Res Off J Am Assoc Cancer Res. 2019;25:2560–74. https://doi.org/10.1158/1078-0432.CCR-18-0432.

    Article  CAS  Google Scholar 

  473. Crispen PL, et al. Tumor cell and tumor vasculature expression of B7–H3 predict survival in clear cell renal cell carcinoma. Clin Cancer Res Off J Am Assoc Cancer Res. 2008;14:5150–7. https://doi.org/10.1158/1078-0432.CCR-08-0536.

    Article  CAS  Google Scholar 

  474. Seaman S, et al. Eradication of tumors through simultaneous ablation of CD276/B7-H3-positive tumor cells and tumor vasculature. Cancer Cell. 2017. https://doi.org/10.1016/j.ccell.2017.03.005.

    Article  PubMed  PubMed Central  Google Scholar 

  475. Mao Y, et al. Cancer cell-expressed B7–H3 regulates the differentiation of tumor-associated macrophages in human colorectal carcinoma. Oncol Lett. 2017;14:6177–83. https://doi.org/10.3892/ol.2017.6935.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  476. Zang X, et al. Tumor associated endothelial expression of B7–H3 predicts survival in ovarian carcinomas. Mod Pathol. 2010;23:1104–12. https://doi.org/10.1038/modpathol.2010.95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  477. Visvader JE. Cells of origin in cancer. Nature. 2011;469:314–22. https://doi.org/10.1038/nature09781.

    Article  CAS  PubMed  Google Scholar 

  478. Wang R, et al. B7–H3 promotes colorectal cancer angiogenesis through activating the NF-κB pathway to induce VEGFA expression. Cell Death Dis. 2020;11:55. https://doi.org/10.1038/s41419-020-2252-3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  479. Xie C, et al. Soluble B7–H3 promotes the invasion and metastasis of pancreatic carcinoma cells through the TLR4/NF-κB pathway. Sci Rep. 2016;6:27528. https://doi.org/10.1038/srep27528.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  480. Flem-Karlsen K, et al. p38 MAPK activation through B7–H3-mediated DUSP10 repression promotes chemoresistance. Sci Rep. 2019;9:5839. https://doi.org/10.1038/s41598-019-42303-w.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  481. Lim S, et al. Immunoregulatory protein B7–H3 reprograms glucose metabolism in cancer cells by ROS-mediated stabilization of HIF1α. Can Res. 2016;76:2231–42. https://doi.org/10.1158/0008-5472.CAN-15-1538.

    Article  CAS  Google Scholar 

  482. Ingebrigtsen VA, et al. B7–H3 expression in colorectal cancer: nuclear localization strongly predicts poor outcome in colon cancer. Int J Cancer. 2012;131:2528–36. https://doi.org/10.1002/ijc.27566.

    Article  CAS  PubMed  Google Scholar 

  483. Suh W-K, et al. The B7 family member B7–H3 preferentially down-regulates T helper type 1-mediated immune responses. Nat Immunol. 2003;4:899–906.

    Article  CAS  PubMed  Google Scholar 

  484. Prasad DVR, et al. Murine B7–H3 is a negative regulator of T cells. J Immunol. 2004;173:2500–6.

    Article  CAS  PubMed  Google Scholar 

  485. Jin M-Z, Jin W-L. The updated landscape of tumor microenvironment and drug repurposing. Signal Transduct Target Ther. 2020;5:166. https://doi.org/10.1038/s41392-020-00280-x.

    Article  PubMed  PubMed Central  Google Scholar 

  486. Zekri L, et al. An optimized IgG-based B7–H3xCD3 bispecific antibody for treatment of gastrointestinal cancers. Mol Ther J Am Soc Gene Ther. 2023;31:1033–45. https://doi.org/10.1016/j.ymthe.2023.02.010.

    Article  CAS  Google Scholar 

  487. Kelly RJ, et al. Neoadjuvant nivolumab or nivolumab plus LAG-3 inhibitor relatlimab in resectable esophageal/gastroesophageal junction cancer: a phase Ib trial and ctDNA analyses. Nat Med. 2024;30:1023–34. https://doi.org/10.1038/s41591-024-02877-z.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  488. André T, et al. KEYSTEP-008: phase II trial of pembrolizumab-based combination in MSI-H/dMMR metastatic colorectal cancer. Future Oncol. 2023;19:2445–52. https://doi.org/10.2217/fon-2022-1105.

    Article  CAS  PubMed  Google Scholar 

  489. Lanier LL. NKG2D Receptor and Its Ligands in Host Defense. Cancer Immunol Res. 2015;3:575–82. https://doi.org/10.1158/2326-6066.CIR-15-0098.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  490. Gasser S, Orsulic S, Brown EJ, Raulet DH. The DNA damage pathway regulates innate immune system ligands of the NKG2D receptor. Nature. 2005;436:1186–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  491. Le Bert N, et al. STING-dependent cytosolic DNA sensor pathways regulate NKG2D ligand expression. Oncoimmunology. 2014;3: e29259.

    Article  PubMed  PubMed Central  Google Scholar 

  492. Salih HR, Rammensee H-G, Steinle A. Cutting edge: down-regulation of MICA on human tumors by proteolytic shedding. J Immunol. 2002;169:4098–102.

    Article  CAS  PubMed  Google Scholar 

  493. Boutet P, et al. Cutting edge: the metalloproteinase ADAM17/TNF-alpha-converting enzyme regulates proteolytic shedding of the MHC class I-related chain B protein. J Immunol. 2009;182:49–53.

    Article  CAS  PubMed  Google Scholar 

  494. Guerra N, et al. NKG2D-deficient mice are defective in tumor surveillance in models of spontaneous malignancy. Immunity. 2008;28:571–80. https://doi.org/10.1016/j.immuni.2008.02.016.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  495. FerrarideAndrade L, et al. Inhibition of MICA and MICB Shedding Elicits NK-Cell-Mediated Immunity against Tumors Resistant to Cytotoxic T Cells. Cancer Immunol Res. 2020;8:769–80. https://doi.org/10.1158/2326-6066.CIR-19-0483.

    Article  Google Scholar 

  496. Capuano C, et al. Harnessing CD16-mediated NK Cell functions to enhance therapeutic efficacy of tumor-targeting mAbs. Cancers. 2021. https://doi.org/10.3390/cancers13102500.

    Article  PubMed  PubMed Central  Google Scholar 

  497. Courau T, et al. Cocultures of human colorectal tumor spheroids with immune cells reveal the therapeutic potential of MICA/B and NKG2A targeting for cancer treatment. J Immunother Cancer. 2019;7:74. https://doi.org/10.1186/s40425-019-0553-9.

    Article  PubMed  PubMed Central  Google Scholar 

  498. Curtsinger JM, Lins DC, Mescher MF. Signal 3 determines tolerance versus full activation of naive CD8 T cells: dissociating proliferation and development of effector function. J Exp Med. 2003;197:1141–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  499. Curtsinger JM, Mescher MF. Inflammatory cytokines as a third signal for T cell activation. Curr Opin Immunol. 2010;22:333–40. https://doi.org/10.1016/j.coi.2010.02.013.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  500. Razak AR, et al. Safety and efficacy of AMG 820, an anti-colony-stimulating factor 1 receptor antibody, in combination with pembrolizumab in adults with advanced solid tumors. J Immunother Cancer. 2020. https://doi.org/10.1136/jitc-2020-001006.

    Article  PubMed  PubMed Central  Google Scholar 

  501. Hajda J, et al. Phase 2 trial of oncolytic h-1 parvovirus therapy shows safety and signs of immune system activation in patients with metastatic pancreatic ductal adenocarcinoma. Clin Cancer Res Off J Am Assoc Cancer Res. 2021;27:5546–56. https://doi.org/10.1158/1078-0432.CCR-21-1020.

    Article  CAS  Google Scholar 

  502. Eriksson E, et al. Shaping the tumor stroma and sparking immune activation by CD40 and 4–1BB Signaling induced by an armed oncolytic virus. Clin Cancer Res Off J Am Assoc Cancer Res. 2017;23:5846–57. https://doi.org/10.1158/1078-0432.CCR-17-0285.

    Article  CAS  Google Scholar 

  503. Musher BL, et al. LOAd703, an oncolytic virus-based immunostimulatory gene therapy, combined with chemotherapy for unresectable or metastatic pancreatic cancer (LOKON001): results from arm 1 of a non-randomised, single-centre, phase 1/2 study. Lancet Oncol. 2024;25:488–500. https://doi.org/10.1016/S1470-2045(24)00079-2.

    Article  CAS  PubMed  Google Scholar 

  504. Shen Y, et al. VG161 activates systemic antitumor immunity in pancreatic cancer models as a novel oncolytic herpesvirus expressing multiple immunomodulatory transgenes. J Med Virol. 2023;95: e28108. https://doi.org/10.1002/jmv.28108.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

We deeply appreciate Capital's Funds for Health Improvement and Research (2024-1-1021), Beijing Natural Science Foundation (Z20J00105), National Natural Science Foundation of China (82272627) and Beijing Natural Science Foundation (L234003).

Author information

Authors and Affiliations

Authors

Contributions

X. C., Y. M., J. C., and W. L., L. C. contributed equally to this paper and shared first authorship. Y. W. and X. Z. designed the outlines of the review. X. C., Y. M., J. C., W. L., L.C.,J.L,L.M.,H.L.,S.H.,C.Y.,Z.Z.and B.W. collected and summarized researches. X. C., Y. M., J. C., W. L., L.C. drafted the manuscript. Y.W, X.Z Y.C., Z.W., S.L., supervised and revised the manuscript. All the authors have read and approved the final version of the manuscript.

Corresponding authors

Correspondence to Yakun Wang or Xiaotian Zhang.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chong, X., Madeti, Y., Cai, J. et al. Recent developments in immunotherapy for gastrointestinal tract cancers. J Hematol Oncol 17, 65 (2024). https://doi.org/10.1186/s13045-024-01578-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13045-024-01578-x

Keywords