Next Article in Journal
Correction: Lebbink et al. Opposite Incidence Trends for Differentiated and Medullary Thyroid Cancer in Young Dutch Patients over a 30-Year Time Span. Cancers 2021, 13, 5104
Next Article in Special Issue
Novel Dicarboximide BK124.1 Breaks Multidrug Resistance and Shows Anticancer Efficacy in Chronic Myeloid Leukemia Preclinical Models and Patients’ CD34+/CD38 Leukemia Stem Cells
Previous Article in Journal
Volatilome Analysis in Prostate Cancer by Electronic Nose: A Pilot Monocentric Study
Previous Article in Special Issue
New Therapy Options for Neuroendocrine Carcinoma of the Pancreas—The Emergent Substance GP-2250 and Gemcitabine Prove to Be Highly Effective without the Development of Secondary Resistances In Vitro and In Vivo
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Drug Resistance in Colorectal Cancer: From Mechanism to Clinic

1
The Second School of Clinical Medical, Shanxi Medical University, Taiyuan 030001, China
2
Department of General Surgery, Affiliated Drum Tower Hospital of Nanjing University Medical School, Nanjing 210008, China
3
Medical Department of General Surgery, The 1st Medical Center, Chinese PLA General Hospital, Beijing 100853, China
4
Department of General Surgery, The 7th Medical Center, Chinese PLA General Hospital, Beijing 100700, China
5
The Second School of Clinical Medicine, Southern Medical University, Guangzhou 510515, China
*
Authors to whom correspondence should be addressed.
These authors have contributed equally to this work.
Cancers 2022, 14(12), 2928; https://doi.org/10.3390/cancers14122928
Submission received: 25 April 2022 / Revised: 3 June 2022 / Accepted: 7 June 2022 / Published: 14 June 2022
(This article belongs to the Collection Drug Resistance and Novel Therapies in Cancers)

Abstract

:

Simple Summary

Chemotherapy, radiotherapy and molecularly targeted therapy could improve the prognosis of colorectal cancer (CRC) patients. Recently, immunotherapy, especially immune checkpoint inhibitors, has significantly improved the prognosis of some patients. However, drug resistance significantly reduces the usefulness of these drugs. Although research on the molecular mechanisms underlying the emergence of drug resistance is ongoing, the specific molecular mechanisms remain unclear. This article reviews the findings on the mechanisms of drug resistance in CRC patients in preclinical and clinical studies, which may provide valuable directions for future in-depth study of drug resistance.

Abstract

Colorectal cancer (CRC) is one of the leading causes of death worldwide. The 5-year survival rate is 90% for patients with early CRC, 70% for patients with locally advanced CRC, and 15% for patients with metastatic CRC (mCRC). In fact, most CRC patients are at an advanced stage at the time of diagnosis. Although chemotherapy, molecularly targeted therapy and immunotherapy have significantly improved patient survival, some patients are initially insensitive to these drugs or initially sensitive but quickly become insensitive, and the emergence of such primary and secondary drug resistance is a significant clinical challenge. The most direct cause of resistance is the aberrant anti-tumor drug metabolism, transportation or target. With more in-depth research, it is found that cell death pathways, carcinogenic signals, compensation feedback loop signal pathways and tumor immune microenvironment also play essential roles in the drug resistance mechanism. Here, we assess the current major mechanisms of CRC resistance and describe potential therapeutic interventions.

1. Introduction

Colorectal cancer (CRC) is the third most common and second deadliest malignancy in the world [1]. Over the past decade, the development and use of chemotherapy, molecularly targeted therapy and immunotherapy have significantly improved patient outcomes (Figure 1). These anti-cancer drugs are first metabolized into an active form, then transported to the appropriate location and used on the appropriate target, ultimately leading to tumor cell death. The aberration of these processes could lead to the emergence of drug resistance in cancer treatment and significantly reduce its utility. Here, we not only explain the role of the pathway mentioned above that is aberrant in drug resistance, but also highlight the influence of compensatory feedback loop signal path and tumor microenvironment (TME) on CRC resistance, thus providing possible strategies for reversing drug resistance in CRC (Table 1 and Table 2).

2. Drug Metabolism

Only when cancer drugs are transformed into active forms can they maximize their anti-tumor effects. Thus, reduced activity or increased degradation of anti-cancer drugs would lead to drug resistance (Figure 2).

2.1. Reduced Drug Activation

Capecitabine (CAP) needs to be transformed into 5-fluorouracil (5-FU) for direct cytotoxicity, which is mediated by thymine phosphorylase (TP) [2]. Preclinical studies show that methylation of extracellular growth factor-1 (ECGF-1 gene, encoding TP protein) inhibits TP gene transcription and leads to drug resistance to CAP, while DNA methyltransferase (DNMT) inhibitors can reverse drug resistance to CAP [3].
5-FU exerts its anti-cancer effects by inhibiting thymidylate synthase (TS) and incorporating its metabolites into RNA and DNA [4]. Orotate phosphoribosyl-transferase (OPRT), uridine monophosphate (UMP) synthetase (UMPS) and UMP kinase (UMPK) are responsible for the conversion of 5-FU into active anti-cancer metabolites in tumor cells [5]. It has been proven that the lower expression of these metabolic enzymes is associated with 5-FU resistance in CRC patients [6,7,8]. In addition, TP can convert 5-FU into 5-fluoro-2′-deoxyuridine (FdU), which is a form that converts 5-FU into an active form. Studies have shown that the higher expression of TP predicts a good response to 5-FU [9,10]. Although another study has shown that there is no correlation between the level of TP expression and tumor response rate, the progression-free survival of patients with higher expression of TP is significantly prolonged [11]. These data suggest that patients with high expression of these enzymes that convert 5-FU to the active form may be more likely to benefit from 5-FU.
Table 1. Promising strategies for reversing drug resistance in preclinical studies.
Table 1. Promising strategies for reversing drug resistance in preclinical studies.
TherapyTreatment RegimenTargetSubpopulationModel SystemReference
ABC inhibitorsCryptotanshinone/dihydrotanshinone/cinobufagin/8-oxocoptisineP-gp/Cell[12,13,14]
Liposomal antisense oligonucleotidesP-gp, MRP1, MRP2, Bcl-2, Bcl-XL/Cell[15,16]
Y-box binding protein 1 (YB-1)TOP-1/Cell[17].
New anti-EGFR mAbsSym004EGFR ECDEGFR G465E mutantCell/mice[18]
MM-151EGFR ECDEGFR G465E mutantPDX[19]
NecitumumabEGFR ECDEGFR S468R mutantCell[20]
MEK inhibitorsBAY86-9766 + cetuximabMEKKRAS mutantPDX[21]
Pimasertib + cetuximabMEKKRAS, NRAS or BRAF mutantCell/mice/PDX[22]
Trametinib + Olaparib + αPD-L1MEKKRAS mutantMice[23]
G-38963 + αPD-1MEKKRAS mutantMice[24]
New anti-MEK mAbs SCH772984MEKKRAS or BRAF mutantPDX[25]
IGF-IR inhibitorsNVP-AEW541(IGF-IR i) + aEGFRIGF-IR/CRC primary cell line[26]
HER2 mABsPertuzumab + cetuximab/lapatinibHER2HER2 amplificationPDX[27]
Trastuzumab + cetuximabHER2/Cell[28]
4D5 + cetuximabHER2/PDX[29]
HER3 mABsU3-1402HER3High HER3 expression;CDX[30]
β-catenin/CBP inhibitorsTrpC5β-catenin/CBP/Cell[31]
AQP5β-catenin/CBP/Cell[32]
ICG-001β-catenin/CBP/Cell/mice[32]
TLR9 agonistSD-101 + PD-1TLR-9/Mice[33]
IMO + cetuximab/bevacizumabTLR-9/Mice[34]
IMO + cetuximab + irinotecanTLR-9/Cell and mice[35]
IMO + cetuximabTLR-9/Cell and mice[36]
TGF-β inhibitorsLY2157299 + GefitinibTGF-βIR/Cell[37]
PF-03446962 + BevacizumabALK-1/Mice and PDX[38]
Galunisertib +αPD-L1TGFBR1low mutational burden, T-cell exclusion and TGFβ-activated stromaMice[39]
Apoptotic inhibitorsABT-263 + selumetinibBax, BakBRAF(V600E) or RAS mutantCell[40]
AEG35156 + taxanesXIAP/Mice[41]
Birinapant + oxaliplatin/5-FUXIAP/Cell[42]
Autophagy inhibitorsChloroquine/3-MA + 5-FUAutophagy/Cell[43,44]
Chloroquine/3-MA + cetuximabAutophagy/Cell[45]
3-MA + αEGFRAutophagy/Cell[46]
SB02024/SAR405 + αPD-1/PD-L1Vps34/Mice[47]
Ferroptosis agonistβ-elemene + cetuximab/KRAS mutantCell and mice[48]
RSL3 + cisplatin//Cell and mice[49]
CAFBGJ398 + 5-FU + oxaliplatinFGFR4/Cell[50]
BLU9931 + cetuximabFGFR4/Cell and mice[51]
Regorafenib + cetuximabFGFR, VEGF, PDGFR-β/Cell and mice[52]
GKT137831NOX4/Mice[53]
MDSCKTN0158 + αPD-1 + αCTLA-4KIT/Mice[54]
TAMDFMO + 5-FUODC/Mice[55]
RG7155CSF1R/Mice[56]
MDSCR848 + oxaliplatinTLR 7/8 agonist/Mice[57]
MDSC and TAMTP-16 + αPD-1EP4/Mice[58]
VEGF inhibitorsαVEGF(A) + αPD-1VEGF-A/Mice[59,60,61]
Vargatef + afatinibVEGF, EGFR/Mice[62]
Regorafenib + cetuximabFGFR, VEGF, PDGFR-βKRAS or BRAF mutantCell and mice[52]
Table 2. Promising strategies for reversing drug resistance in clinical trials.
Table 2. Promising strategies for reversing drug resistance in clinical trials.
TherapyTreatment RegimenTargetSubpopulationSpeciesSettingEfficiencyReference
New anti-EGFR mAbsSym004EGFR ECD/anti-EGFR-refractory/unselected mCRCPhase I trial studiesORCD: (67%)NCT01117428 [18]
Sym004EGFR ECDKRAS WTanti-EGFR-refractory mCRCPhase II trial studiesmOS: (10.3 m:9.6 m)2013-003829-29 [63]
MM-151 + irinotecan vs. MM-151EGFR ECDKRAS WTRefractory advanced CRCPhase I trial studiesMM-151-SD: 31%
MM-151+ irinotecan- PR: 1/3
NCT01520389 [64]
NecitumumabEGFR ECD/first-line treatment for locally advanced or metastatic CRCPhase II trial studiesORR: (63.6%)
OS: (22.5 m)
[65]
BRAF inhibitorVemurafenib + panitumumabBRAF V600EBRAF-mutantRefractory mCRCPilot trialORR: (16.7%)[66]
Vemurafenib + cetuximab vs. vemurafenibBRAF V600EBRAF V600E mutantCRCPhase II trial studiesORR: (0%:4%)NCT01524978 [66]
Vemurafenib + irinotecan + cetuximabBRAF V600EBRAF V600E mutantmCRCPhase Ib trial studiesORR: (35%)NCT01787500 [67].
PI3K inhibitorcetuximab+PX-866 vs. cetuximabPI3KKRAS WTirinotecan- and oxaliplatin-refractory mCRCPhase II trial studiesmOS: (266d:333d)[68]
Alpelisib + cetuximab + encorafenib + vs. cetuximab + encorafenibPI3KBRAF-MutantmCRCPhase Ib trial studiesmOS: (4.2 m:3.7 m)NCT0171938 [69]
IGF-IR inhibitorIMC-A12 + cetuximab vs. IMC-A12IGF-IR/anti-EGFR-refractory mCRCPhase II trial studiesORR: (0%:5%)NCT00503685 [70]
Dalotuzumab + Cetuximab + Irinotecan vs. Cetuximab + IrinotecanIGF-IRKRAS WTchemo-refractory mCRCPhase II/III trial studiesmOS: (10.8 m:14 m)NCT00614393 [71]
HGF/IGF-IR inhibitorRilotumumab + panitumumab vs. ganitumab + panitumumab vs. panitumumabHGF/IGF-IRKRAS WTpreviously treated mCRCPhase Ib/II trial studiesmOS: (13.8 m:10.6 m:11.6 m)NCT00788957 [72]
HER2 inhibitorNeratinib + CetuximabHER2Quadruple-WT(KRAS, NRAS, BRAF, PIK3CA)anti-EGFR-refractory mCRCPhase Ib trial studiesThere were no objective responsesNCT01960023 [73]
Trastuzumab + lapatinibHER2KRAS WT, HER2-positiveanti-EGFR-refractory mCRCPhase II trial studiesORR: (30%)[74]
Trastuzumab + PertuzumabHER2HER2 amplificationtreatment-refractory mCRCPhase IIa trial studiesORR: (32%)NCT02091141 [75]
HER3 inhibitorLumretuzumabHER3HER3-positivetreatment-refractory mCRCPhase I trial studiesSD: (21.3%)NCT01482377 [76]
MEHD7945AHER3/EGFR/treatment-refractory mCRCPhase I trial studiesSD: (28.6%)NCT01207323 [77]
MET inhibitorTivantinib + CetuximabMETKRAS WTanti-EGFR-refractory mCRCPhase II trial studiesORR: (9.8%)NCT01892527 [78]
Tivantinib + irinotecan + cetuximab vs. irinotecan + cetuximabMETKRAS WTlocally advanced or metastatic CRCPhase I/II trial studiesmPFS: (8.3 m:7.3 m)
mOS: (19.8 m:16.9 m)
ORR: 45%:33%
NCT01075048 [79]
Capmatinib + cetuximabMETK/NRAS WT, MET-positiveanti-EGFR-refractory mCRCPhase Ib trial studiesSD: (46.2%)NCT02205398 [80]
VEGF/ALK-1 inhibitorRegorafenib + PF-03446962VEGF/ALK-1/treatment-refractory mCRCPhase Ib trial studiesSD: (18.2%)[81]
MEK inhibitorCobimetinib + atezolizumabMEK/mCRCPhase I/Ib trial studiesORR: (8%)NCT01988896 [82]
Atezolizumab + cobimetinib vs. Atezolizumab vs. regorafenibMEK/previously treated mCRCPhase III trial studiesmPFS: (1.91 m:1.94 m:2.0 m)
mOS: (8.87 m:7.10 m:8.51 m)
NCT02788279 [83]
Cobimetinib + Atezolizumab + HydroxychloroquineMEKKRAS-MutantAdvanced MalignanciesPhase I/II trial studiesNoneNCT04214418
CSF1R inhibitorPexidartinib + DurvalumabCSF1R/Metastatic/Advanced CRCPhase I trial studiesNoneNCT02777710
ARRY-382 + PembrolizumabCSF1R/Advanced Solid TumorsPhase I/II trial studiesNoneNCT02880371
anti-VEGF mAbsBevacizumab + AtezolizumabVEGF-AMSI-HmCRCPhase Ib trial studiesORR: (30%)
SD: (90%)
NCT01633970 [84]
bevacizumab + capecitabine + atezolizumab vs. bevacizumab + capecitabineVEGF-A/refractory mCRCPhase II trial studiesmPFS: (3.3 m:4.4 m)
ORR: (4.35%:8.45%)
12 m OS: (43%:52%)
NCT01633970 [85]
nivolumab + standard of care chemotherapy + bevacizumabVEGF-A/mCRCPhase II/III trial studiesNoneNCT03414983
anti-VEGFR mAbsRegorafenib + cetuximab/PanitumumabVEGFR/Unresectable, Locally Advanced, or Metastatic Colorectal CancerPhase I trial studiesClinical Benefit: (53%)NCT02095054 [86]
Phase II trial studiesNoneNCT04117945
Irinotecan exerts a cytotoxic effect by inhibiting DNA replication and transcription by targeting topoisomerase 1 (TOP-1) [87]. Carboxylesterase 2 (CES2) is the critical enzyme for the activation of irinotecan, which can catalyze the conversion of irinotecan to the more active form of SN-38 (a topoisomerase inhibitor) [88]. Multiple studies have shown that the expression of CES2 in CRC is positively correlated with irinotecan sensitivity [89,90]. Multiple preclinical studies have shown that selectively increasing the activity of CES in CRC by viruses can improve the efficacy of irinotecan [91,92,93]. Although some studies can indeed enhance the sensitivity of irinotecan in vitro and in vivo, this mechanism has not been studied in a clinical setting. In addition, it has been reported that genetically engineered human neural stem cells can increase CES activity in other tumors [94,95,96]. Gene therapy may be a future approach to increase the sensitivity of CRC to irinotecan chemotherapy.

2.2. Increased Drug Inactivation

The vast majority of 5-FU is catabolized into inactive metabolites by dihydropyrimidine dehydrogenase (DPD) in the liver [97]. As the name implies, high expression of DPD mRNA can confer resistance to 5-FU in CRC [9,10]. Therefore, interference with DPD is a way to overcome drug resistance in these patients. Preclinical studies have shown that methylation of the DPYD gene promoter, which encodes DPD protein, can inhibit DPYD transcription [98]. In addition, methylation of the DPYD promoter region can reduce the activity of DPD in clinical samples [99]. In addition, studies have reported that DPD inhibitors such as 5-ethynyluracil and 5-chloro-2,4-dihydroxypyridine (CDHP) can improve the anti-tumor activity of 5-FU to a greater extent [100,101,102]. However, patients with DPD deficiency will experience severe systemic poisoning symptoms when using 5-FU [103]. Therefore, the expression level of DPD should be further standardized to maximize the benefit of patients.
Irinotecan or its activated form SN38 can be metabolically inactivated by uridine diphosphate glucuronosyltransferase 1A1 (UGT1A1), cytochrome P450-3A4 (CYP3A4) and β-glucuronidase [104]. These enzymes involved in irinotecan metabolism are susceptible to environmental factors such as smoking, herbal supplements and drug therapy that alter the irinotecan metabolic pathway, resulting in altered irinotecan clearance [105,106,107]. Therefore, targeting these enzymes may be an effective way to reverse irinotecan resistance. Methimazole (CYP3A inhibitor) [108], neomycin (β-glucuronidase inhibitor) [109] and amoxapine (β-glucuronidase inhibitor) [110] have been shown to reduce SN-38 degradation effectively.
In addition, the transcription level of UGT1A1 is negatively regulated by DNA methylation [111]. Therefore, UGT1A1 promoter methylation and subsequent inhibition of UGT1A1-related metabolic pathways participate in the retention of active SN-38 in CRC, enhancing the therapeutic effect of irinotecan [112].

3. Drug Transport

When the membrane transporter pumps chemotherapeutic drugs or molecularly targeted drugs from the inside of the cell out of the cell, it would decrease the concentration of anti-tumor drugs in the tumor cell and the development of multi-drug resistance (MDR) [113]. To date, the human genome has been proven to have more than 400 membrane transporters, which are divided into two families: ATP binding cassette (ABC) and solute carrier (SLC) transporters [114]. Representative ABC transporters are multi-drug resistance protein 1 (MDR1; P-gp; ABCB1), breast cancer resistance protein (BCRP; ABCG2) and multi-drug-resistance-associated protein 1 (MRP1, ABCC1), while solute carrier (SLC) transporters include organic anion transporters, organic cation transporters and organic anion transport polypeptides [115]. In fact, ABC transporter is the main membrane transporter for MDR and plays an important role in CRC resistance to chemotherapy (Figure 2).
P-glycoprotein (P-gp) is the first identified ABC superfamily member and exists in various normal cells of the intestine [116]. P-gp is the most classical efflux pump mediating multi-drug resistance in tumor therapy; the ATP-driven hydrolysis of P-gp may pump chemotherapeutic drugs out of the cell. Preclinical and clinical studies have demonstrated that overexpression of P-gp in a variety of solid tumors, including CRC, leads to multi-drug resistance and chemotherapy failure. For example, P-gp is overexpressed in CRC cell lines and specimens resistant to 5-FU/Dox/oxaliplatin [31,117,118]. The overexpression of BCRP and MRP1 has subsequently been identified to be associated with drug resistance in various tumors, including CRC [119,120,121].
Given the vital position of ABC transporters in CRC resistance, inhibiting these transporters is an effective way to reverse resistance. Therefore, ABC transporter inhibitors have emerged. Since the emergence of the first-generation P-gp inhibitor verapamil in 1981, P-gp inhibitors have been studied in depth, and the third-generation P-gp inhibitors have appeared today [122]. Although the second- and third-generation P-gp inhibitors have higher affinity and lower side effects, their therapeutic effects have not been improved as expected in preclinical and clinical studies. Therefore, it will be exciting to develop an effective, high-affinity and relatively non-toxic natural substance as an inhibitor. P-gp modulators derived from natural products belong to the fourth generation of P-gp inhibitors [123]. A variety of studies have reported natural products as P-gp inhibitors. For example, cryptotanshinone and dihydrotanshinone, the extracts of danshen, can reverse the sensitivity to doxorubicin and irinotecan by downregulating the expression of P-gp mRNA and protein in CRC cell lines [12]. Natural products cinobufagin [13] and 8-oxocoptisine [14] have also been shown to reverse chemotherapy resistance for CRC by inhibiting P-gp. This evidence indicates that natural products may be a target for solving chemotherapy resistance in the future. In addition to P-gp inhibitors, other modulators of ABC transporters have achieved good results. Most of the experimental data about ABC transporter inhibitors are obtained based on cell experiments, and the specific effects in vivo have not been well-proven. Therefore, the main work included preclinical or clinical research based on pharmacology.
Another way to inhibit the ABC transporter is to reduce the expression level of the transporter. Studies have shown that antisense oligonucleotides (AOSs), miRNA and LncRNA can mediate the expression of ABC transporters to reverse drug resistance. For example, liposome AOSs reverse multi-drug resistance in vivo and in vitro by inhibiting the expression of P-GP, MRP1, MRP2 and Bcl-2/Bcl-XL, which may enhance the sensitivity of CRC to chemotherapy by decreasing the activity level of the MDR1 promoter [15,16]. MiR-506 and miR-26b are downregulated in chemotherapy-resistant CRC cell lines and specimens. Overexpression of miR-506 and miR-26b can reverse drug resistance of CRC cells to oxaliplatin and 5-FU by downregulating P-gp [117,124]. Recent studies have shown that LncRNA CASC15 is highly expressed in oxaliplatin-resistant CRC cell lines and specimens. LncRNA CASC15 can enhance the expression of ABCC1 by inhibiting the expression of miR-145 from conferring oxaliplatin resistance to CRC in vivo and in vitro [125]. LncRNA PVT1 was overexpressed in CRC cells and specimens with 5-FU resistance, which might be attributed to the fact that PVT1 increased CRC resistance to 5-FU by inhibiting apoptosis and upregulating the expression of MRP1, P-GP and Bcl-2 mRNA and protein. Silencing LncRNA PVT1 increased apoptosis and downregulated ABC transporter, thus increasing the sensitivity of CRC to 5-FU [126]. Therefore, it is necessary to develop strategies based on antisense oligonucleotides, miRNA and LncRNA to reduce the expression level of ABC transporters to reverse the drug resistance of CRC cells.

4. Changes of Drug Targets

Aberration of corresponding drug targets and/or ligands can mediate the generation of drug resistance, no matter whether for chemotherapy drugs or molecular targeted therapy drugs (Figure 2).

4.1. Thymidylate Synthase (TS)

Changes in target mutations or expression levels of chemotherapeutic drugs can influence drug responses and subsequently induce drug resistance. 5-FU mainly destroys the deoxynucleotide pool required for DNA replication by inhibiting thymidylate synthase (TS), and mixes its metabolites into RNA and DNA to exert anti-tumor effects [4,127]. Several studies have shown that the high expression of TS is a crucial determinant of CRC resistance to 5-FU, and 5-FU intervention can also increase the expression of TS [4,9,10]. Under normal circumstances, unbound TS can be combined with its own mRNA, thus inhibiting its translation through the activity of decreased negative feedback, but when TS with 5-FU metabolites FdU experiences steady binding, TS can no longer bind to its own mRNA, resulting in increased protein expression because the sharp increase in TS will promote the recovery of enzyme activity, which constitutes a potential resistance mechanism [128]. Therefore, developed methods to inhibit this adverse feedback pathway will be a new way to combat resistance. In addition, the addition of folic acid sources such as leucovorin to exogenous therapy enhances the cytotoxicity of 5-FU by increasing TS inhibition in vitro and in vivo [129].

4.2. Topoisomerase 1 (TOP-1)

Given the targeting properties of irinotecan or its active metabolite SN-38, mutations or lower expression of TOP-1 may cause irinotecan resistance to CRC [89]. The rearrangement and hypermethylation of TOP-1 genes can be found in drug-resistant cells, leading to transcriptional silencing [130]. In addition, the amount of TOP-1 bound to DNA is reduced, which may also lead to CRC resistance to irinotecan [88,131]. Therefore, the increased expression of TOP-1 may increase irinotecan sensitivity. Y-box binding protein 1 (YB-1), as a promoter of intracellular TOP-1 catalytic activity, can improve the sensitivity of irinotecan by directly interacting with TOP-1 [17].

4.3. Epidermal Growth Factor Receptor and Its Ligand

Cetuximab specifically binds epidermal growth factor receptor (EGFR) and blocks the EGFR signaling pathway, thereby exerting anti-tumor activity [132]. However, multiple ligands of EGFR can activate EGFR, such as epidermal growth factor (EGF), transforming growth factor alpha (TGF-α), heparin-binding-EGF (HB-EGF), betacellulin (BTC), amphiregulin (AREG) and epiregulin (EREG) (Figure 2) [133,134]. Several clinical studies have confirmed that the expression levels of AREG and EREG are related to the treatment response of cetuximab [135,136]. However, these ligands have no predictive power in KRAS-mutations patients [136]. Similar results were obtained in a subsequent randomized clinical trial of panitumumab, irinotecan and cyclosporine to treat colorectal cancer [137]. This evidence suggests that upregulation of AREG and EREG may be one of the mechanisms of anti-EGFR mAbs resistance in KRAS-WT (wild-type) CRC patients.
Mutations or low expression of EGFR are detrimental to the efficacy of anti-EGFR therapy in lung cancer [138]. However, clinical studies have shown that metastatic colorectal cancer (mCRC) patients respond to anti-EGFR mAbs regardless of EGFR status [139,140]. A growing number of studies suggest that mutations in the extracellular domain (ECD) of the EGFR, such as S492R, G465E, R451C and K467T, prevent binding cetuximab and make CRC resistant to anti-EGFR mAbs [141,142,143,144]. In addition, methylation at R198 and R200 of the EGFR ECD plays an essential role in regulating EGFR functionality and resistance to cetuximab treatment in CRC patients [145].
Therefore, the development of new mAbs that can bind to different or mutated EGFR ECDs is expected to improve the efficiency of anti-EGFR mAbs. MM-151 is a drug consisting of three anti-EGFR mAbs. Preclinical studies have shown that CRC with resistance to cetuximab or panitumumab due to EGFR ECD mutations is sensitive to EGFR blockade by MM-151 [19]. A Phase I clinical trial reported that MM-151 has an acceptable tolerability profile and objective clinical activity alone and in combination with irinotecan [64]. Necitumumab is another FDA-approved anti-EGFR mAbs. A preclinical study has shown that necitumumab can bind and inhibit the most common mutated form of EGFR (S468R) that is resistant to cetuximab and other EGFR epitope variants that are resistant to cetuximab [20]. A Phase II trial study evaluated necitumumab plus modified FOLFOX6 as a first-line treatment for locally advanced or metastatic CRC [65]. Compared with unselected mCRC patients who received mFOLFOX6 alone in a randomized Phase III TREE study, the addition of necitumumab significantly improved objective response rate (ORR, 63.6% vs. 41%) and overall survival (OS, 22.5 m vs. 19.2 m) [65,146]. Sym004 is a 1:1 mixture of two recombinant human-mouse chimeric mAbs directed against non-overlapping EGFR epitopes. Preclinical studies have shown that Sym004 could effectively inhibit cells stimulated by EGFR ligands and overcome cetuximab resistance mediated by EGFR ECD mutations in CRC [18,147]. However, in a Phase II clinical study, Sym004 effectively targeted cancer cells with EGFR ECD mutations, but did not improve the OS of mCRC [63]. Nevertheless, EGFR-related testing in mCRC is not currently recommended [148].

5. Aberration in Downstream Signaling Pathways

The EGFR signaling pathway has two downstream pathways: RAS/RAF/MEK/ERK and PI3K/AKT/mTOR. Anti-EGFR mAbs, such as cetuximab, inhibit these downstream pathways by inhibiting EGFR, so aberrant activation of these downstream pathways can mediate anti-EGFR mAbs resistance by bypassing EGFR (Figure 2).

5.1. RAS/RAF/MEK/ERK

RAS mutations are well-known predictors of resistance to anti-EGFR mAbs [135]. Some large clinical trials have shown that FOLFIRI (leucovorin + 5-FU + irinotecan) or FOLFOX regimens combined with EGFR inhibitors can benefit KRAS-WT CRC. However, CRC with KRAS mutations do not benefit from EGFR inhibitors, possibly because activation of KRAS does not depend upstream on EGFR [149,150,151,152]. KRAS mutations are the most common RAS mutations, occurring in 40% of mCRC [149]. Mutations in exons 2–4 of KRAS and NRAS can lead to mCRC resistance to cetuximab and panitumumab [153,154]. However, there are still no drugs that specifically target RAS. Some preclinical and clinical studies have reported that some drugs could reverse resistance induced by RAS, such as sotorasib, dasatinib and metformin [155,156,157].
BRAF mutations are another critical factor in the poor prognosis of mCRC patients treated with anti-EGFR mAbs [149,152]. The hotspot of BRAF mutations in CRC is the substitution from valine to glutamic acid at codon 600 (V600E), located in exon 15, leading to 130- to 700-fold increased BRAF kinase activity compared with that of BRAF-WT [158]. The prevalence of BRAF V600E mutations in mCRC is 8–10%, and they occur only in KRAS-WT mCRC [159,160]. Vemurafenib is an oral inhibitor that binds explicitly to BRAF V600 kinase. A pilot trial in patients with BRAF-mutations mCRC after chemotherapy showed that the combination of vemurafenib and panitumumab limited tumor progression and produced moderate clinical activity [66]. However, in another clinical study, CRC patients did not benefit from vemurafenib monotherapy and/or cetuximab [161]. In a Phase Ib trial study, vemurafenib combined with irinotecan and cetuximab had favorable survival outcomes and response rates in patients with refractory BRAF-mutant mCRC [67]. Encorafenib, another BRAF inhibitor, has been shown to reverse anti-EGFR resistance by targeting both EGFR and BRAF. In the recently published BEACON CRC study, encorafenib plus cetuximab improved OS, ORR and progression-free survival in 444 patients with previously treated BRAF V600E-mutant mCRC compared with standard chemotherapy [162].
MEK/ERK is a downstream cascade of RAS/RAF. We should emphasize that RAS/RAF mutations ultimately lead to cell proliferation and survival by activating MEK. It is exciting that multiple preclinical studies have demonstrated that MEK inhibitors can inhibit the proliferation of CRC cells with KRAS or BRAF mutations in vitro and in vivo, reversing drug resistance mediated by KRAS or BRAF mutations [21,22]. However, KRAS or BRAF mutant CRC usually develops resistance to MEK inhibitors [25,163]. Although the mechanism of this resistance is unclear, preclinical studies suggest that it may be related to the reactivation of the ERK signaling pathway [25]. SCH772984, a novel ERK inhibitor, effectively inhibited MAPK signaling and cell proliferation in BRAF or MEK inhibitor resistance models and reversed the sensitivity of including CRC to BRAF and MEK inhibitors [25].

5.2. PI3K/AKT/mTOR

PI3K/AKT/mTOR is another downstream parallel pathway of EGFR. There is increasing evidence that aberrant activation of this pathway is associated with anti-EGFR treatment resistance [164]. In a clinical study of 110 mCRC patients treated with anti-EGFR mAbs, the authors found that PIK3CA mutations were significantly associated with clinical resistance to panitumumab or cetuximab, particularly in KRAS wild-type mCRC [165]. PIK3CA mutations occurring in the “hotspots” located in exon 9 (E542K, E545K) and exon 20 (H1047R) are oncogenic in CRC cellular models [166]. However, PIK3CA exon 20 mutations significantly reduced the anti-EGFR response rate in the case of KRAS-WT, while exon 9 mutations had no significant effect [167,168]. In addition, tumor suppressor gene phosphatase and tensin homolog (PTEN) is a negative regulator of this pathway and aberrant PTEN may lead to anti-EGFR resistance [169]. Preclinical and clinical data show that loss of PTEN protein significantly reduces the sensitivity of CRC to cetuximab [170,171].
In summary, these changes confer CRC resistance to anti-EGFR mAbs. Whereas PI3K inhibitors and mTOR inhibitors have been developed and approved for clinical trials in PIK3CA-mutated solid tumors [164]. However, two clinical trials showed that adding PI3K inhibitors to cetuximab did not improve PFS, ORR and OS in mCRC patients [68,69]. More and more clinical trials are needed to verify its safety and efficacy.

6. Aberrant Activations of Alternative Receptors

Insulin-like growth factor 1 (IGF1)/IGF2, HGF/MET and HER2/HER3 have been reported to activate the PI3K/AKT and RAS/RAF/MEK/ERK axis independently of EGFR [172,173]. Alterations in these pathways play an essential role in primary and secondary resistance to anti-EGFR mAbs (Figure 2).

6.1. IGF1 and IGF2

IGF1 and IGF2 have substantial effects on cell proliferation and differentiation and are effective inhibitors of apoptosis [174]. The role of IGF1 and IGF2 is mediated mainly by the IGF1 receptor (IGF-1R). Multiple clinical studies have demonstrated that IGF1-negative patients with CRC in KRAS-WT treated with irinotecan-cetuximab have higher responses, higher median-PFS and higher median-OS than IGF1-positive patients [172,175]. In addition, preclinical studies have shown that IGF2 overexpression attenuates the efficacy of cetuximab in CRC patient-derived xenografts (PDXs) [176].
Targeting IGF1/IGF2 may be an effective way to overcome anti-EGFR resistance. In preclinical studies, co-targeting EGFR and IGF-1R can synergistically enhance the anti-tumor effect of CRC cells [26]. Unfortunately, IGF-1R inhibitor IMC-A12 [70], dalotuzumab (MK-0646) [71] and ganitumab [72] did not induce effective ORR and improved median-OS in mCRC. This is inconsistent with the theoretical efficacy of IGF-1R inhibitors, and more clinical evidence and new targeted drugs are needed to demonstrate the possibility of IGF-1R as a therapeutic target after anti-EGFR mAbs resistance.

6.2. HER2 and HER3

The cancer genome atlas (TCGA) colorectal cancer project found that 7% of CRC patients have HER2 somatic mutations or HER2 gene amplification [177]. A series of studies have demonstrated that HER2 gene amplification or activating mutations can mediate resistance to anti-EGFR mAbs in CRC [27,178,179]. It may be that HER2 gene amplification or activating mutations mediate aberrant activation of HER2, leading to persistent ERK1/2 signaling mediated resistance to cetuximab [178,179].
Therefore, targeting HER2 may be a strategy to reverse CRC resistance to EGFR. Current HER2-targeted drugs, such as trastuzumab, pertuzumab and lapatinib, have significantly improved HER2-amplified breast cancer and CRC. Preclinical studies have shown that the combination of HER2 and EGFR blockade can lead to CRC tumor regression [27,28,29]. In addition, in CRC, the HER2-mutated PDXs model showed that trastuzumab plus neratinib caused sustained tumor regression compared to a single HER2-targeted agent [179]. In two large clinical studies, trastuzumab and lapatinib/pertuzumab combinations showed good activity and tolerability in EGFR-resistant HER2-positive mCRC [74,75]. However, data from a recently published Phase Ib clinical study suggest that neratinib-plus-cetuximab did not show an objective response in quadruple-WT (KRAS, NRAS, BRAF and PIK3CA) mCRC resistant to cetuximab or panitumumab [73].
HER3 mutations are present in 5.7% to 11% of CRC [180,181]. Extensive data suggest that aberrant HER3 in mCRC may also reduce the efficacy of anti-EGFR therapy [172,182]. Currently, HER3-targeted drugs U3-1402 [30], lumretuzumab [76] and MEHD7945A [77] have been used in preclinical and clinical trials with good safety and efficacy. Further clinical studies are needed to demonstrate the clinical value of HER2/HER3 combined with anti-EGFR mAbs.

6.3. HGF/MET

The proto-oncogene MET, which enforces hepatocyte growth factor receptor (HGFR, also known as MET tyrosine kinase receptor), shares several downstream pathways with EGFR, including RAS/RAF/MAPK and PI3K/AKT signaling pathways [173,183]. A clinical study has demonstrated that MET amplification is associated with acquired resistance to cetuximab or panitumumab in patients with KRAS-WT mCRC [184]. In addition, anti-EGFR therapy leads to MET amplification, contributing to the development of CRC-acquired resistance [185,186]. The possible reason is that EGFR ligand TGF-α can induce cross-interaction between EGFR and MET, which increases the phosphorylation of MET and its downstream MAPK and AKT, leading to the generation of drug resistance [187].
Therefore, targeting HGF-MET may be a way to overcome resistance to anti-EGFR therapy. Tivantinib is a selective oral MET inhibitor that has been used in monotherapy and combination therapy for solid tumors, including mCRC [188]. Data from a Phase II clinical study showed that tivantinib combined with cetuximab showed good toleration and activity in 41 mCRC patients with MET-high who relapsed after cetuximab or panitumumab treatment [78]. However, in another clinical study, the combination of tivantinib, cetuximab and irinotecan was well-tolerated in patients with KRAS-WT mCRC, but did not significantly improve PFS [79]. In addition, rilotumumab (AMG 102) is an all-human HGF-targeting IgG 2 mAbs that neutralizes HGF-dependent MET signaling. In a randomized Phase Ib/II trial, the combination of panitumumab and rilotumumab significantly increased the anti-tumor activity of patients compared with panitumumab alone [72].

7. Persistent Activation of Oncogenic/Bypass Signaling

7.1. Wnt/β-Catenin

As an evolutionarily conserved signal, Wnt/β-catenin signaling is involved in various processes of embryonic development and also plays a crucial role in cancer [189]. Sustained activation of Wnt/β-catenin signaling endows cancer cells with sustained self-renewal growth and is associated with tumor treatment resistance [190].
Current studies have shown that activation of the Wnt/β-catenin signaling pathway can mediate CRC resistance to chemotherapy through several pathways, such as upregulation of MDR1, inhibition of apoptosis, maintenance of stem cell dryness, regulation of epithelial-mesenchymal transformation (EMT) and regulation of tumor microenvironment [191]. In addition, Wnt and EGFR signals are closely related. For example, EGFR can form a complex with β-catenin to activate the Wnt pathway, and the Wnt ligand can confer anti-EGFR resistance through its Frizzled activation of EGFR signaling [192,193]. The overexpression of MiR100HG and miR-100/125b was also observed in cetuximab-resistant CRC cell lines and tumors from CRC patients that progressed on cetuximab [194]. Specifically, miR-100/125b coordinately represses five Wnt/β-catenin negative regulators, resulting in increased Wnt signaling, and inhibition of Wnt restored cetuximab reactivity in cetuximab-resistant cells [194].
Immune checkpoint inhibitors (ICIs) such as nivolumab and pembrolizumab could reverse T cell dysfunction and apoptosis by inhibiting programmed death-1 (PD-1), which enhances T cell activation and cytotoxicity to tumor cells [195]. The Wnt/β-catenin signaling pathway is also associated with ICIs resistance. In primary CRC tumor samples, Wnt signaling reduced the frequency of tumor-infiltrating T cells, regardless of mutation load [196]. In addition, APC, as a gene that upregulates WNT signaling, is negatively correlated with the frequency of T cell infiltration [196,197,198]. It may be that the loss of APC upregulates the expression of Dickkopf-related protein 2 (DKK2), which, together with its receptor LRP5, inhibits STAT5 signaling, leading to immunosuppression. Induction of DKK2 ablation activates natural killer cells (NK) and CD8+ T cells in tumors and enhances the efficacy of anti-PD-1 therapy [198].
Inhibition of the Wnt pathway may reverse CRC sensitivity to treatment. For example, inhibiting the expression of transient receptor potential channel 5 (TRPC5) can weaken the ABCB1 efferent pump by inhibiting the Wnt/β-catenin signaling pathway, thus reversing 5-FU resistance in CRC [31]. The silencing of aquaporin 5 (AQP5) enhances CRC sensitivity to 5-FU by inhibiting the Wnt/β-catenin signaling pathway [32]. Blocking of Wnt/β-catenin signaling improves CD8 + T cell infiltration and initiation, creating a more favorable situation for immune checkpoint inhibition [199]. Combination therapy with Wnt inhibitors and immune checkpoint inhibition is an exciting strategy that should be evaluated in preclinical and clinical studies. However, active Wnt signaling disrupts the transcriptional activity of Foxp3, which is critical for the development and function of regulatory T cells (Tregs) [200]. Therefore, it must be recognized that blocking Wnt for immune regulation is still controversial at present. In conclusion, the Wnt/β-catenin signaling pathway may be a promising target for reversing tumor resistance.

7.2. JAK/STAT

The Janus kinases (JAK) and signal transducer and activator of transcription (STAT) are essential in regulating cell survival, proliferation, differentiation and apoptosis as the convergence points of many oncogenic signaling pathways [201]. STAT3, a member of the STAT transcription factor family, is upregulated in various tumors, including CRC [202]. It has been shown that STAT3 expression level in CRC is positively correlated with chemoradiotherapy resistance [203]. This may be due to IL-6-induced STAT3 phosphorylation, and inhibition of STAT3 reverses CRC resistance to chemotherapy in vivo and vitro [203]. In addition, nuclear pyruvate kinase isoform M2 (PKM2) promotes gefitinib resistance by upregulation of STAT3 activation in CRC, and inhibition of STAT3 or PKM2 can reverse CRC against EGFR resistance [204].
IFN-γ is mainly produced by T cells, NK cells and NK T cells, and IFN-γ signal must be activated through JAK1/2/STAT1 pathway, suggesting that IFN-γ/JAK1/2/STAT1 pathway may be involved in anti-tumor immune response [205,206]. Initially, IFN-γ signaling can play an anti-tumor role by increasing antigen presentation, such as upregulation of major histocompatibility complex (MHC) molecules and recruitment of immune cells [207]. However, continued IFN-γ activation causes immune reprogramming of tumor cells, leading to tumor immune escape [208]. Tumor cells evade IFN-γ by downregulating or mutating the molecules involved in the IFN-γ signaling pathway, such as JAK1/2 and STAT, thus blocking the IFN-γ signaling pathway. Preclinical studies have shown that defects in IFNGR1 (IFN-γ receptors) and JAK1/2 genes cause tumor cells to be unresponsive to IFN- γ, which allows tumor cells to grow in the presence of IFN- γ [209]. JAK1/2-inactivating mutations reduce programmed death ligand 1 (PD-L1) and MHC I expression in mouse and human CRC [33,210]. This implies that the inactivation mutations of JAK1 and JAK2 block the IFN-γ signaling pathway, leading to downregulation of tumor PD-L1 expression, so it would be unnecessary to attempt anti-PD-L1 therapy in this case. Inactivation mutations of JAK1 and JAK2 have been observed in a biopsy specimen of high TMB colon cancer that did not respond to PD-1 therapy, further highlighting the importance of IFN-γ deficiency against PD-1 resistance [210]. In addition, the expression of MHC class I molecules in a variety of tumors mainly depends on IFN-γ signal [205]. Preclinical studies have shown that hypoxia induces a decrease in IFN-γ signaling in a mouse CRC model, leading to a decrease in MHC class I molecule expression [211]. A positive correlation between IRF2 expression and MHC I has been observed in many human tumors, including CRC [212].
Multiple pathways have been developed, such as toll-like receptor (TLR) agonists, oncolytic viruses or other pathways that activate alternative interferon pathways (type I IFN) that lead to signal transduction and transcriptional activator 1 (STAT 1) and STAT 2 signaling, thereby promoting transcription of PD-L1 and MHC classes by inducing IRF1 [33,213].

7.3. TGF-β

TGF-β signaling activation plays a key role in cancer progression and is associated with tumor treatment resistance [214]. Preclinical studies have shown that 5-FU can activate the TGF-βRI/Smad3 pathway to confer chemotherapeutic resistance to CRC; TGF-βRI inhibition reduces the proliferation of chemoresistant cancer cells and increases cell death [215]. In addition, it has been reported that the effect of TGF-β on chemotherapy resistance may require the coordination of protease-activated receptor 2 (PAR2) because the activation of PAR2 eliminates the inhibition of TGF-β [216]. In addition, TGF-β receptor maturation may be disrupted by mediator protein complex subunit 12 (MED12), which leads to activation of TGF-β signaling pathway and upregulation of mesenchymal marker expression, further conferring chemotherapeutic tolerance in CRC patients [217]. MED12 deletion, TGF-β receptor overexpression or recombinant TGF-β therapy induces TKI resistance in various solid tumors, including CRC [37]. This may be due to TGF-β signaling inducing MEK/ERK signal activation, thus restoring the reduction in TKI-mediated MAPK pathway inhibition. Therefore, MED12 may be a promising target in the future.
TGF-β can provide an escape route for tumor angiogenesis by influencing and regulating tumor angiogenesis [218]. Preclinical studies have shown that TGF-β receptor I inhibitors (SD-208) inhibit anti-angiogenesis in hepatocellular carcinoma (HCC) and glioblastoma multiforme (GBM) xenograft tumors [219,220]. However, SD-208 did not cause significant vascular inhibition in the CRC model, at least in the SW-48 cell line [221]. This suggests that TGF-β-mediated angiogenesis may be related to the corresponding tumor microenvironment. Activin receptor-like kinase 1 (ALK-1) is a TGF-β transmembrane receptor that plays a crucial role in TGF-β -mediated angiogenesis [218]. PF-03446962 is an IgG2 monoclonal antibody against human ALK-1. PF-03446962 has been shown in preclinical studies to reverse the sensitivity of tumors to antiangiogenic therapy [38]. Regorafenib is a multikinase inhibitor that targets VEGF R1-3, KIT and PDGFR-α/β. However, in a Phase Ib trial study, regorafenib combined with PF-03446962 did not show significant clinical activity in patients with refractory mCRC [81].
TGF-β signaling can produce profound immunosuppressive activity on crucial cell types of innate and adaptive immunity, thereby attenuating the inherent anti-tumor potential of immune cells within TME [214]. For example, mouse CRC models with low mutation, T cell rejection and TGF-β activation have difficulty benefiting from immune checkpoint inhibitors. This may be due to increased TGF-β promoting T cell rejection and blocking the Th1 effector phenotype. Blockade of the TGFβ signaling pathway promotes enough T cell infiltration to make tumors sensitive to PD-1 therapy [39]. Similar therapeutic responses to combined TGF-β and PD-L1 blockade were observed in other mouse CRC models [222,223].

7.4. EGFR

7.4.1. PI3K/AKT/mTOR

The activation of PI3K/AKT/mTOR signaling pathway would occur in 20% of CRC patients, and the activation of this pathway may be related to the expression of PD-L1 [224,225]. A series of preclinical studies has shown that PI3K/AKT induces high expression of PD-L1 in colon cancer stem cells and mediates immune escape [226,227]. In addition, as a negative regulator of PI3K/Akt/mTOR signaling pathway, loss of PTEN leads to upregulation of PD-L1 protein, decrease of CD8 + T cell and increase of TAM, which are negative predictors of PD-1 blocking response in microsatellite instability-high (MSI-H) CRC patients [228,229]. Inhibition of the PI3K/AKT/mTOR pathway improves the efficacy of anti-PD-1 and anti-CTLA-4 antibodies in mouse models [226,230,231].

7.4.2. RAS/RAF/MEK/ERK

The increased expression of PD-L1 has been observed in the specimens and cell lines of CRC patients with KRAS mutation, which may increase the stability of PD-L1 mRNA by regulating the AU-rich element-binding protein tristetraprolin (TTP), and the combined blocking of anti-EGFR and PD-1/PD-L1 significantly inhibits tumor growth [232,233]. In addition, recently published preclinical data suggest that KRAS mutations enhance resistance to PARPi and anti-PD-L1 therapy in CRC PDX models, which may be related to PMEK1/2 and pMAPK, and the addition of MEKi to the combination of PARPi and anti-PD-L1 therapy reverses drug resistance in KRAS-mutated tumors [23].
The MAPK axis is the main downstream pathway triggered by normal T cell receptors (TCRs) [234]. Continuous TCR stimulation drives negative regulator PD-1 expression and inhibits effector T cell function [235]. MEK is an essential part of MAPK signaling pathway, and inhibition of MEK may rejuvenate anti-tumor T cells [236]. In a mouse CRC model, MEK inhibition protects CD8+T cells from chronic TCR-stimulus-driven death while preserving cytotoxic activity. Combining MEK inhibition with anti-PD-L1 produces collaborative and durable tumor regression [24]. Moreover, MEK inhibits apoptosis-inducing markers, upregulates HLA expression and downregulates immunosuppressive factors such as PD-L1, IL1, IL8, NT5E and VEGFA [237]. This evidence suggests that immunoregulation by MEK inhibitors combined with immune checkpoint inhibitors is a more effective anti-tumor therapy. A Phase Ib trial study evaluated atezolizumab plus cobimetinib in chemical-refractory mCRC. This combination therapy had manageable safety and clinical activity [82]. However, in another Phase III trial study, the atezolizumab plus cobimetinib did not increase the tumor response or survival benefit of mCRC compared to treatment alone [83]. Moreover, a deleterious effect of MEK inhibition on T lymphocyte proliferation and antigen-specific T lymphocyte activation has been observed in melanoma [238,239]. This may partly explain why CRC patients do not benefit from MEK inhibitors and immune checkpoint inhibitors.

8. Pathway of Cell Death

When enough active drugs accumulate and inhibit their corresponding targets, treatment outcomes depend on how the cancer cells respond. Ideally, drug-induced damage is closely associated with cell death. Apoptosis, autophagy and ferroptosis are the main modes of cell death pathway [240]. The dysfunction of programmed death pathways is related to tumor drug resistance (Figure 2).

8.1. Apoptosis

Intrinsic or acquired resistance to apoptosis is one of the characteristics of human cancers [241]. Apoptosis can pass through two signal transduction pathways, such as the death receptor (external) pathway and the mitochondrial (internal) pathway, both of which ultimately lead to the cleavage of the executioner caspase-3 [241]. Therefore, aberration of these pathways may amplify the degree of tumor treatment resistance by inhibiting apoptosis pathways.
The internal apoptosis pathway is mainly regulated by members of the Bcl-2 family [242]. The Bcl-2 protein family consists of anti-apoptotic proteins, such as Bcl-2, Bcl-XL and McL-1, and pro-apoptotic molecules, such as Bax and BH3-only proteins [243]. Dysregulation of antiapoptotic protein and proapoptotic protein may induce drug resistance in the tumor. Preclinical studies have shown that 5-FU leakage will increase the expression of Bcl-2 and Bax in CRC, and the decrease or deletion of Bax expression will increase the resistance of CRC to 5-FU and OXA [244].
Therefore, targeting the Bcl-2 protein family may be an effective way to reverse drug resistance. ABT-737 is a small molecule BH3 analog. ABT-737 and its oral derivative ABT-263 (navitoclax) can target anti-apoptotic proteins and promote the pro-apoptotic ability of Bax and Bak [245]. ABT-263 in combination with selumetinib increased caspase-dependent cell death through Bax and effectively reduced the incidence of CRC-acquired resistance to selumetinib [40]. However, resistance mechanisms limit the effectiveness of these drugs. The most apparent cause is another anti-apoptotic member of the Bcl-2 protein family, McL-1 [245,246]. Obatoclax is a promising novel agent that simultaneously targets members of the entire anti-apoptotic Bcl-2 protein family, including McL-1 [247]. However, there are few studies on CRC by obatoclax. Timme et al. [248] showed that obatoclax could restore oxaliplatin-induced CRC cell apoptosis by blocking McL-1 and/or Bcl-XL.
The inhibitor of apoptosis protein (IAP) family is a crucial regulator of programmed death pathway, which can block the activation of caspase and negatively regulate apoptosis pathway [249]. The IAP family mainly includes neuronal IAP (NIAP), cellular IAP1 (cIAP1), cellular IAP2 (cIAP2), X chromosome link IAP (XIAP) and survivalin [250]. Preclinical studies have shown that XIAP can promote resistance of CRC cells to apoptosis induced by multiple tumor treatments [251,252,253]. miR-587 enhances AKT activation and XIAP signal transduction by downregulating the expression of negative AKT regulator PPP2R1B, thereby inhibiting 5-FU-induced apoptosis [254].
AEG35156 is a new second-generation antisense oligonucleotide targeting XIAP. In the mice CRC model, AEG35156 can reduce XIAP mRNA and significantly enhance the therapeutic effect of taxanes [41]. Second mitochondria-derived activator of caspases (SMAC) can inhibit IAP function, thereby inhibiting cell apoptosis. In the past 20 years, several small molecules simulating SMAC function have been developed [255]. Some studies have confirmed that birinapant, a SMAC analog, can improve the response of CRC cells to oxaliplatin/5-FU [42]. XIAP inhibitors and SMAC activators are promising drugs that enhance anti-cancer drug-mediated apoptosis of CRC cell.
DNA damage caused by chemotherapeutic drugs always leads to apoptosis of CRC cells. As a tumor suppressor gene, P53 mutation could disrupt DNA-damage-induced cell cycle arrest and inhibit tumor cell apoptosis, which could induce chemotherapy resistance in various solid tumors, including CRC [256,257]. A series of clinical studies have confirmed that CRC patients with wild-type P53 can benefit from 5-FU, and CRC patients with mutant P53 have a poor prognosis when receiving adjuvant chemotherapy [258,259,260].
Enhancing the activity of P53 may be a way to reverse chemotherapy resistance in CRC. Murine double minute 2 (MDM2) is a primary physiological antagonist of P53, which is inhibited by hiding the activation domain of P53 or by ubiquitin-ligase [261,262]. Therefore, the development of MDM2-based inhibitors may help reverse drug resistance. Nutlin-3 is a specific antagonist of MDM2-p53 interaction. Increased p53 activity has been reported after nutlin-3 treatment in breast, prostate and CRC cells [263,264]. In addition, TRIM67 interacted directly with the C-terminus of p53, inhibiting p53 degradation by its ubiquitin ligase MDM2 [265]. Preclinical studies have shown that TRIM67 can enhance the anti-proliferation effect of adriamycin, 5-FU or oxaliplatin in p53 wild-type CRC cells [265].

8.2. Autophagy

Autophagy is a self-degradative process that is important for balancing energy sources at critical times in development and in response to nutrient stress [266]. The role of autophagy in tumors is controversial and highly dependent on tumor stage and classification [267]. Autophagy acts as a tumor suppressor in precancerous lesions and inhibits tumorigenesis through type II programmed cell death [268]. Autophagy acts as a tumor promoter in identified tumors, which can degrade damaged proteins and organelles as reserve energy to supply tumor cells for survival in anti-cancer therapy and endue tumor cells with resistance to tumor therapy [269]. In the process of autophagy, different autophagy-related genes (ATG) are always regulated, such as ATG1 (ULK1), Vps34 (PIK3C3), Beclin1 (BECN1), LC3 and ATG5 [267].
In CRC patients undergoing chemotherapy, overexpression of autophagy-associated proteins, such as Beclin-1, was associated with low survival [270]. These results suggest that autophagy-associated proteins may be associated with chemotherapy resistance in CRC patients. Also, abundant Clostridium nuclei have been detected in CRC specimens of patients, with recurrence after chemotherapy. Clostridium nuclei activate autophagy pathways and alter the chemotherapy response of CRC by targeting TLR4 and MYD88 innate immune signaling and specific microRNA [271]. IL-6 activates autophagy and promotes chemotherapy resistance in CRC through the IL-6/JAK2/BECN1 pathway [272]. Mechanistically, IL-6 promotes the phosphorylation of BECN1-Y333 through JAK2 signaling, leading to autophagy [272]. In addition, non-coding RNA can mediate CRC resistance to chemotherapy by enhancing autophagy pathways, such as circHIPK3, LncRNA SNHG6, LncRNA SNHG14, LncRNA H19 and LncRNA. NEAT1 has been shown to mediate CRC resistance to chemotherapy by enhancing the autophagy pathway [273,274,275,276,277]. A series of preclinical studies have demonstrated that autophagy inhibitors, such as chloroquine/3-methyladenine, can reverse CRC resistance by inhibiting 5-FU-induced activation of autophagy [43,44].
Elevated levels of autophagy-associated protein have also been observed in patients with advanced CRC treated with cetuximab, and autophagy-associated protein levels predict patient response to cetuximab [278]. It is possible that cetuximab induces autophagy in cancer cells by downregulating HIF-1α and Bcl-2 and activating the Beclin1/hVps34 complex. Autophagy inhibitor, such as chloroquine/3-MA, could enhance the sensitivity of CRC cells to cetuximab-induced cell death [45,46].
There is increasing evidence that autophagy is involved in immune system development and the survival and function of effector T cells [279]. Therefore, autophagy may be involved in tumor immune escape, which is based on immunogenicity loss. For example, in a mouse CT26 colon tumor model, mitoxantrone and oxaliplatin activate autophagy and enhance tumor-infiltrating dendritic cells and T cells, and depletion of autophagy-associated proteins (ATG5 or ATG7) reduces chemotherapy-induced anti-tumor immune responses [280]. In a mouse CRC model, the autophagy inhibitors, such as SB02024 and SAR405, increased NK and T cell infiltration in TME and reversed resistance to anti-PD-1 or anti-PD-L1 [47]. A clinical trial evaluates the combination of the autophagy inhibitor hydroxychloroquine, immune checkpoint inhibitor atezolizumab and MEK inhibitor cobimetinib in patients with advanced gastrointestinal tumors with KRAS mutations (NCT04214418).

8.3. Ferroptosis

Unlike autophagy and apoptosis, ferroptosis is a type of cell death caused by iron-dependent and lipid reactive oxygen species (ROS), first proposed by Dixon and colleagues in 2012 [281]. Several drugs, such as cisplatin and cetuximab, could induce ferroptosis, thus exerting its anti-tumor effect, which involves several vital molecules, NOX and P53, glutathione peroxidase 4 (GPX4), nuclear factor E2-related factor 2 (Nrf2) and solute carrier (SLC) family members. For example, GPX4, as a negative regulator of iron apoptosis, plays a major antioxidant role in anti-cancer activity and the inhibition of GPX4 can enhance ferroptosis and thus enhance the sensitivity of CRC to cisplatin/oxaliplatin [49,282]. Lipocalin 2 is an iron carrier binding protein that can regulate ferroptosis. It has been proven that lipocalin 2 inhibits ferroptosis and leads to 5-FU resistance in colon cancer cell lines [283]. The transcriptional regulator Nrf2 is another crucial inhibitor of iron apoptosis because it inhibits iron uptake, restricts ROS accumulation and upregulates SLC7A11 [284]. Higher Nrf2 was observed in CRC cells with 5-FU resistance, which further led to increased heme oxygenase-1 (HO-1) expression and reduced CRC sensitivity to 5-FU [285].
As mentioned earlier, RAS mutations limit the usefulness of anti-EGFR in mCRC patients. Ras-selective lethal 3 (RSL3) is a small molecule that kills RAS-mutated cancer cells and activates iron apoptosis in RAS-mutated cancer cells [286]. In KRAS mutant CRC cell lines and mouse CRC models, cetuximab promotes RSL3-induced ferroptosis by activating P38 MAPK to inhibit Nrf2/HO-1 signaling pathway [287]. β-elemene can be used as an inducer of iron apoptosis and the combination of cetuximab and β-elemene can be sensitive to KRAS mutant CRC cells by inducing ferroptosis and inhibiting EMT [48].
In addition, ferroptosis is thought to be involved in anti-tumor immune responses. For example, in melanoma models, CD8+ T cells modulate tumor ferroptosis during immunotherapy, which contributes to the anti-tumor efficacy of immunotherapy [288]. Dihydroartemisinin (DHA) is an active metabolite of artemisinin. Immune-stimulating nanodrugs (DHA and oxaliplatin) have been shown to induce ROS production in tumor cells, resulting in immunogenic programmed death and enhanced checkpoint blocking immunotherapy [289]. Recent literature suggests that DHA and pyropheophorbide-iron can produce and aggregate ROS to induce ferroptosis in an immunogenic manner, sensitizing non-immunogenic CRC, and can enhance the therapeutic effect of CRC anti-PD-L1 [290].

9. Tumor Microenvironment

In solid tumors, the TME consists of extracellular matrix, cancer-associated fibroblasts (CAFs), immune and inflammatory cells [291]. Components of the TME can act as therapeutic barriers for treating solid tumors [291]. The different cell types in TME are recruited to the tumor site in response to a series of therapeutic treatments that provide a sanctuary for tumor cells to escape drug-mediated cell death and develop resistance to tumor treatment drugs (Figure 3) [292].

9.1. Cancer-Associated Fibroblasts

CAF is a significant component of tumor stroma and expresses specific markers such as α smooth muscle actin (αSMA), fibroblast activation protein α (FAP) and fibroblast growth factor receptor (FGFR) [293]. In CRC, CAF-derived CCL2 upregulates FGFR4, which plays a role in tumor-matrix interactions and promotes EMT, which may constitute an essential factor in CRC resistance to 5-FU and oxaliplatin [294]. Transcriptional analysis in drug-resistant CRC cell lines showed that enhanced FGFR4 activity confers chemotherapeutic resistance, and FGFR4 silencing downregulation of anti-apoptotic proteins c-FLIP and Bcl-2 and decreasing STAT3 activity reverse the sensitivity of CRC to chemotherapy [50]. CAFs also mediate tumor resistance by secreting pro-inflammatory factors that alter tumor cell sensitivity and protect CRC cells from chemotherapeutic agents [295]. CAF can release TGF-β2 under hypoxia, synergize with HIF-1α, activate GLI2 expression through the non-canonical Hedgehog pathway and promote CRC resistance to 5-FU/oxaliplatin [296]. CAF-expressing Snail participates in CRC resistance to 5-FU/paclitaxel chemotherapy by secreting CCL1 to activate TGF-β and NF-κB signaling pathways, and inhibition of TGF-β and/or NF-κB signaling pathway can reverse CCL1-mediated multi-drug resistance [297]. Not only that, cytotoxicity itself activates CAF, which further promotes the self-renewal of CRC initiation cells, endowing CRC with chemotherapy resistance. This may be related to the secretion of interleukin-17A (IL-17A) induced by chemotherapy [298]. IL1β/TGFβ1 is a trigger for fibroblast recruitment and transformation into cancer-associated fibroblasts (CAF) in CRC and endows chemotherapy resistance [295]. However, the blocking of IL1β/TGFβ1 signaling did not increase sensitivity to chemotherapy, which may be a compensatory activation of the tak1-mediated non-canonical TGFβ pathway. Combined blocking of TGFβ -activated kinase 1 (TAK1) and TGFBR1 inhibits IL1β/TGFβ1-mediated fibroblast activation, reduces secretion of pro-inflammatory cytokines and makes tumor cells more sensitive to chemotherapy [295]. CAF-derived exosomes are also active agents of drug resistance. A series of studies have shown that CAF-derived exosomes, such as miR-92a-3p, H19 and CCAL, promote CRC EMT and lead to CRC metastasis and chemotherapy resistance by activating the Wnt/β-catenin pathway. Targeting these exosomes can effectively reduce autophagy activity and reverse CRC chemotherapy sensitivity [299,300,301].
During anti-EGFR therapy in CRC patients, CAF promotes tumor resistance to EGFR by secreting growth factors, such as FGF1, FGF2, HGF, TGF-β1 and TGF-β2 [302]. Conditional medium (CM) from CAF can confer anti-EGFR resistance in CRC stem-cell-like cells, which may be wild-type MET activated by CAF-derived HGF [303]. In addition, preclinical studies have demonstrated that FGFR4 overexpression secretes EGFR ligands, such as AREG, which activate EGFR and ERBB3, and FGFR4 overexpression reduces cetuximab-induced cytotoxicity. Compared with cetuximab, FGFR4 inhibitor (BLU9931) combined with cetuximab has a stronger anti-tumor effect [51].
Cell analysis of human and mouse CRC specimens showed that CAF is a major contributor to TGF-β production [39]. As previously mentioned, TGF-β activation in the TME can produce profound immunosuppressive activity against crucial cell types of innate and adaptive immunity. It has been demonstrated in a variety of tumor models, including CRC models, that activation of CAF may induce PD-1 and CTLA-4 expression on relevant immune cells, which leads to tumor insensitivity to immunotherapy or recurrence [39,304,305,306]. In addition, NADPH oxidase 4 (NOX4) is a common regulator of myofibroblast accumulation in many human cancers. NOX4 expression is closely related to CAF accumulation in myofibroblasts in patients with CRC, which may be attributed to the fact that NOX4 regulates the transdifferentiation from fibroblasts to myofibroblasts through the delayed period of intracellular ROS production, and the inhibition of NOX4 can effectively reduce the accumulation of CAF [53]. Preclinical studies have shown that MC38 models in CAF-rich mice have difficulty benefiting from tumor vaccines and anti-PD-1 therapy, possibly because CAFs exclude CD8+ T cells (not CD4+ T cells) from tumors by producing NOX4 T cells (or macrophages) to inhibit the response extensively, and pharmacological inhibition of NOX4 restores the immunotherapeutic response of CAF-rich tumors [306]. In addition, PD-L1 expression in CRC specimens is positively correlated with FGFR2 expression, which promotes PD-L1 expression in CRC by activating the JAK/STAT3 signaling pathway, which further promotes T cell apoptosis and leads to tumor immunosuppression. This phenomenon can be reversed by JAK/STAT3 signaling pathway inhibitors [233].

9.2. Myeloid-Derived Suppressor Cells

Myeloid-derived suppressor cells (MDSCs), a heterogeneous population of immature bone marrow cells, are one of the major immunosuppressive cell subpopulations in the tumor microenvironment and contribute to cancer immune avoidance by inhibiting the function of T and NK cells [307,308,309]. MDSCs are divided into three groups according to their histological characteristics: early or immature MDSC (E-MDSC/I-MDSC), mononuclear MDSCs (M-MDSCs) and polymorphonuclear/granulocyte MDSCs (PMN-MDSCs/G-MDSCs) [309,310]. The frequency of G-MDSCs was significantly higher than that of M-MDSCs in several CRC mouse models [308,311,312,313].
Immunosuppression is the main feature of MDSC in CRC TME. In the mouse CRC model, G-MDSCs expressed high arginase I (ARG1) in both peripheral blood and TME [313]; ARG1 depletes L-arginine, which maintains T cell proliferation and function, inhibits T cell proliferation and induces T cell apoptosis [314,315]. In addition, MDSC impairs Fc-receptor-mediated NK cell function by producing NO [316].
Cytokines play an important role in MDSC recruitment [317]. Preclinical studies have demonstrated that STAT1 recruits G-MDSC through IL-17 in colitis-associated colorectal cancer (CAC) models. Blockade of IL-17 reduced the recruitment of G-MDSC into intestinal specimen of STAT1 -/- mice and significantly decreased the expression of ARG1 and inducible nitric oxide synthase (iNOS) [317]. Other studies have reported that IL-8, GM-CSF, TNF-α, YAP1, CXCR2 and CCL2 promote G-MDSC recruitment in CRC and confer tumor immunotherapy resistance to CRC [308,311,318,319,320].
Inhibition of MDSC may be a promising way to reverse drug resistance to anti-PD-1. In a mouse colon-26 tumor model, inhibition of receptor tyrosine kinase KIT enhances the anti-tumor activity of immune checkpoint inhibitors (anti-CTLA-4 and anti-PD-1) by selectively reducing immunosuppressive M-MDSC clusters [54]. The humanized anti-Kit mAbs KTN0158 is currently being evaluated in patients with Kit-positive advanced solid tumors (NCT02642016). In addition, preclinical studies have demonstrated that TRAIL-R2 agonist antibodies can selectively target MDSC [321]. In a Phase I trial study, the efficacy of the novel agonist TRAIL-R antibody DS-8273A was evaluated, which caused a temporary reduction in the number of elevated MDSC in peripheral blood in most patients with advanced cancer, including CRC, to levels observed in healthy volunteers. In contrast, DS-8273A did not affect the number of neutrophil monocytes and other myeloid and lymphocyte populations (NCT02076451) [322].

9.3. Tumor-Associated Macrophages

Tumor-associated macrophages (TAMs) are one of the most abundant immune types in TME, influencing tumor progression by producing pro-inflammatory and anti-inflammatory cytokines [323]. In general, macrophages are divided into two subpopulations: classically activated, M1 macrophages and alternatively activated, M2 macrophages [323]. M1 macrophages (NOS2) are critical cellular components involved in inflammatory response and anti-tumor immunity. In contrast, M2 macrophages (CD163) play an anti-inflammatory and pro-tumor role, and TAMs often exhibit an “M2-like” phenotype [324]. Preclinical studies have shown that TAM (CD68) can be activated during 5-FU treatment and can give CRC cells against chemotherapy with 5-FU during 5-FU treatment, which may be caused by TAM secreting ornithine decarboxylase (ODC)-dependent decay activation JNK-caspase-3, which further leads to the CRC cells against chemotherapy with 5-FU [55]. Pharmacological or genetic blockade of ODC reverses TAM-induced chemotherapy resistance to 5-Fu in vitro and in vivo [55]. Data in the mouse MC38 tumor model suggest that TAM-derived milk-fat globule-epidermal growth factor-VIII (MFG-E8) mediates tumor chemotherapy resistance by activating STAT3 and Hedgehog signals in tumor stem cells [325]. In stage III CRC patients treated with 5-FU, L-OHP and leucovorin, high TAM (CD68) density is associated with treatment insensitivity. TAM-derived IL6 induces CRC chemotherapy resistance by activating the IL6R/STAT3/miR-204-5p axis, which can be reversed by miR-204-5P overexpression or IL6 blockers [326]. However, high densities of TAMs have been shown to significantly benefit stage III CRC patients treated with 5-FU [327]. It may be that 5-FU exposure is beneficial to the polarization of unpolarized macrophages towards M1, and that 5-FU and TAM (CD68) show a synergistic effect in increasing CRC cell death [327]. CD68 pan-macrophage marker does not allow subclassification of TAM, which may be responsible for the contradiction between prognosis and chemotherapy sensitivity. One instance of clinical data suggests that the M1/M2 ratio reflects the sensitivity of CRC patients to adjuvant therapy [328]. Therefore, the sensitivity of TAM to CRC chemotherapy depends on the M1/M2 ratio in TME.
TAM can mediate tumor immunosuppression by expressing immune checkpoints. In human and mouse CRC tumor models, high levels of PD-1 expression on TAM have low tumor phagocytosis, and treatment with checkpoint inhibitors targeting PD-1 or PD-L1 increases phagocytosis and reduces tumor growth [329]. Colony-stimulating factor 1 (CSF1) and its receptor CSF-1R play an essential role in the recruitment, differentiation and maintenance of immunosuppressive macrophages in tumors [330]. It has been shown in tumor-conditioned medium (TCM) differentiated macrophages that CSF-1 differentiated macrophages affect the frequency and the activity of T cells [56]. CSF-1R inhibition can specifically eliminate immunosuppressive M2-like macrophages and positively transfer the ratio of CD8 to CD4 to cytotoxic effector T cells [56]. An ongoing clinical study is evaluating the efficacy of anti-CSF1R (pexidartinib) in combination with anti-PD-L1 (durvalumab) in advanced/metastatic CRC or pancreatic cancer (NCT02777710). In addition, regorafenib also reduces TAMs by inhibiting the CSF-1R [331]. A combination of regorafenib and nivolumab showed activity in Japanese third-line or later pMMR mCRC with an ORR of 33.3% [332]. However, this strategy has not been proven effective in the North American population, with an ORR of 7.1% [333].
Therefore, TAM consumption or blocking cancer-induced M2-like macrophage programming has the potential to improve treatment efficiency and reduce anti-cancer drug resistance. Macrophages have been shown to be plastic [334,335,336]. Preclinical studies show that prostaglandin E2 (PGE2) receptor 4 (EP4), as the primary regulator of IMC, can mediate drug resistance by promoting the differentiation and proliferation of M2 TAM and MDSC [58]. TP-16 (EP4 antagonist) treatment polarized TAM from immunosuppressive M2 type to pro-inflammatory M1 type, which enhanced T-cell-mediated anti-tumor immunity and improved the efficacy of anti-PD-1 [58]. In addition, a series of studies have proved that Ganoderma lucidum spore polysaccharide (GLSP) [337], CMPB90-1 [338], natural M1 exosomes [339], conditional deletion of the microRNA (miRNA)-processing enzyme DICER [334] and TLR 7/8 agonist [57] can specifically eliminate M2-like phenotype or reset TAM from M2-like phenotype to M1-like phenotype to inhibit tumor development and reverse tumor treatment resistance.

9.4. Angiogenesis

Vascular endothelial growth factors (VEGFs) and their receptors (VEGFRs) are uniquely required to balance the formation of new blood vessels with the maintenance and remodeling of existing ones, during development and in adult tissues [340]. However, VEGF in the TME could drive immunosuppression by inducing vascular aberrations, inhibiting antigen-presenting and immune effector cells or enhancing the immunosuppressive activity of Tregs, MDSCs and TAMs [341]. In turn, immunosuppressive cells can drive angiogenesis, thereby creating a vicious cycle of suppressed anti-tumor immunity [341]. Bevacizumab is a humanized monoclonal antibody that binds to the VEGF-A isoform, thereby preventing the interaction of VEGF-A with VEGFR, thereby inhibiting the activation of the VEGF signaling pathway that promotes neovascularization [342]. Preclinical studies have demonstrated that VEGF blockade can reverse T cell failure and enhance the therapeutic effect of ICI [59,60]. However, in clinical studies, only patients with MSI-H CRC benefited from the combined blockade of VEGF and PD-1. For example, in a Phase Ib trial study evaluating the efficacy of a combination of atezolizumab and bevacizumab in MSI-H CRC patients, the disease control rate was 90% [84]. However, in another clinical study, the addition of atezolizumab did not improve survival compared with microsatellite-stable (MSS) CRC treated with bevacizumab and capecitabine [85]. Similar results were obtained in the NSABP C-08 study [343]. Since MSI-H tumors are hypermutated and highly immunogenic, they should evade the attack of the immune system in order to progress, and VEGF may be one of the critical molecules mediating immune evasion. Therefore, using VEGF inhibitors while inducing more immunogenicity of MSS tumors may be a promising approach. Phase 2 study results of CheckMate 9X8 show nivolumab plus standard of care (SOC, fluorouracil/leucovorin/oxaliplatin/bevacizumab) This shows higher response rates in first-line treatment of mCRC compared to SOC [344].
High expression of VEGF is also associated with drug resistance to EGFR therapy. Cetuximab can reduce VEGF expression, and high VEGF expression is associated with lower response rates and shortened PFS in mCRC patients treated with cetuximab [345]. Therefore, anti-VEGF mAbs may reverse EGFR resistance in mCRC patients. In a mouse CRC xenograft model, combined blocking of VEGF and EGFR significantly inhibited tumor growth and angiogenesis [61]. However, in another preclinical study, the combination of bevacizumab and cetuximab did not increase apoptotic tumor cell death compared to either drug alone [62]. In a clinical study, a combination of cetuximab and regorafenib was used in 17 mCRC patients who had previously received both anti-VEGF and anti-EGFR therapy. Eight mCRC patients showed clinical benefit, including partial response in one patient and stable disease in seven [86].

10. Conclusions

Although there is use of multiple agents in clinical practice, such as chemotherapy, targeted therapy and immunotherapy, the emergence of drug resistance is the most critical factor in the poor prognosis of CRC. Tumor cells can protect themselves from treatment effects by changing drug metabolism, drug transport, drug targets and cell death pathways. In addition, changes in these pathways can also mediate drug resistance by controlling several basic signaling pathways involved in drug resistance. Recently, more and more attention has been paid to the role of TME in drug resistance. We reviewed the data to show that changes in pathways from drug metabolism to eventual cell death play a key role in CRC cell drug resistance. Although many potential therapeutic targets for these pathways are described in our review, most of the data are based on limited preclinical data, and larger, multi-institutional studies including more preclinical and clinical analyses are needed to assess clinical significance of these potential therapeutic targets. In addition, many preclinical studies have shown that multi-drug combination therapy may be an effective way to overcome drug resistance, and further clinical studies are urgently needed to evaluate the effectiveness of combination therapy in improving drug resistance.

Author Contributions

Q.W. and X.S. designed and wrote the manuscript, drafted the figures and searched relevant literature. G.C. and J.D. supervised this review. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (No. 81970500, X.S. and 81870393, J.D.).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  2. Malet-Martino, M.; Martino, R. Clinical studies of three oral prodrugs of 5-fluorouracil (capecitabine, UFT, S-1): A review. Oncologist 2002, 7, 288–323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Kosuri, K.V.; Wu, X.; Wang, L.; Villalona-Calero, M.A.; Otterson, G.A. An epigenetic mechanism for capecitabine resistance in mesothelioma. Biochem. Biophys. Res. Commun. 2010, 391, 1465–1470. [Google Scholar] [CrossRef] [PubMed]
  4. Longley, D.B.; Harkin, D.P.; Johnston, P.G. 5-fluorouracil: Mechanisms of action and clinical strategies. Nat. Rev. Cancer 2003, 3, 330–338. [Google Scholar] [CrossRef]
  5. Blondy, S.; David, V.; Verdier, M.; Mathonnet, M.; Perraud, A.; Christou, N. 5-Fluorouracil resistance mechanisms in colorectal cancer: From classical pathways to promising processes. Cancer Sci. 2020, 111, 3142–3154. [Google Scholar] [CrossRef]
  6. Fujii, R.; Seshimo, A.; Kameoka, S. Relationships between the expression of thymidylate synthase, dihydropyrimidine dehydrogenase, and orotate phosphoribosyltransferase and cell proliferative activity and 5-fluorouracil sensitivity in colorectal carcinoma. Int. J. Clin. Oncol. 2003, 8, 72–78. [Google Scholar] [CrossRef]
  7. Humeniuk, R.; Menon, L.G.; Mishra, P.J.; Gorlick, R.; Sowers, R.; Rode, W.; Pizzorno, G.; Cheng, Y.-C.; Kemeny, N.; Bertino, J.R.; et al. Decreased levels of UMP kinase as a mechanism of fluoropyrimidine resistance. Mol. Cancer Ther. 2009, 8, 1037–1044. [Google Scholar] [CrossRef] [Green Version]
  8. Griffith, M.; Mwenifumbo, J.C.; Cheung, P.Y.; Paul, J.E.; Pugh, T.J.; Tang, M.J.; Chittaranjan, S.; Morin, R.D.; Asano, J.K.; Ally, A.A.; et al. Novel mRNA isoforms and mutations of uridine monophosphate synthetase and 5-fluorouracil resistance in colorectal cancer. Pharm. J. 2013, 13, 148–158. [Google Scholar] [CrossRef] [Green Version]
  9. Salonga, D.; Danenberg, K.D.; Johnson, M.; Metzger, R.; Groshen, S.; Tsao-Wei, D.D.; Lenz, H.J.; Leichman, C.G.; Leichman, L.; Diasio, R.B.; et al. Colorectal tumors responding to 5-fluorouracil have low gene expression levels of dihydropyrimidine dehydrogenase, thymidylate synthase, and thymidine phosphorylase. Clin. Cancer Res. 2000, 6, 1322–1327. [Google Scholar]
  10. Soong, R.; Shah, N.; Salto-Tellez, M.; Tai, B.C.; Soo, R.A.; Han, H.C.; Ng, S.S.; Tan, W.L.; Zeps, N.; Joseph, D.; et al. Prognostic significance of thymidylate synthase, dihydropyrimidine dehydrogenase and thymidine phosphorylase protein expression in colorectal cancer patients treated with or without 5-fluorouracil-based chemotherapy. Ann. Oncol. 2008, 19, 915–919. [Google Scholar] [CrossRef]
  11. Lindskog, E.B.; Derwinger, K.; Gustavsson, B.; Falk, P.; Wettergren, Y. Thymidine phosphorylase expression is associated with time to progression in patients with metastatic colorectal cancer. BMC Clin. Pathol. 2014, 14, 25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Hu, T.; To, K.K.W.; Wang, L.; Zhang, L.; Lu, L.; Shen, J.; Chan, R.L.Y.; Li, M.; Yeung, J.H.K.; Cho, C.H. Reversal of P-glycoprotein (P-gp) mediated multidrug resistance in colon cancer cells by cryptotanshinone and dihydrotanshinone of Salvia miltiorrhiza. Phytomedicine 2014, 21, 1264–1272. [Google Scholar] [CrossRef] [PubMed]
  13. Yuan, Z.; Shi, X.; Qiu, Y.; Jia, T.; Yuan, X.; Zou, Y.; Liu, C.; Yu, H.; Yuan, Y.; He, X.; et al. Reversal of P-gp-mediated multidrug resistance in colon cancer by cinobufagin. Oncol. Rep. 2017, 37, 1815–1825. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Min, Y.D.; Yang, M.C.; Lee, K.H.; Kim, K.R.; Choi, S.U.; Lee, K.R. Protoberberine alkaloids and their reversal activity of P-gp expressed multidrug resistance (MDR) from the rhizome of Coptis japonica Makino. Arch. Pharm. Res. 2006, 29, 757–761. [Google Scholar] [CrossRef] [PubMed]
  15. Lo, Y.-L.; Liu, Y. Reversing multidrug resistance in Caco-2 by silencing MDR1, MRP1, MRP2, and BCL-2/BCL-xL using liposomal antisense oligonucleotides. PLoS ONE 2014, 9, e90180. [Google Scholar] [CrossRef]
  16. Lo, Y.-L.; Liu, Y.; Tsai, J.-C. Overcoming multidrug resistance using liposomal epirubicin and antisense oligonucleotides targeting pump and nonpump resistances in vitro and in vivo. Biomed. Pharmacother. 2013, 67, 261–267. [Google Scholar] [CrossRef]
  17. Wu, Y.; Wang, K.-y.; Li, Z.; Liu, Y.-p.; Izumi, H.; Uramoto, H.; Nakayama, Y.; Ito, K.-i.; Kohno, K. Y-box binding protein 1 enhances DNA topoisomerase 1 activity and sensitivity to camptothecin via direct interaction. J. Exp. Clin. Cancer Res. 2014, 33, 112. [Google Scholar] [CrossRef] [Green Version]
  18. Dienstmann, R.; Patnaik, A.; Garcia-Carbonero, R.; Cervantes, A.; Benavent, M.; Roselló, S.; Tops, B.B.J.; van der Post, R.S.; Argilés, G.; Skartved, N.J.Ø.; et al. Safety and Activity of the First-in-Class Sym004 Anti-EGFR Antibody Mixture in Patients with Refractory Colorectal Cancer. Cancer Discov. 2015, 5, 598–609. [Google Scholar] [CrossRef] [Green Version]
  19. Arena, S.; Siravegna, G.; Mussolin, B.; Kearns, J.D.; Wolf, B.B.; Misale, S.; Lazzari, L.; Bertotti, A.; Trusolino, L.; Adjei, A.A.; et al. MM-151 overcomes acquired resistance to cetuximab and panitumumab in colorectal cancers harboring EGFR extracellular domain mutations. Sci. Transl. Med. 2016, 8, 324ra14. [Google Scholar] [CrossRef] [Green Version]
  20. Bagchi, A.; Haidar, J.N.; Eastman, S.W.; Vieth, M.; Topper, M.; Iacolina, M.D.; Walker, J.M.; Forest, A.; Shen, Y.; Novosiadly, R.D.; et al. Molecular Basis for Necitumumab Inhibition of EGFR Variants Associated with Acquired Cetuximab Resistance. Mol. Cancer Ther. 2018, 17, 521–531. [Google Scholar] [CrossRef] [Green Version]
  21. Troiani, T.; Napolitano, S.; Martini, G.; Martinelli, E.; Cardone, C.; Normanno, N.; Vitagliano, D.; Morgillo, F.; Fenizia, F.; Lambiase, M.; et al. Maintenance Treatment with Cetuximab and BAY86-9766 Increases Antitumor Efficacy of Irinotecan plus Cetuximab in Human Colorectal Cancer Xenograft Models. Clin. Cancer Res. 2015, 21, 4153–4164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Misale, S.; Arena, S.; Lamba, S.; Siravegna, G.; Lallo, A.; Hobor, S.; Russo, M.; Buscarino, M.; Lazzari, L.; Sartore-Bianchi, A.; et al. Blockade of EGFR and MEK intercepts heterogeneous mechanisms of acquired resistance to anti-EGFR therapies in colorectal cancer. Sci. Transl. Med. 2014, 6, 224ra26. [Google Scholar] [CrossRef] [PubMed]
  23. Yang, B.; Li, X.; Fu, Y.; Guo, E.; Ye, Y.; Li, F.; Liu, S.; Xiao, R.; Liu, C.; Lu, F.; et al. MEK Inhibition Remodels the Immune Landscape of Mutant Tumors to Overcome Resistance to PARP and Immune Checkpoint Inhibitors. Cancer Res. 2021, 81, 2714–2729. [Google Scholar] [CrossRef]
  24. Ebert, P.J.R.; Cheung, J.; Yang, Y.; McNamara, E.; Hong, R.; Moskalenko, M.; Gould, S.E.; Maecker, H.; Irving, B.A.; Kim, J.M.; et al. MAP Kinase Inhibition Promotes T Cell and Anti-tumor Activity in Combination with PD-L1 Checkpoint Blockade. Immunity 2016, 44, 609–621. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Morris, E.J.; Jha, S.; Restaino, C.R.; Dayananth, P.; Zhu, H.; Cooper, A.; Carr, D.; Deng, Y.; Jin, W.; Black, S.; et al. Discovery of a novel ERK inhibitor with activity in models of acquired resistance to BRAF and MEK inhibitors. Cancer Discov. 2013, 3, 742–750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Cunningham, M.P.; Thomas, H.; Marks, C.; Green, M.; Fan, Z.; Modjtahedi, H. Co-targeting the EGFR and IGF-IR with anti-EGFR monoclonal antibody ICR62 and the IGF-IR tyrosine kinase inhibitor NVP-AEW541 in colorectal cancer cells. Int. J. Oncol. 2008, 33, 1107–1113. [Google Scholar]
  27. Bertotti, A.; Migliardi, G.; Galimi, F.; Sassi, F.; Torti, D.; Isella, C.; Corà, D.; Di Nicolantonio, F.; Buscarino, M.; Petti, C.; et al. A molecularly annotated platform of patient-derived xenografts (“xenopatients”) identifies HER2 as an effective therapeutic target in cetuximab-resistant colorectal cancer. Cancer Discov. 2011, 1, 508–523. [Google Scholar] [CrossRef] [Green Version]
  28. Luca, T.; Barresi, V.; Privitera, G.; Musso, N.; Caruso, M.; Condorelli, D.F.; Castorina, S. In vitro combined treatment with cetuximab and trastuzumab inhibits growth of colon cancer cells. Cell Prolif. 2014, 47, 435–447. [Google Scholar] [CrossRef]
  29. Kuwada, S.K.; Scaife, C.L.; Kuang, J.; Li, X.; Wong, R.F.; Florell, S.R.; Coffey, R.J.; Gray, P.D. Effects of trastuzumab on epidermal growth factor receptor-dependent and -independent human colon cancer cells. Int. J. Cancer 2004, 109, 291–301. [Google Scholar] [CrossRef]
  30. Koganemaru, S.; Kuboki, Y.; Koga, Y.; Kojima, T.; Yamauchi, M.; Maeda, N.; Kagari, T.; Hirotani, K.; Yasunaga, M.; Matsumura, Y.; et al. U3-1402, a Novel HER3-Targeting Antibody-Drug Conjugate, for the Treatment of Colorectal Cancer. Mol. Cancer Ther. 2019, 18, 2043–2050. [Google Scholar] [CrossRef] [Green Version]
  31. Wang, T.; Chen, Z.; Zhu, Y.; Pan, Q.; Liu, Y.; Qi, X.; Jin, L.; Jin, J.; Ma, X.; Hua, D. Inhibition of transient receptor potential channel 5 reverses 5-Fluorouracil resistance in human colorectal cancer cells. J. Biol. Chem. 2015, 290, 448–456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Li, Q.; Yang, T.; Li, D.; Ding, F.; Bai, G.; Wang, W.; Sun, H. Knockdown of aquaporin-5 sensitizes colorectal cancer cells to 5-fluorouracil via inhibition of the Wnt-β-catenin signaling pathway. Biochem. Cell Biol. 2018, 96, 572–579. [Google Scholar] [CrossRef] [PubMed]
  33. Torrejon, D.Y.; Abril-Rodriguez, G.; Champhekar, A.S.; Tsoi, J.; Campbell, K.M.; Kalbasi, A.; Parisi, G.; Zaretsky, J.M.; Garcia-Diaz, A.; Puig-Saus, C.; et al. Overcoming Genetically Based Resistance Mechanisms to PD-1 Blockade. Cancer Discov. 2020, 10, 1140–1157. [Google Scholar] [CrossRef] [PubMed]
  34. Damiano, V.; Caputo, R.; Garofalo, S.; Bianco, R.; Rosa, R.; Merola, G.; Gelardi, T.; Racioppi, L.; Fontanini, G.; De Placido, S.; et al. TLR9 agonist acts by different mechanisms synergizing with bevacizumab in sensitive and cetuximab-resistant colon cancer xenografts. Proc. Natl. Acad. Sci. USA 2007, 104, 12468–12473. [Google Scholar] [CrossRef] [Green Version]
  35. Damiano, V.; Caputo, R.; Bianco, R.; D’Armiento, F.P.; Leonardi, A.; De Placido, S.; Bianco, A.R.; Agrawal, S.; Ciardiello, F.; Tortora, G. Novel toll-like receptor 9 agonist induces epidermal growth factor receptor (EGFR) inhibition and synergistic antitumor activity with EGFR inhibitors. Clin. Cancer Res. 2006, 12, 577–583. [Google Scholar] [CrossRef] [Green Version]
  36. Rosa, R.; Melisi, D.; Damiano, V.; Bianco, R.; Garofalo, S.; Gelardi, T.; Agrawal, S.; Di Nicolantonio, F.; Scarpa, A.; Bardelli, A.; et al. Toll-like receptor 9 agonist IMO cooperates with cetuximab in K-ras mutant colorectal and pancreatic cancers. Clin. Cancer Res. 2011, 17, 6531–6541. [Google Scholar] [CrossRef] [Green Version]
  37. Huang, S.; Hölzel, M.; Knijnenburg, T.; Schlicker, A.; Roepman, P.; McDermott, U.; Garnett, M.; Grernrum, W.; Sun, C.; Prahallad, A.; et al. MED12 controls the response to multiple cancer drugs through regulation of TGF-β receptor signaling. Cell 2012, 151, 937–950. [Google Scholar] [CrossRef] [Green Version]
  38. Hu-Lowe, D.D.; Chen, E.; Zhang, L.; Watson, K.D.; Mancuso, P.; Lappin, P.; Wickman, G.; Chen, J.H.; Wang, J.; Jiang, X.; et al. Targeting activin receptor-like kinase 1 inhibits angiogenesis and tumorigenesis through a mechanism of action complementary to anti-VEGF therapies. Cancer Res. 2011, 71, 1362–1373. [Google Scholar] [CrossRef] [Green Version]
  39. Tauriello, D.V.F.; Palomo-Ponce, S.; Stork, D.; Berenguer-Llergo, A.; Badia-Ramentol, J.; Iglesias, M.; Sevillano, M.; Ibiza, S.; Cañellas, A.; Hernando-Momblona, X.; et al. TGFβ drives immune evasion in genetically reconstituted colon cancer metastasis. Nature 2018, 554, 538–543. [Google Scholar] [CrossRef] [Green Version]
  40. Sale, M.J.; Cook, S.J. The BH3 mimetic ABT-263 synergizes with the MEK1/2 inhibitor selumetinib/AZD6244 to promote BIM-dependent tumour cell death and inhibit acquired resistance. Biochem. J. 2013, 450, 285–294. [Google Scholar] [CrossRef]
  41. LaCasse, E.C.; Cherton-Horvat, G.G.; Hewitt, K.E.; Jerome, L.J.; Morris, S.J.; Kandimalla, E.R.; Yu, D.; Wang, H.; Wang, W.; Zhang, R.; et al. Preclinical characterization of AEG35156/GEM 640, a second-generation antisense oligonucleotide targeting X-linked inhibitor of apoptosis. Clin. Cancer Res. 2006, 12, 5231–5241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Fichtner, M.; Bozkurt, E.; Salvucci, M.; McCann, C.; McAllister, K.A.; Halang, L.; Düssmann, H.; Kinsella, S.; Crawford, N.; Sessler, T.; et al. Molecular subtype-specific responses of colon cancer cells to the SMAC mimetic Birinapant. Cell Death Dis. 2020, 11, 1020. [Google Scholar] [CrossRef] [PubMed]
  43. Sasaki, K.; Tsuno, N.H.; Sunami, E.; Tsurita, G.; Kawai, K.; Okaji, Y.; Nishikawa, T.; Shuno, Y.; Hongo, K.; Hiyoshi, M.; et al. Chloroquine potentiates the anti-cancer effect of 5-fluorouracil on colon cancer cells. BMC Cancer 2010, 10, 370. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Li, J.; Hou, N.; Faried, A.; Tsutsumi, S.; Takeuchi, T.; Kuwano, H. Inhibition of autophagy by 3-MA enhances the effect of 5-FU-induced apoptosis in colon cancer cells. Ann. Surg. Oncol. 2009, 16, 761–771. [Google Scholar] [CrossRef] [PubMed]
  45. Li, X.; Fan, Z. The epidermal growth factor receptor antibody cetuximab induces autophagy in cancer cells by downregulating HIF-1alpha and Bcl-2 and activating the beclin 1/hVps34 complex. Cancer Res. 2010, 70, 5942–5952. [Google Scholar] [CrossRef] [Green Version]
  46. Chen, Z.; Gao, S.; Wang, D.; Song, D.; Feng, Y. Colorectal cancer cells are resistant to anti-EGFR monoclonal antibody through adapted autophagy. Am. J. Transl. Res. 2016, 8, 1190–1196. [Google Scholar]
  47. Noman, M.Z.; Parpal, S.; Van Moer, K.; Xiao, M.; Yu, Y.; Viklund, J.; De Milito, A.; Hasmim, M.; Andersson, M.; Amaravadi, R.K.; et al. Inhibition of Vps34 reprograms cold into hot inflamed tumors and improves anti-PD-1/PD-L1 immunotherapy. Sci. Adv. 2020, 6, eaax7881. [Google Scholar] [CrossRef]
  48. Chen, P.; Li, X.; Zhang, R.; Liu, S.; Xiang, Y.; Zhang, M.; Chen, X.; Pan, T.; Yan, L.; Feng, J.; et al. Combinative treatment of β-elemene and cetuximab is sensitive to KRAS mutant colorectal cancer cells by inducing ferroptosis and inhibiting epithelial-mesenchymal transformation. Theranostics 2020, 10, 5107–5119. [Google Scholar] [CrossRef]
  49. Zhang, X.; Sui, S.; Wang, L.; Li, H.; Zhang, L.; Xu, S.; Zheng, X. Inhibition of tumor propellant glutathione peroxidase 4 induces ferroptosis in cancer cells and enhances anticancer effect of cisplatin. J. Cell Physiol. 2020, 235, 3425–3437. [Google Scholar] [CrossRef]
  50. Turkington, R.C.; Longley, D.B.; Allen, W.L.; Stevenson, L.; McLaughlin, K.; Dunne, P.D.; Blayney, J.K.; Salto-Tellez, M.; Van Schaeybroeck, S.; Johnston, P.G. Fibroblast growth factor receptor 4 (FGFR4): A targetable regulator of drug resistance in colorectal cancer. Cell Death Dis. 2014, 5, e1046. [Google Scholar] [CrossRef] [Green Version]
  51. Hong, C.-S.; Sun, E.-G.; Choi, J.-N.; Kim, D.-H.; Kim, J.-H.; Ryu, K.-H.; Shim, H.-J.; Hwang, J.-E.; Bae, W.-K.; Kim, H.-R.; et al. Fibroblast growth factor receptor 4 increases epidermal growth factor receptor (EGFR) signaling by inducing amphiregulin expression and attenuates response to EGFR inhibitors in colon cancer. Cancer Sci. 2020, 111, 3268–3278. [Google Scholar] [CrossRef] [PubMed]
  52. Napolitano, S.; Martini, G.; Rinaldi, B.; Martinelli, E.; Donniacuo, M.; Berrino, L.; Vitagliano, D.; Morgillo, F.; Barra, G.; De Palma, R.; et al. Primary and Acquired Resistance of Colorectal Cancer to Anti-EGFR Monoclonal Antibody Can Be Overcome by Combined Treatment of Regorafenib with Cetuximab. Clin. Cancer Res. 2015, 21, 2975–2983. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Hanley, C.J.; Mellone, M.; Ford, K.; Thirdborough, S.M.; Mellows, T.; Frampton, S.J.; Smith, D.M.; Harden, E.; Szyndralewiez, C.; Bullock, M.; et al. Targeting the Myofibroblastic Cancer-Associated Fibroblast Phenotype Through Inhibition of NOX4. J. Natl. Cancer Inst. 2018, 110, 109–120. [Google Scholar] [CrossRef] [PubMed]
  54. Garton, A.J.; Seibel, S.; Lopresti-Morrow, L.; Crew, L.; Janson, N.; Mandiyan, S.; Trombetta, E.S.; Pankratz, S.; LaVallee, T.M.; Gedrich, R. Anti-KIT Monoclonal Antibody Treatment Enhances the Antitumor Activity of Immune Checkpoint Inhibitors by Reversing Tumor-Induced Immunosuppression. Mol. Cancer Ther. 2017, 16, 671–680. [Google Scholar] [CrossRef] [Green Version]
  55. Zhang, X.; Chen, Y.; Hao, L.; Hou, A.; Chen, X.; Li, Y.; Wang, R.; Luo, P.; Ruan, Z.; Ou, J.; et al. Macrophages induce resistance to 5-fluorouracil chemotherapy in colorectal cancer through the release of putrescine. Cancer Lett. 2016, 381, 305–313. [Google Scholar] [CrossRef]
  56. Ries, C.H.; Cannarile, M.A.; Hoves, S.; Benz, J.; Wartha, K.; Runza, V.; Rey-Giraud, F.; Pradel, L.P.; Feuerhake, F.; Klaman, I.; et al. Targeting tumor-associated macrophages with anti-CSF-1R antibody reveals a strategy for cancer therapy. Cancer Cell 2014, 25, 846–859. [Google Scholar] [CrossRef] [Green Version]
  57. Liu, Z.; Xie, Y.; Xiong, Y.; Liu, S.; Qiu, C.; Zhu, Z.; Mao, H.; Yu, M.; Wang, X. TLR 7/8 agonist reverses oxaliplatin resistance in colorectal cancer via directing the myeloid-derived suppressor cells to tumoricidal M1-macrophages. Cancer Lett. 2020, 469, 173–185. [Google Scholar] [CrossRef]
  58. Lu, W.; Yu, W.; He, J.; Liu, W.; Yang, J.; Lin, X.; Zhang, Y.; Wang, X.; Jiang, W.; Luo, J.; et al. Reprogramming immunosuppressive myeloid cells facilitates immunotherapy for colorectal cancer. EMBO Mol. Med. 2021, 13, e12798. [Google Scholar] [CrossRef]
  59. Voron, T.; Colussi, O.; Marcheteau, E.; Pernot, S.; Nizard, M.; Pointet, A.-L.; Latreche, S.; Bergaya, S.; Benhamouda, N.; Tanchot, C.; et al. VEGF-A modulates expression of inhibitory checkpoints on CD8+ T cells in tumors. J. Exp. Med. 2015, 212, 139–148. [Google Scholar] [CrossRef]
  60. Kim, C.G.; Jang, M.; Kim, Y.; Leem, G.; Kim, K.H.; Lee, H.; Kim, T.-S.; Choi, S.J.; Kim, H.-D.; Han, J.W.; et al. VEGF-A drives TOX-dependent T cell exhaustion in anti-PD-1-resistant microsatellite stable colorectal cancers. Sci. Immunol. 2019, 4, eaay0555. [Google Scholar] [CrossRef]
  61. Ding, C.; Li, L.; Yang, T.; Fan, X.; Wu, G. Combined application of anti-VEGF and anti-EGFR attenuates the growth and angiogenesis of colorectal cancer mainly through suppressing AKT and ERK signaling in mice model. BMC Cancer 2016, 16, 791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Poindessous, V.; Ouaret, D.; El Ouadrani, K.; Battistella, A.; Mégalophonos, V.F.; Kamsu-Kom, N.; Petitprez, A.; Escargueil, A.E.; Boudou, P.; Dumont, S.; et al. EGFR- and VEGF(R)-targeted small molecules show synergistic activity in colorectal cancer models refractory to combinations of monoclonal antibodies. Clin. Cancer Res. 2011, 17, 6522–6530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Montagut, C.; Argilés, G.; Ciardiello, F.; Poulsen, T.T.; Dienstmann, R.; Kragh, M.; Kopetz, S.; Lindsted, T.; Ding, C.; Vidal, J.; et al. Efficacy of Sym004 in Patients With Metastatic Colorectal Cancer with Acquired Resistance to Anti-EGFR Therapy and Molecularly Selected by Circulating Tumor DNA Analyses: A Phase 2 Randomized Clinical Trial. JAMA Oncol. 2018, 4, e175245. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Lieu, C.H.; Beeram, M.; Harb, W.A.; Kearns, J.D.; Sloss, C.M.; Nering, R.; Adjei, A.A.; Wolf, B.B. Safety, pharmacology, and preliminary clinical activity of MM-151: An oligocolnal anti-EGFR theraputic in patients with cetuximab-resistant CRC and other refractory solid tumors. J. Clin. Oncol. 2015, 33, 647. [Google Scholar] [CrossRef]
  65. Elez, E.; Hendlisz, A.; Delaunoit, T.; Sastre, J.; Cervantes, A.; Varea, R.; Chao, G.; Wallin, J.; Tabernero, J. Phase II study of necitumumab plus modified FOLFOX6 as first-line treatment in patients with locally advanced or metastatic colorectal cancer. Br. J. Cancer 2016, 114, 372–380. [Google Scholar] [CrossRef] [Green Version]
  66. Yaeger, R.; Cercek, A.; O’Reilly, E.M.; Reidy, D.L.; Kemeny, N.; Wolinsky, T.; Capanu, M.; Gollub, M.J.; Rosen, N.; Berger, M.F.; et al. Pilot trial of combined BRAF and EGFR inhibition in BRAF-mutant metastatic colorectal cancer patients. Clin. Cancer Res. 2015, 21, 1313–1320. [Google Scholar] [CrossRef] [Green Version]
  67. Hong, D.S.; Morris, V.K.; El Osta, B.; Sorokin, A.V.; Janku, F.; Fu, S.; Overman, M.J.; Piha-Paul, S.; Subbiah, V.; Kee, B.; et al. Phase IB Study of Vemurafenib in Combination with Irinotecan and Cetuximab in Patients with Metastatic Colorectal Cancer with BRAFV600E Mutation. Cancer Discov. 2016, 6, 1352–1365. [Google Scholar] [CrossRef] [Green Version]
  68. Bowles, D.W.; Kochenderfer, M.; Cohn, A.; Sideris, L.; Nguyen, N.; Cline-Burkhardt, V.; Schnadig, I.; Choi, M.; Nabell, L.; Chaudhry, A.; et al. A Randomized, Phase II Trial of Cetuximab With or Without PX-866, an Irreversible Oral Phosphatidylinositol 3-Kinase Inhibitor, in Patients With Metastatic Colorectal Carcinoma. Clin. Colorectal Cancer 2016, 15, 337–344.e2. [Google Scholar] [CrossRef]
  69. van Geel, R.M.J.M.; Tabernero, J.; Elez, E.; Bendell, J.C.; Spreafico, A.; Schuler, M.; Yoshino, T.; Delord, J.-P.; Yamada, Y.; Lolkema, M.P.; et al. A Phase Ib Dose-Escalation Study of Encorafenib and Cetuximab with or without Alpelisib in Metastatic -Mutant Colorectal Cancer. Cancer Discov. 2017, 7, 610–619. [Google Scholar] [CrossRef] [Green Version]
  70. Reidy, D.L.; Vakiani, E.; Fakih, M.G.; Saif, M.W.; Hecht, J.R.; Goodman-Davis, N.; Hollywood, E.; Shia, J.; Schwartz, J.; Chandrawansa, K.; et al. Randomized, phase II study of the insulin-like growth factor-1 receptor inhibitor IMC-A12, with or without cetuximab, in patients with cetuximab- or panitumumab-refractory metastatic colorectal cancer. J. Clin. Oncol. 2010, 28, 4240–4246. [Google Scholar] [CrossRef] [Green Version]
  71. Sclafani, F.; Kim, T.Y.; Cunningham, D.; Kim, T.W.; Tabernero, J.; Schmoll, H.J.; Roh, J.K.; Kim, S.Y.; Park, Y.S.; Guren, T.K.; et al. A Randomized Phase II/III Study of Dalotuzumab in Combination With Cetuximab and Irinotecan in Chemorefractory, KRAS Wild-Type, Metastatic Colorectal Cancer. J. Natl. Cancer Inst. 2015, 107, djv258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Van Cutsem, E.; Eng, C.; Nowara, E.; Swieboda-Sadlej, A.; Tebbutt, N.C.; Mitchell, E.; Davidenko, I.; Stephenson, J.; Elez, E.; Prenen, H.; et al. Randomized phase Ib/II trial of rilotumumab or ganitumab with panitumumab versus panitumumab alone in patients with wild-type KRAS metastatic colorectal cancer. Clin. Cancer Res. 2014, 20, 4240–4250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Jacobs, S.A.; Lee, J.J.; George, T.J.; Wade, J.L.; Stella, P.J.; Wang, D.; Sama, A.R.; Piette, F.; Pogue-Geile, K.L.; Kim, R.S.; et al. Neratinib-Plus-Cetuximab in Quadruple-WT () Metastatic Colorectal Cancer Resistant to Cetuximab or Panitumumab: NSABP FC-7, A Phase Ib Study. Clin. Cancer Res. 2021, 27, 1612–1622. [Google Scholar] [CrossRef] [PubMed]
  74. Sartore-Bianchi, A.; Trusolino, L.; Martino, C.; Bencardino, K.; Lonardi, S.; Bergamo, F.; Zagonel, V.; Leone, F.; Depetris, I.; Martinelli, E.; et al. Dual-targeted therapy with trastuzumab and lapatinib in treatment-refractory, KRAS codon 12/13 wild-type, HER2-positive metastatic colorectal cancer (HERACLES): A proof-of-concept, multicentre, open-label, phase 2 trial. Lancet Oncol. 2016, 17, 738–746. [Google Scholar] [CrossRef]
  75. Meric-Bernstam, F.; Hurwitz, H.; Raghav, K.P.S.; McWilliams, R.R.; Fakih, M.; VanderWalde, A.; Swanton, C.; Kurzrock, R.; Burris, H.; Sweeney, C.; et al. Pertuzumab plus trastuzumab for HER2-amplified metastatic colorectal cancer (MyPathway): An updated report from a multicentre, open-label, phase 2a, multiple basket study. Lancet Oncol. 2019, 20, 518–530. [Google Scholar] [CrossRef]
  76. Meulendijks, D.; Jacob, W.; Martinez-Garcia, M.; Taus, A.; Lolkema, M.P.; Voest, E.E.; Langenberg, M.H.G.; Fleitas Kanonnikoff, T.; Cervantes, A.; De Jonge, M.J.; et al. First-in-Human Phase I Study of Lumretuzumab, a Glycoengineered Humanized Anti-HER3 Monoclonal Antibody, in Patients with Metastatic or Advanced HER3-Positive Solid Tumors. Clin. Cancer Res. 2016, 22, 877–885. [Google Scholar] [CrossRef] [Green Version]
  77. Juric, D.; Dienstmann, R.; Cervantes, A.; Hidalgo, M.; Messersmith, W.; Blumenschein, G.R.; Tabernero, J.; Roda, D.; Calles, A.; Jimeno, A.; et al. Safety and Pharmacokinetics/Pharmacodynamics of the First-in-Class Dual Action HER3/EGFR Antibody MEHD7945A in Locally Advanced or Metastatic Epithelial Tumors. Clin. Cancer Res. 2015, 21, 2462–2470. [Google Scholar] [CrossRef] [Green Version]
  78. Rimassa, L.; Bozzarelli, S.; Pietrantonio, F.; Cordio, S.; Lonardi, S.; Toppo, L.; Zaniboni, A.; Bordonaro, R.; Di Bartolomeo, M.; Tomasello, G.; et al. Phase II Study of Tivantinib and Cetuximab in Patients With KRAS Wild-type Metastatic Colorectal Cancer With Acquired Resistance to EGFR Inhibitors and Emergence of MET Overexpression: Lesson Learned for Future Trials With EGFR/MET Dual Inhibition. Clin. Colorectal Cancer 2019, 18, 125–132.e2. [Google Scholar] [CrossRef]
  79. Eng, C.; Bessudo, A.; Hart, L.L.; Severtsev, A.; Gladkov, O.; Müller, L.; Kopp, M.V.; Vladimirov, V.; Langdon, R.; Kotiv, B.; et al. A randomized, placebo-controlled, phase 1/2 study of tivantinib (ARQ 197) in combination with irinotecan and cetuximab in patients with metastatic colorectal cancer with wild-type KRAS who have received first-line systemic therapy. Int. J. Cancer 2016, 139, 177–186. [Google Scholar] [CrossRef] [Green Version]
  80. Delord, J.-P.; Argilés, G.; Fayette, J.; Wirth, L.; Kasper, S.; Siena, S.; Mesia, R.; Berardi, R.; Cervantes, A.; Dekervel, J.; et al. A phase 1b study of the MET inhibitor capmatinib combined with cetuximab in patients with MET-positive colorectal cancer who had progressed following anti-EGFR monoclonal antibody treatment. Investig. New Drugs 2020, 38, 1774–1783. [Google Scholar] [CrossRef]
  81. Clarke, J.M.; Blobe, G.C.; Strickler, J.H.; Uronis, H.E.; Zafar, S.Y.; Morse, M.; Dropkin, E.; Howard, L.; O’Neill, M.; Rushing, C.N.; et al. A phase Ib study of the combination regorafenib with PF-03446962 in patients with refractory metastatic colorectal cancer (REGAL-1 trial). Cancer Chemother. Pharmacol. 2019, 84, 909–917. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Hellmann, M.D.; Kim, T.W.; Lee, C.B.; Goh, B.C.; Miller, W.H.; Oh, D.Y.; Jamal, R.; Chee, C.E.; Chow, L.Q.M.; Gainor, J.F.; et al. Phase Ib study of atezolizumab combined with cobimetinib in patients with solid tumors. Ann. Oncol. 2019, 30, 1134–1142. [Google Scholar] [CrossRef] [PubMed]
  83. Eng, C.; Kim, T.W.; Bendell, J.; Argilés, G.; Tebbutt, N.C.; Di Bartolomeo, M.; Falcone, A.; Fakih, M.; Kozloff, M.; Segal, N.H.; et al. Atezolizumab with or without cobimetinib versus regorafenib in previously treated metastatic colorectal cancer (IMblaze370): A multicentre, open-label, phase 3, randomised, controlled trial. Lancet Oncol. 2019, 20, 849–861. [Google Scholar] [CrossRef]
  84. Hochster, H.S.; Bendell, J.C.; Cleary, J.M.; Foster, P.; Zhang, W.; He, X.; Hernandez, G.; Iizuka, K.; Eckhardt, S.G. Efficacy and safety of atezolizumab (atezo) and bevacizumab (bev) in a phase Ib study of microsatellite instability (MSI)-high metastatic colorectal cancer (mCRC). J. Clin. Oncol. 2017, 35, 673. [Google Scholar] [CrossRef]
  85. Mettu, N.B.; Twohy, E.; Ou, F.S.; Halfdanarson, T.R.; Lenz, H.J.; Breakstone, R.; Boland, P.M.; Crysler, O.; Wu, C.; Grothey, A.; et al. 533PD-BACCI: A phase II randomized, double-blind, multicenter, placebo-controlled study of capecitabine (C) bevacizumab (B) plus atezolizumab (A) or placebo (P) in refractory metastatic colorectal cancer (mCRC): An ACCRU network study. Ann. Oncol. 2019, 30, v203. [Google Scholar] [CrossRef]
  86. Subbiah, V.; Khawaja, M.R.; Hong, D.S.; Amini, B.; Yungfang, J.; Liu, H.; Johnson, A.; Schrock, A.B.; Ali, S.M.; Sun, J.X.; et al. First-in-human trial of multikinase VEGF inhibitor regorafenib and anti-EGFR antibody cetuximab in advanced cancer patients. JCI Insight 2017, 2, e90380. [Google Scholar] [CrossRef]
  87. Kciuk, M.; Marciniak, B.; Kontek, R. Irinotecan-Still an Important Player in Cancer Chemotherapy: A Comprehensive Overview. Int. J. Mol. Sci. 2020, 21, 4919. [Google Scholar] [CrossRef]
  88. Xu, Y.; Villalona-Calero, M.A. Irinotecan: Mechanisms of tumor resistance and novel strategies for modulating its activity. Ann. Oncol. 2002, 13, 1841–1851. [Google Scholar] [CrossRef]
  89. Shaojun, C.; Li, H.; Haixin, H.; Guisheng, L. Expression of Topoisomerase 1 and carboxylesterase 2 correlates with irinotecan treatment response in metastatic colorectal cancer. Cancer Biol. Ther. 2018, 19, 153–159. [Google Scholar] [CrossRef] [Green Version]
  90. Xu, G.; Zhang, W.; Ma, M.K.; McLeod, H.L. Human carboxylesterase 2 is commonly expressed in tumor tissue and is correlated with activation of irinotecan. Clin. Cancer Res. 2002, 8, 2605–2611. [Google Scholar]
  91. Uchino, J.; Takayama, K.; Harada, A.; Sone, T.; Harada, T.; Curiel, D.T.; Kuroki, M.; Nakanishi, Y. Tumor targeting carboxylesterase fused with anti-CEA scFv improve the anticancer effect with a less toxic dose of irinotecan. Cancer Gene Ther. 2008, 15, 94–100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Oosterhoff, D.; Overmeer, R.M.; de Graaf, M.; van der Meulen, I.H.; Giaccone, G.; van Beusechem, V.W.; Haisma, H.J.; Pinedo, H.M.; Gerritsen, W.R. Adenoviral vector-mediated expression of a gene encoding secreted, EpCAM-targeted carboxylesterase-2 sensitises colon cancer spheroids to CPT-11. Br. J. Cancer 2005, 92, 882–887. [Google Scholar] [CrossRef] [Green Version]
  93. Wierdl, M.; Morton, C.L.; Weeks, J.K.; Danks, M.K.; Harris, L.C.; Potter, P.M. Sensitization of human tumor cells to CPT-11 via adenoviral-mediated delivery of a rabbit liver carboxylesterase. Cancer Res. 2001, 61, 5078–5082. [Google Scholar] [PubMed]
  94. Gutova, M.; Najbauer, J.; Chen, M.Y.; Potter, P.M.; Kim, S.U.; Aboody, K.S. Therapeutic targeting of melanoma cells using neural stem cells expressing carboxylesterase, a CPT-11 activating enzyme. Curr. Stem Cell Res. Ther. 2010, 5, 273–276. [Google Scholar] [CrossRef] [PubMed]
  95. Yi, B.-R.; Kim, S.U.; Choi, K.-C. Co-treatment with therapeutic neural stem cells expressing carboxyl esterase and CPT-11 inhibit growth of primary and metastatic lung cancers in mice. Oncotarget 2014, 5, 12835–12848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Choi, S.S.; Yoon, K.; Choi, S.-A.; Yoon, S.-B.; Kim, S.U.; Lee, H.J. Tumor-specific gene therapy for pancreatic cancer using human neural stem cells encoding carboxylesterase. Oncotarget 2016, 7, 75319–75327. [Google Scholar] [CrossRef] [Green Version]
  97. Saif, M.W.; Syrigos, K.N.; Katirtzoglou, N.A. S-1: A promising new oral fluoropyrimidine derivative. Expert Opin. Investig. Drugs 2009, 18, 335–348. [Google Scholar] [CrossRef]
  98. Zhang, X.; Soong, R.; Wang, K.; Li, L.; Davie, J.R.; Guarcello, V.; Diasio, R.B. Suppression of DPYD expression in RKO cells via DNA methylation in the regulatory region of the DPYD promoter: A potentially important epigenetic mechanism regulating DPYD expression. Biochem. Cell Biol. 2007, 85, 337–346. [Google Scholar] [CrossRef]
  99. Ezzeldin, H.H.; Lee, A.M.; Mattison, L.K.; Diasio, R.B. Methylation of the DPYD promoter: An alternative mechanism for dihydropyrimidine dehydrogenase deficiency in cancer patients. Clin. Cancer Res. 2005, 11, 8699–8705. [Google Scholar] [CrossRef] [Green Version]
  100. Porter, D.J.; Chestnut, W.G.; Merrill, B.M.; Spector, T. Mechanism-based inactivation of dihydropyrimidine dehydrogenase by 5-ethynyluracil. J. Biol. Chem. 1992, 267, 5236–5242. [Google Scholar] [CrossRef]
  101. Spector, T.; Cao, S.; Rustum, Y.M.; Harrington, J.A.; Porter, D.J. Attenuation of the antitumor activity of 5-fluorouracil by (R)-5-fluoro-5,6-dihydrouracil. Cancer Res. 1995, 55, 1239–1241. [Google Scholar] [PubMed]
  102. Takechi, T.; Fujioka, A.; Matsushima, E.; Fukushima, M. Enhancement of the antitumour activity of 5-fluorouracil (5-FU) by inhibiting dihydropyrimidine dehydrogenase activity (DPD) using 5-chloro-2,4-dihydroxypyridine (CDHP) in human tumour cells. Eur. J. Cancer 2002, 38, 1271–1277. [Google Scholar] [CrossRef]
  103. Johnson, M.R.; Hageboutros, A.; Wang, K.; High, L.; Smith, J.B.; Diasio, R.B. Life-threatening toxicity in a dihydropyrimidine dehydrogenase-deficient patient after treatment with topical 5-fluorouracil. Clin Cancer Res. 1999, 5, 2006–2011. [Google Scholar]
  104. De Man, F.M.; Goey, A.K.L.; van Schaik, R.H.N.; Mathijssen, R.H.J.; Bins, S. Individualization of Irinotecan Treatment: A Review of Pharmacokinetics, Pharmacodynamics, and Pharmacogenetics. Clin. Pharmacokinet. 2018, 57, 1229–1254. [Google Scholar] [CrossRef] [Green Version]
  105. Van der Bol, J.M.; Mathijssen, R.H.J.; Loos, W.J.; Friberg, L.E.; van Schaik, R.H.N.; de Jonge, M.J.A.; Planting, A.S.T.; Verweij, J.; Sparreboom, A.; de Jong, F.A. Cigarette smoking and irinotecan treatment: Pharmacokinetic interaction and effects on neutropenia. J. Clin. Oncol. 2007, 25, 2719–2726. [Google Scholar] [CrossRef] [PubMed]
  106. Mathijssen, R.H.J.; Verweij, J.; de Bruijn, P.; Loos, W.J.; Sparreboom, A. Effects of St. John’s wort on irinotecan metabolism. J. Natl. Cancer Inst. 2002, 94, 1247–1249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Innocenti, F.; Undevia, S.D.; Ramírez, J.; Mani, S.; Schilsky, R.L.; Vogelzang, N.J.; Prado, M.; Ratain, M.J. A phase I trial of pharmacologic modulation of irinotecan with cyclosporine and phenobarbital. Clin. Pharmacol. Ther. 2004, 76, 490–502. [Google Scholar] [CrossRef]
  108. van der Bol, J.M.; Visser, T.J.; Loos, W.J.; de Jong, F.A.; Wiemer, E.A.C.; van Aken, M.O.; Planting, A.S.; Schellens, J.H.; Verweij, J.; Mathijssen, R.H.J. Effects of methimazole on the elimination of irinotecan. Cancer Chemother. Pharmacol. 2011, 67, 231–236. [Google Scholar] [CrossRef] [Green Version]
  109. Kehrer, D.F.; Sparreboom, A.; Verweij, J.; de Bruijn, P.; Nierop, C.A.; van de Schraaf, J.; Ruijgrok, E.J.; de Jonge, M.J. Modulation of irinotecan-induced diarrhea by cotreatment with neomycin in cancer patients. Clin. Cancer Res. 2001, 7, 1136–1141. [Google Scholar]
  110. Kong, R.; Liu, T.; Zhu, X.; Ahmad, S.; Williams, A.L.; Phan, A.T.; Zhao, H.; Scott, J.E.; Yeh, L.-A.; Wong, S.T.C. Old drug new use—Amoxapine and its metabolites as potent bacterial β-glucuronidase inhibitors for alleviating cancer drug toxicity. Clin. Cancer Res. 2014, 20, 3521–3530. [Google Scholar] [CrossRef] [Green Version]
  111. Gagnon, J.-F.; Bernard, O.; Villeneuve, L.; Têtu, B.; Guillemette, C. Irinotecan inactivation is modulated by epigenetic silencing of UGT1A1 in colon cancer. Clin. Cancer Res. 2006, 12, 1850–1858. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Bélanger, A.-S.; Tojcic, J.; Harvey, M.; Guillemette, C. Regulation of UGT1A1 and HNF1 transcription factor gene expression by DNA methylation in colon cancer cells. BMC Mol. Biol. 2010, 11, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Li, W.; Zhang, H.; Assaraf, Y.G.; Zhao, K.; Xu, X.; Xie, J.; Yang, D.-H.; Chen, Z.-S. Overcoming ABC transporter-mediated multidrug resistance: Molecular mechanisms and novel therapeutic drug strategies. Drug Resist. Updates 2016, 27, 14–29. [Google Scholar] [CrossRef] [PubMed]
  114. Giacomini, K.M.; Huang, S.-M.; Tweedie, D.J.; Benet, L.Z.; Brouwer, K.L.R.; Chu, X.; Dahlin, A.; Evers, R.; Fischer, V.; Hillgren, K.M.; et al. Membrane transporters in drug development. Nat. Rev. Drug Discov. 2010, 9, 215–236. [Google Scholar] [CrossRef]
  115. Lin, L.; Yee, S.W.; Kim, R.B.; Giacomini, K.M. SLC transporters as therapeutic targets: Emerging opportunities. Nat. Rev. Drug Discov. 2015, 14, 543–560. [Google Scholar] [CrossRef] [Green Version]
  116. Gottesman, M.M.; Ling, V. The molecular basis of multidrug resistance in cancer: The early years of P-glycoprotein research. FEBS Lett. 2006, 580, 998–1009. [Google Scholar] [CrossRef] [Green Version]
  117. Zhou, H.; Lin, C.; Zhang, Y.; Zhang, X.; Zhang, C.; Zhang, P.; Xie, X.; Ren, Z. miR-506 enhances the sensitivity of human colorectal cancer cells to oxaliplatin by suppressing MDR1/P-gp expression. Cell Prolif. 2017, 50, e12341. [Google Scholar] [CrossRef]
  118. Du, J.; He, Y.; Li, P.; Wu, W.; Chen, Y.; Ruan, H. IL-8 regulates the doxorubicin resistance of colorectal cancer cells via modulation of multidrug resistance 1 (MDR1). Cancer Chemother. Pharmacol. 2018, 81, 1111–1119. [Google Scholar] [CrossRef]
  119. Robey, R.W.; Shukla, S.; Steadman, K.; Obrzut, T.; Finley, E.M.; Ambudkar, S.V.; Bates, S.E. Inhibition of ABCG2-mediated transport by protein kinase inhibitors with a bisindolylmaleimide or indolocarbazole structure. Mol. Cancer Ther. 2007, 6, 1877–1885. [Google Scholar] [CrossRef] [Green Version]
  120. Liu, Z.; Qiu, M.; Tang, Q.-L.; Liu, M.; Lang, N.; Bi, F. Establishment and biological characteristics of oxaliplatin-resistant human colon cancer cell lines. Chin. J. Cancer 2010, 29, 661–667. [Google Scholar] [CrossRef] [Green Version]
  121. Herraez, E.; Gonzalez-Sanchez, E.; Vaquero, J.; Romero, M.R.; Serrano, M.A.; Marin, J.J.G.; Briz, O. Cisplatin-induced chemoresistance in colon cancer cells involves FXR-dependent and FXR-independent up-regulation of ABC proteins. Mol. Pharm. 2012, 9, 2565–2576. [Google Scholar] [CrossRef] [PubMed]
  122. Palmeira, A.; Sousa, E.; Vasconcelos, M.H.; Pinto, M.M. Three decades of P-gp inhibitors: Skimming through several generations and scaffolds. Curr. Med. Chem. 2012, 19, 1946–2025. [Google Scholar] [CrossRef] [PubMed]
  123. Wu, C.-P.; Ohnuma, S.; Ambudkar, S.V. Discovering natural product modulators to overcome multidrug resistance in cancer chemotherapy. Curr. Pharm. Biotechnol. 2011, 12, 609–620. [Google Scholar] [CrossRef] [PubMed]
  124. Wang, B.; Lu, F.-Y.; Shi, R.-H.; Feng, Y.-D.; Zhao, X.-D.; Lu, Z.-P.; Xiao, L.; Zhou, G.-Q.; Qiu, J.-M.; Cheng, C.-E. MiR-26b regulates 5-FU-resistance in human colorectal cancer via down-regulation of Pgp. Am. J. Cancer Res. 2018, 8, 2518–2527. [Google Scholar] [PubMed]
  125. Gao, R.; Fang, C.; Xu, J.; Tan, H.; Li, P.; Ma, L. LncRNA CACS15 contributes to oxaliplatin resistance in colorectal cancer by positively regulating ABCC1 through sponging miR-145. Arch. Biochem. Biophys. 2019, 663, 183–191. [Google Scholar] [CrossRef]
  126. Fan, H.; Zhu, J.-H.; Yao, X.-Q. Knockdown of long non-coding RNA PVT1 reverses multidrug resistance in colorectal cancer cells. Mol. Med. Rep. 2018, 17, 8309–8315. [Google Scholar] [CrossRef] [Green Version]
  127. Vodenkova, S.; Buchler, T.; Cervena, K.; Veskrnova, V.; Vodicka, P.; Vymetalkova, V. 5-fluorouracil and other fluoropyrimidines in colorectal cancer: Past, present and future. Pharmacol. Ther. 2020, 206, 107447. [Google Scholar] [CrossRef]
  128. Longley, D.B.; Johnston, P.G. Molecular mechanisms of drug resistance. J. Pathol. 2005, 205, 275–292. [Google Scholar] [CrossRef]
  129. Grogan, L.; Sotos, G.A.; Allegra, C.J. Leucovorin modulation of fluorouracil. Oncology 1993, 7, 63–72. [Google Scholar]
  130. Tan, K.B.; Mattern, M.R.; Eng, W.K.; McCabe, F.L.; Johnson, R.K. Nonproductive rearrangement of DNA topoisomerase I and II genes: Correlation with resistance to topoisomerase inhibitors. J. Natl. Cancer Inst. 1989, 81, 1732–1735. [Google Scholar] [CrossRef]
  131. Li, X.G.; Haluska, P.; Hsiang, Y.H.; Bharti, A.; Kufe, D.W.; Rubin, E.H. Identification of topoisomerase I mutations affecting both DNA cleavage and interaction with camptothecin. Ann. N. Y. Acad. Sci. 1996, 803, 111–127. [Google Scholar] [CrossRef] [PubMed]
  132. Fornasier, G.; Francescon, S.; Baldo, P. An Update of Efficacy and Safety of Cetuximab in Metastatic Colorectal Cancer: A Narrative Review. Adv. Ther. 2018, 35, 1497–1509. [Google Scholar] [CrossRef] [PubMed]
  133. Loupakis, F.; Cremolini, C.; Fioravanti, A.; Orlandi, P.; Salvatore, L.; Masi, G.; Schirripa, M.; Di Desidero, T.; Antoniotti, C.; Canu, B.; et al. EGFR ligands as pharmacodynamic biomarkers in metastatic colorectal cancer patients treated with cetuximab and irinotecan. Target Oncol. 2014, 9, 205–214. [Google Scholar] [CrossRef] [PubMed]
  134. Xian, C.J. Roles of epidermal growth factor family in the regulation of postnatal somatic growth. Endocr Rev 2007, 28, 284–296. [Google Scholar] [CrossRef] [Green Version]
  135. Khambata-Ford, S.; Garrett, C.R.; Meropol, N.J.; Basik, M.; Harbison, C.T.; Wu, S.; Wong, T.W.; Huang, X.; Takimoto, C.H.; Godwin, A.K.; et al. Expression of epiregulin and amphiregulin and K-ras mutation status predict disease control in metastatic colorectal cancer patients treated with cetuximab. J. Clin. Oncol. 2007, 25, 3230–3237. [Google Scholar] [CrossRef]
  136. Jacobs, B.; De Roock, W.; Piessevaux, H.; Van Oirbeek, R.; Biesmans, B.; De Schutter, J.; Fieuws, S.; Vandesompele, J.; Peeters, M.; Van Laethem, J.-L.; et al. Amphiregulin and epiregulin mRNA expression in primary tumors predicts outcome in metastatic colorectal cancer treated with cetuximab. J. Clin. Oncol. 2009, 27, 5068–5074. [Google Scholar] [CrossRef]
  137. Seligmann, J.F.; Elliott, F.; Richman, S.D.; Jacobs, B.; Hemmings, G.; Brown, S.; Barrett, J.H.; Tejpar, S.; Quirke, P.; Seymour, M.T. Combined Epiregulin and Amphiregulin Expression Levels as a Predictive Biomarker for Panitumumab Therapy Benefit or Lack of Benefit in Patients With RAS Wild-Type Advanced Colorectal Cancer. JAMA Oncol. 2016, 2, 633–642. [Google Scholar] [CrossRef] [Green Version]
  138. Fukuoka, M.; Wu, Y.-L.; Thongprasert, S.; Sunpaweravong, P.; Leong, S.-S.; Sriuranpong, V.; Chao, T.-Y.; Nakagawa, K.; Chu, D.-T.; Saijo, N.; et al. Biomarker analyses and final overall survival results from a phase III, randomized, open-label, first-line study of gefitinib versus carboplatin/paclitaxel in clinically selected patients with advanced non-small-cell lung cancer in Asia (IPASS). J. Clin. Oncol. 2011, 29, 2866–2874. [Google Scholar] [CrossRef]
  139. Chung, K.Y.; Shia, J.; Kemeny, N.E.; Shah, M.; Schwartz, G.K.; Tse, A.; Hamilton, A.; Pan, D.; Schrag, D.; Schwartz, L.; et al. Cetuximab shows activity in colorectal cancer patients with tumors that do not express the epidermal growth factor receptor by immunohistochemistry. J. Clin. Oncol. 2005, 23, 1803–1810. [Google Scholar] [CrossRef]
  140. Hecht, J.R.; Mitchell, E.; Neubauer, M.A.; Burris, H.A.; Swanson, P.; Lopez, T.; Buchanan, G.; Reiner, M.; Gansert, J.; Berlin, J. Lack of correlation between epidermal growth factor receptor status and response to Panitumumab monotherapy in metastatic colorectal cancer. Clin. Cancer Res. 2010, 16, 2205–2213. [Google Scholar] [CrossRef] [Green Version]
  141. Montagut, C.; Dalmases, A.; Bellosillo, B.; Crespo, M.; Pairet, S.; Iglesias, M.; Salido, M.; Gallen, M.; Marsters, S.; Tsai, S.P.; et al. Identification of a mutation in the extracellular domain of the Epidermal Growth Factor Receptor conferring cetuximab resistance in colorectal cancer. Nat. Med. 2012, 18, 221–223. [Google Scholar] [CrossRef] [PubMed]
  142. Arena, S.; Bellosillo, B.; Siravegna, G.; Martínez, A.; Cañadas, I.; Lazzari, L.; Ferruz, N.; Russo, M.; Misale, S.; González, I.; et al. Emergence of Multiple EGFR Extracellular Mutations during Cetuximab Treatment in Colorectal Cancer. Clin. Cancer Res. 2015, 21, 2157–2166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Van Emburgh, B.O.; Arena, S.; Siravegna, G.; Lazzari, L.; Crisafulli, G.; Corti, G.; Mussolin, B.; Baldi, F.; Buscarino, M.; Bartolini, A.; et al. Acquired RAS or EGFR mutations and duration of response to EGFR blockade in colorectal cancer. Nat. Commun. 2016, 7, 13665. [Google Scholar] [CrossRef]
  144. Bertotti, A.; Papp, E.; Jones, S.; Adleff, V.; Anagnostou, V.; Lupo, B.; Sausen, M.; Phallen, J.; Hruban, C.A.; Tokheim, C.; et al. The genomic landscape of response to EGFR blockade in colorectal cancer. Nature 2015, 526, 263–267. [Google Scholar] [CrossRef] [PubMed]
  145. Liao, H.-W.; Hsu, J.-M.; Xia, W.; Wang, H.-L.; Wang, Y.-N.; Chang, W.-C.; Arold, S.T.; Chou, C.-K.; Tsou, P.-H.; Yamaguchi, H.; et al. PRMT1-mediated methylation of the EGF receptor regulates signaling and cetuximab response. J. Clin. Investig. 2015, 125, 4529–4543. [Google Scholar] [CrossRef] [Green Version]
  146. Hochster, H.S.; Hart, L.L.; Ramanathan, R.K.; Childs, B.H.; Hainsworth, J.D.; Cohn, A.L.; Wong, L.; Fehrenbacher, L.; Abubakr, Y.; Saif, M.W.; et al. Safety and efficacy of oxaliplatin and fluoropyrimidine regimens with or without bevacizumab as first-line treatment of metastatic colorectal cancer: Results of the TREE Study. J. Clin. Oncol. 2008, 26, 3523–3529. [Google Scholar] [CrossRef]
  147. Sánchez-Martín, F.J.; Bellosillo, B.; Gelabert-Baldrich, M.; Dalmases, A.; Cañadas, I.; Vidal, J.; Martinez, A.; Argilés, G.; Siravegna, G.; Arena, S.; et al. The First-in-class Anti-EGFR Antibody Mixture Sym004 Overcomes Cetuximab Resistance Mediated by EGFR Extracellular Domain Mutations in Colorectal Cancer. Clin. Cancer Res. 2016, 22, 3260–3267. [Google Scholar] [CrossRef] [Green Version]
  148. El-Deiry, W.S.; Goldberg, R.M.; Lenz, H.-J.; Shields, A.F.; Gibney, G.T.; Tan, A.R.; Brown, J.; Eisenberg, B.; Heath, E.I.; Phuphanich, S.; et al. The current state of molecular testing in the treatment of patients with solid tumors, 2019. CA Cancer J. Clin. 2019, 69, 305–343. [Google Scholar] [CrossRef]
  149. Van Cutsem, E.; Köhne, C.-H.; Láng, I.; Folprecht, G.; Nowacki, M.P.; Cascinu, S.; Shchepotin, I.; Maurel, J.; Cunningham, D.; Tejpar, S.; et al. Cetuximab plus irinotecan, fluorouracil, and leucovorin as first-line treatment for metastatic colorectal cancer: Updated analysis of overall survival according to tumor KRAS and BRAF mutation status. J. Clin. Oncol. 2011, 29, 2011–2019. [Google Scholar] [CrossRef] [Green Version]
  150. Van Cutsem, E.; Köhne, C.-H.; Hitre, E.; Zaluski, J.; Chang Chien, C.-R.; Makhson, A.; D’Haens, G.; Pintér, T.; Lim, R.; Bodoky, G.; et al. Cetuximab and chemotherapy as initial treatment for metastatic colorectal cancer. N. Engl. J. Med. 2009, 360, 1408–1417. [Google Scholar] [CrossRef] [Green Version]
  151. Peeters, M.; Price, T.J.; Cervantes, A.; Sobrero, A.F.; Ducreux, M.; Hotko, Y.; André, T.; Chan, E.; Lordick, F.; Punt, C.J.A.; et al. Randomized phase III study of panitumumab with fluorouracil, leucovorin, and irinotecan (FOLFIRI) compared with FOLFIRI alone as second-line treatment in patients with metastatic colorectal cancer. J. Clin. Oncol. 2010, 28, 4706–4713. [Google Scholar] [CrossRef] [PubMed]
  152. Ciardiello, F.; Normanno, N.; Maiello, E.; Martinelli, E.; Troiani, T.; Pisconti, S.; Giuliani, F.; Barone, C.; Cartenì, G.; Rachiglio, A.M.; et al. Clinical activity of FOLFIRI plus cetuximab according to extended gene mutation status by next-generation sequencing: Findings from the CAPRI-GOIM trial. Ann. Oncol. 2014, 25, 1756–1761. [Google Scholar] [CrossRef] [PubMed]
  153. Bokemeyer, C.; Köhne, C.H.; Ciardiello, F.; Lenz, H.J.; Heinemann, V.; Klinkhardt, U.; Beier, F.; Duecker, K.; van Krieken, J.H.; Tejpar, S. FOLFOX4 plus cetuximab treatment and RAS mutations in colorectal cancer. Eur. J. Cancer 2015, 51, 1243–1252. [Google Scholar] [CrossRef] [PubMed]
  154. Blons, H.; Emile, J.F.; Le Malicot, K.; Julié, C.; Zaanan, A.; Tabernero, J.; Mini, E.; Folprecht, G.; Van Laethem, J.L.; Thaler, J.; et al. Prognostic value of KRAS mutations in stage III colon cancer: Post hoc analysis of the PETACC8 phase III trial dataset. Ann. Oncol. 2014, 25, 2378–2385. [Google Scholar] [CrossRef] [PubMed]
  155. Ye, H.; Liu, Y.; Wu, K.; Luo, H.; Cui, L. AMPK activation overcomes anti-EGFR antibody resistance induced by KRAS mutation in colorectal cancer. Cell Commun. Signal. 2020, 18, 115. [Google Scholar] [CrossRef]
  156. Dunn, E.F.; Iida, M.; Myers, R.A.; Campbell, D.A.; Hintz, K.A.; Armstrong, E.A.; Li, C.; Wheeler, D.L. Dasatinib sensitizes KRAS mutant colorectal tumors to cetuximab. Oncogene 2011, 30, 561–574. [Google Scholar] [CrossRef] [Green Version]
  157. Skoulidis, F.; Li, B.T.; Dy, G.K.; Price, T.J.; Falchook, G.S.; Wolf, J.; Italiano, A.; Schuler, M.; Borghaei, H.; Barlesi, F.; et al. Sotorasib for Lung Cancers with p.G12C Mutation. N. Engl. J. Med. 2021, 384, 2371–2381. [Google Scholar] [CrossRef]
  158. Shinozaki, E.; Yoshino, T.; Yamazaki, K.; Muro, K.; Yamaguchi, K.; Nishina, T.; Yuki, S.; Shitara, K.; Bando, H.; Mimaki, S.; et al. Clinical significance of BRAF non-V600E mutations on the therapeutic effects of anti-EGFR monoclonal antibody treatment in patients with pretreated metastatic colorectal cancer: The Biomarker Research for anti-EGFR monoclonal Antibodies by Comprehensive Cancer genomics (BREAC) study. Br. J. Cancer 2017, 117, 1450–1458. [Google Scholar] [CrossRef]
  159. Prahallad, A.; Sun, C.; Huang, S.; Di Nicolantonio, F.; Salazar, R.; Zecchin, D.; Beijersbergen, R.L.; Bardelli, A.; Bernards, R. Unresponsiveness of colon cancer to BRAF(V600E) inhibition through feedback activation of EGFR. Nature 2012, 483, 100–103. [Google Scholar] [CrossRef] [Green Version]
  160. Rajagopalan, H.; Bardelli, A.; Lengauer, C.; Kinzler, K.W.; Vogelstein, B.; Velculescu, V.E. Tumorigenesis: RAF/RAS oncogenes and mismatch-repair status. Nature 2002, 418, 934. [Google Scholar] [CrossRef]
  161. Hyman, D.M.; Puzanov, I.; Subbiah, V.; Faris, J.E.; Chau, I.; Blay, J.-Y.; Wolf, J.; Raje, N.S.; Diamond, E.L.; Hollebecque, A.; et al. Vemurafenib in Multiple Nonmelanoma Cancers with BRAF V600 Mutations. N. Engl. J. Med. 2015, 373, 726–736. [Google Scholar] [CrossRef] [PubMed]
  162. Tabernero, J.; Grothey, A.; Van Cutsem, E.; Yaeger, R.; Wasan, H.; Yoshino, T.; Desai, J.; Ciardiello, F.; Loupakis, F.; Hong, Y.S.; et al. Encorafenib Plus Cetuximab as a New Standard of Care for Previously Treated V600E-Mutant Metastatic Colorectal Cancer: Updated Survival Results and Subgroup Analyses from the BEACON Study. J. Clin. Oncol. 2021, 39, 273–284. [Google Scholar] [CrossRef] [PubMed]
  163. Sun, C.; Hobor, S.; Bertotti, A.; Zecchin, D.; Huang, S.; Galimi, F.; Cottino, F.; Prahallad, A.; Grernrum, W.; Tzani, A.; et al. Intrinsic resistance to MEK inhibition in KRAS mutant lung and colon cancer through transcriptional induction of ERBB3. Cell Rep. 2014, 7, 86–93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Li, Q.-H.; Wang, Y.-Z.; Tu, J.; Liu, C.-W.; Yuan, Y.-J.; Lin, R.; He, W.-L.; Cai, S.-R.; He, Y.-L.; Ye, J.-N. Anti-EGFR therapy in metastatic colorectal cancer: Mechanisms and potential regimens of drug resistance. Gastroenterol. Rep. 2020, 8, 179–191. [Google Scholar] [CrossRef] [PubMed]
  165. Sartore-Bianchi, A.; Martini, M.; Molinari, F.; Veronese, S.; Nichelatti, M.; Artale, S.; Di Nicolantonio, F.; Saletti, P.; De Dosso, S.; Mazzucchelli, L.; et al. PIK3CA mutations in colorectal cancer are associated with clinical resistance to EGFR-targeted monoclonal antibodies. Cancer Res. 2009, 69, 1851–1857. [Google Scholar] [CrossRef] [Green Version]
  166. Samuels, Y.; Diaz, L.A.; Schmidt-Kittler, O.; Cummins, J.M.; Delong, L.; Cheong, I.; Rago, C.; Huso, D.L.; Lengauer, C.; Kinzler, K.W.; et al. Mutant PIK3CA promotes cell growth and invasion of human cancer cells. Cancer Cell 2005, 7, 561–573. [Google Scholar] [CrossRef] [Green Version]
  167. De Roock, W.; Claes, B.; Bernasconi, D.; De Schutter, J.; Biesmans, B.; Fountzilas, G.; Kalogeras, K.T.; Kotoula, V.; Papamichael, D.; Laurent-Puig, P.; et al. Effects of KRAS, BRAF, NRAS, and PIK3CA mutations on the efficacy of cetuximab plus chemotherapy in chemotherapy-refractory metastatic colorectal cancer: A retrospective consortium analysis. Lancet Oncol. 2010, 11, 753–762. [Google Scholar] [CrossRef]
  168. Mao, C.; Yang, Z.Y.; Hu, X.F.; Chen, Q.; Tang, J.L. PIK3CA exon 20 mutations as a potential biomarker for resistance to anti-EGFR monoclonal antibodies in KRAS wild-type metastatic colorectal cancer: A systematic review and meta-analysis. Ann. Oncol. 2012, 23, 1518–1525. [Google Scholar] [CrossRef]
  169. Malek, M.; Kielkowska, A.; Chessa, T.; Anderson, K.E.; Barneda, D.; Pir, P.; Nakanishi, H.; Eguchi, S.; Koizumi, A.; Sasaki, J.; et al. PTEN Regulates PI(3,4)P Signaling Downstream of Class I PI3K. Mol. Cell 2017, 68, 566–580.e10. [Google Scholar] [CrossRef] [Green Version]
  170. Jhawer, M.; Goel, S.; Wilson, A.J.; Montagna, C.; Ling, Y.-H.; Byun, D.-S.; Nasser, S.; Arango, D.; Shin, J.; Klampfer, L.; et al. PIK3CA mutation/PTEN expression status predicts response of colon cancer cells to the epidermal growth factor receptor inhibitor cetuximab. Cancer Res. 2008, 68, 1953–1961. [Google Scholar] [CrossRef] [Green Version]
  171. Loupakis, F.; Pollina, L.; Stasi, I.; Ruzzo, A.; Scartozzi, M.; Santini, D.; Masi, G.; Graziano, F.; Cremolini, C.; Rulli, E.; et al. PTEN expression and KRAS mutations on primary tumors and metastases in the prediction of benefit from cetuximab plus irinotecan for patients with metastatic colorectal cancer. J. Clin. Oncol. 2009, 27, 2622–2629. [Google Scholar] [CrossRef] [PubMed]
  172. Scartozzi, M.; Giampieri, R.; Maccaroni, E.; Mandolesi, A.; Giustini, L.; Silva, R.; Zaniboni, A.; Biscotti, T.; Biagetti, S.; Galizia, E.; et al. Analysis of HER-3, insulin growth factor-1, nuclear factor-kB and epidermal growth factor receptor gene copy number in the prediction of clinical outcome for K-RAS wild-type colorectal cancer patients receiving irinotecan-cetuximab. Ann. Oncol. 2012, 23, 1706–1712. [Google Scholar] [CrossRef] [PubMed]
  173. Engelman, J.A.; Zejnullahu, K.; Mitsudomi, T.; Song, Y.; Hyland, C.; Park, J.O.; Lindeman, N.; Gale, C.-M.; Zhao, X.; Christensen, J.; et al. MET amplification leads to gefitinib resistance in lung cancer by activating ERBB3 signaling. Science 2007, 316, 1039–1043. [Google Scholar] [CrossRef] [PubMed]
  174. Kasprzak, A. Insulin-Like Growth Factor 1 (IGF-1) Signaling in Glucose Metabolism in Colorectal Cancer. Int. J. Mol. Sci. 2021, 22, 6434. [Google Scholar] [CrossRef]
  175. Scartozzi, M.; Mandolesi, A.; Giampieri, R.; Pierantoni, C.; Loupakis, F.; Zaniboni, A.; Galizia, E.; Giustini, L.; Silva, R.R.; Bisonni, R.; et al. Insulin-like growth factor 1 expression correlates with clinical outcome in K-RAS wild type colorectal cancer patients treated with cetuximab and irinotecan. Int. J. Cancer 2010, 127, 1941–1947. [Google Scholar] [CrossRef]
  176. Zanella, E.R.; Galimi, F.; Sassi, F.; Migliardi, G.; Cottino, F.; Leto, S.M.; Lupo, B.; Erriquez, J.; Isella, C.; Comoglio, P.M.; et al. IGF2 is an actionable target that identifies a distinct subpopulation of colorectal cancer patients with marginal response to anti-EGFR therapies. Sci. Transl. Med. 2015, 7, 272ra12. [Google Scholar] [CrossRef]
  177. Comprehensive molecular characterization of human colon and rectal cancer. Nature 2012, 487, 330–337. [CrossRef] [Green Version]
  178. Yonesaka, K.; Zejnullahu, K.; Okamoto, I.; Satoh, T.; Cappuzzo, F.; Souglakos, J.; Ercan, D.; Rogers, A.; Roncalli, M.; Takeda, M.; et al. Activation of ERBB2 signaling causes resistance to the EGFR-directed therapeutic antibody cetuximab. Sci. Transl. Med. 2011, 3, 99ra86. [Google Scholar] [CrossRef] [Green Version]
  179. Kavuri, S.M.; Jain, N.; Galimi, F.; Cottino, F.; Leto, S.M.; Migliardi, G.; Searleman, A.C.; Shen, W.; Monsey, J.; Trusolino, L.; et al. HER2 activating mutations are targets for colorectal cancer treatment. Cancer Discov. 2015, 5, 832–841. [Google Scholar] [CrossRef] [Green Version]
  180. Jaiswal, B.S.; Kljavin, N.M.; Stawiski, E.W.; Chan, E.; Parikh, C.; Durinck, S.; Chaudhuri, S.; Pujara, K.; Guillory, J.; Edgar, K.A.; et al. Oncogenic ERBB3 mutations in human cancers. Cancer Cell 2013, 23, 603–617. [Google Scholar] [CrossRef] [Green Version]
  181. Loree, J.M.; Bailey, A.M.; Johnson, A.M.; Yu, Y.; Wu, W.; Bristow, C.A.; Davis, J.S.; Shaw, K.R.; Broaddus, R.; Banks, K.C.; et al. Molecular Landscape of ERBB2/ERBB3 Mutated Colorectal Cancer. J. Natl. Cancer Inst. 2018, 110, 1409–1417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Seligmann, J.F.; Hatch, A.J.; Richman, S.D.; Elliott, F.; Jacobs, B.; Brown, S.; Hurwitz, H.; Barrett, J.H.; Quirke, P.; Nixon, A.B.; et al. Association of Tumor HER3 Messenger RNA Expression With Panitumumab Efficacy in Advanced Colorectal Cancer. JAMA Oncol. 2018, 4, 564–568. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Jo, M.; Stolz, D.B.; Esplen, J.E.; Dorko, K.; Michalopoulos, G.K.; Strom, S.C. Cross-talk between epidermal growth factor receptor and c-Met signal pathways in transformed cells. J. Biol. Chem. 2000, 275, 8806–8811. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Bardelli, A.; Corso, S.; Bertotti, A.; Hobor, S.; Valtorta, E.; Siravegna, G.; Sartore-Bianchi, A.; Scala, E.; Cassingena, A.; Zecchin, D.; et al. Amplification of the MET receptor drives resistance to anti-EGFR therapies in colorectal cancer. Cancer Discov. 2013, 3, 658–673. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Pietrantonio, F.; Vernieri, C.; Siravegna, G.; Mennitto, A.; Berenato, R.; Perrone, F.; Gloghini, A.; Tamborini, E.; Lonardi, S.; Morano, F.; et al. Heterogeneity of Acquired Resistance to Anti-EGFR Monoclonal Antibodies in Patients with Metastatic Colorectal Cancer. Clin. Cancer Res. 2017, 23, 2414–2422. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Pietrantonio, F.; Oddo, D.; Gloghini, A.; Valtorta, E.; Berenato, R.; Barault, L.; Caporale, M.; Busico, A.; Morano, F.; Gualeni, A.V.; et al. MET-Driven Resistance to Dual EGFR and BRAF Blockade May Be Overcome by Switching from EGFR to MET Inhibition in BRAF-Mutated Colorectal Cancer. Cancer Discov. 2016, 6, 963–971. [Google Scholar] [CrossRef] [Green Version]
  187. Troiani, T.; Martinelli, E.; Napolitano, S.; Vitagliano, D.; Ciuffreda, L.P.; Costantino, S.; Morgillo, F.; Capasso, A.; Sforza, V.; Nappi, A.; et al. Increased TGF-α as a mechanism of acquired resistance to the anti-EGFR inhibitor cetuximab through EGFR-MET interaction and activation of MET signaling in colon cancer cells. Clin. Cancer Res. 2013, 19, 6751–6765. [Google Scholar] [CrossRef] [Green Version]
  188. Adjei, A.A.; Schwartz, B.; Garmey, E. Early clinical development of ARQ 197, a selective, non-ATP-competitive inhibitor targeting MET tyrosine kinase for the treatment of advanced cancers. Oncologist 2011, 16, 788–799. [Google Scholar] [CrossRef] [Green Version]
  189. Clevers, H. Wnt/beta-catenin signaling in development and disease. Cell 2006, 127, 469–480. [Google Scholar] [CrossRef] [Green Version]
  190. Bugter, J.M.; Fenderico, N.; Maurice, M.M. Mutations and mechanisms of WNT pathway tumour suppressors in cancer. Nat. Rev. Cancer 2021, 21, 5–21. [Google Scholar] [CrossRef]
  191. Yuan, S.; Tao, F.; Zhang, X.; Zhang, Y.; Sun, X.; Wu, D. Role of Wnt/-Catenin Signaling in the Chemoresistance Modulation of Colorectal Cancer. BioMed Res. Int. 2020, 2020, 9390878. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Hu, T.; Li, C. Convergence between Wnt-β-catenin and EGFR signaling in cancer. Mol. Cancer 2010, 9, 236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Grainger, S.; Nguyen, N.; Richter, J.; Setayesh, J.; Lonquich, B.; Oon, C.H.; Wozniak, J.M.; Barahona, R.; Kamei, C.N.; Houston, J.; et al. EGFR is required for Wnt9a-Fzd9b signalling specificity in haematopoietic stem cells. Nat. Cell Biol. 2019, 21, 721–730. [Google Scholar] [CrossRef] [PubMed]
  194. Lu, Y.; Zhao, X.; Liu, Q.; Li, C.; Graves-Deal, R.; Cao, Z.; Singh, B.; Franklin, J.L.; Wang, J.; Hu, H.; et al. lncRNA MIR100HG-derived miR-100 and miR-125b mediate cetuximab resistance via Wnt/β-catenin signaling. Nat. Med. 2017, 23, 1331–1341. [Google Scholar] [CrossRef] [PubMed]
  195. Sharma, P.; Siddiqui, B.A.; Anandhan, S.; Yadav, S.S.; Subudhi, S.K.; Gao, J.; Goswami, S.; Allison, J.P. The Next Decade of Immune Checkpoint Therapy. Cancer Discov. 2021, 11, 838–857. [Google Scholar] [CrossRef]
  196. Grasso, C.S.; Giannakis, M.; Wells, D.K.; Hamada, T.; Mu, X.J.; Quist, M.; Nowak, J.A.; Nishihara, R.; Qian, Z.R.; Inamura, K.; et al. Genetic Mechanisms of Immune Evasion in Colorectal Cancer. Cancer Discov. 2018, 8, 730–749. [Google Scholar] [CrossRef] [Green Version]
  197. Polakis, P. Wnt signaling in cancer. Cold Spring Harb. Perspect. Biol. 2012, 4, a008052. [Google Scholar] [CrossRef] [Green Version]
  198. Xiao, Q.; Wu, J.; Wang, W.-J.; Chen, S.; Zheng, Y.; Yu, X.; Meeth, K.; Sahraei, M.; Bothwell, A.L.M.; Chen, L.; et al. DKK2 imparts tumor immunity evasion through β-catenin-independent suppression of cytotoxic immune-cell activation. Nat. Med. 2018, 24, 262–270. [Google Scholar] [CrossRef]
  199. Chen, D.S.; Mellman, I. Oncology meets immunology: The cancer-immunity cycle. Immunity 2013, 39, 1–10. [Google Scholar] [CrossRef] [Green Version]
  200. van Loosdregt, J.; Fleskens, V.; Tiemessen, M.M.; Mokry, M.; van Boxtel, R.; Meerding, J.; Pals, C.E.G.M.; Kurek, D.; Baert, M.R.M.; Delemarre, E.M.; et al. Canonical Wnt signaling negatively modulates regulatory T cell function. Immunity 2013, 39, 298–310. [Google Scholar] [CrossRef] [Green Version]
  201. Xin, P.; Xu, X.; Deng, C.; Liu, S.; Wang, Y.; Zhou, X.; Ma, H.; Wei, D.; Sun, S. The role of JAK/STAT signaling pathway and its inhibitors in diseases. Int. Immunopharmacol. 2020, 80, 106210. [Google Scholar] [CrossRef] [PubMed]
  202. Ma, X.-T.; Wang, S.; Ye, Y.-J.; Du, R.-Y.; Cui, Z.-R.; Somsouk, M. Constitutive activation of Stat3 signaling pathway in human colorectal carcinoma. World J. Gastroenterol. 2004, 10, 1569–1573. [Google Scholar] [CrossRef] [PubMed]
  203. Spitzner, M.; Roesler, B.; Bielfeld, C.; Emons, G.; Gaedcke, J.; Wolff, H.A.; Rave-Fränk, M.; Kramer, F.; Beissbarth, T.; Kitz, J.; et al. STAT3 inhibition sensitizes colorectal cancer to chemoradiotherapy in vitro and in vivo. Int. J. Cancer 2014, 134, 997–1007. [Google Scholar] [CrossRef] [PubMed]
  204. Li, Q.; Zhang, D.; Chen, X.; He, L.; Li, T.; Xu, X.; Li, M. Nuclear PKM2 contributes to gefitinib resistance via upregulation of STAT3 activation in colorectal cancer. Sci. Rep. 2015, 5, 16082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Burke, J.D.; Young, H.A. IFN-γ: A cytokine at the right time, is in the right place. Semin. Immunol. 2019, 43, 101280. [Google Scholar] [CrossRef] [PubMed]
  206. Zhou, F. Molecular mechanisms of IFN-gamma to up-regulate MHC class I antigen processing and presentation. Int. Rev. Immunol. 2009, 28, 239–260. [Google Scholar] [CrossRef]
  207. Platanias, L.C. Mechanisms of type-I- and type-II-interferon-mediated signalling. Nat. Rev. Immunol. 2005, 5, 375–386. [Google Scholar] [CrossRef]
  208. Kursunel, M.A.; Esendagli, G. The untold story of IFN-γ in cancer biology. Cytokine Growth Factor Rev. 2016, 31, 73–81. [Google Scholar] [CrossRef]
  209. Ikeda, H.; Old, L.J.; Schreiber, R.D. The roles of IFN gamma in protection against tumor development and cancer immunoediting. Cytokine Growth Factor Rev. 2002, 13, 95–109. [Google Scholar] [CrossRef]
  210. Shin, D.S.; Zaretsky, J.M.; Escuin-Ordinas, H.; Garcia-Diaz, A.; Hu-Lieskovan, S.; Kalbasi, A.; Grasso, C.S.; Hugo, W.; Sandoval, S.; Torrejon, D.Y.; et al. Primary Resistance to PD-1 Blockade Mediated by JAK1/2 Mutations. Cancer Discov. 2017, 7, 188–201. [Google Scholar] [CrossRef] [Green Version]
  211. Murthy, A.; Gerber, S.A.; Koch, C.J.; Lord, E.M. Intratumoral Hypoxia Reduces IFN-γ-Mediated Immunity and MHC Class I Induction in a Preclinical Tumor Model. Immunohorizons 2019, 3, 149–160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Kriegsman, B.A.; Vangala, P.; Chen, B.J.; Meraner, P.; Brass, A.L.; Garber, M.; Rock, K.L. Frequent Loss of IRF2 in Cancers Leads to Immune Evasion through Decreased MHC Class I Antigen Presentation and Increased PD-L1 Expression. J. Immunol. 2019, 203, 1999–2010. [Google Scholar] [CrossRef] [PubMed]
  213. Fend, L.; Yamazaki, T.; Remy, C.; Fahrner, C.; Gantzer, M.; Nourtier, V.; Préville, X.; Quéméneur, E.; Kepp, O.; Adam, J.; et al. Immune Checkpoint Blockade, Immunogenic Chemotherapy or IFN-α Blockade Boost the Local and Abscopal Effects of Oncolytic Virotherapy. Cancer Res. 2017, 77, 4146–4157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Derynck, R.; Turley, S.J.; Akhurst, R.J. TGFβ biology in cancer progression and immunotherapy. Nat. Rev. Clin. Oncol. 2021, 18, 9–34. [Google Scholar] [CrossRef]
  215. Romano, G.; Santi, L.; Bianco, M.R.; Giuffrè, M.R.; Pettinato, M.; Bugarin, C.; Garanzini, C.; Savarese, L.; Leoni, S.; Cerrito, M.G.; et al. The TGF-β pathway is activated by 5-fluorouracil treatment in drug resistant colorectal carcinoma cells. Oncotarget 2016, 7, 22077–22091. [Google Scholar] [CrossRef] [Green Version]
  216. Quan, Q.; Zhong, F.; Wang, X.; Chen, K.; Guo, L. PAR2 Inhibition Enhanced the Sensitivity of Colorectal Cancer Cells to 5-FU and Reduced EMT Signaling. Oncol. Res. 2019, 27, 779–788. [Google Scholar] [CrossRef]
  217. Brunen, D.; Willems, S.M.; Kellner, U.; Midgley, R.; Simon, I.; Bernards, R. TGF-β: An emerging player in drug resistance. Cell Cycle 2013, 12, 2960–2968. [Google Scholar] [CrossRef] [Green Version]
  218. Cunha, S.I.; Pietras, K. ALK1 as an emerging target for antiangiogenic therapy of cancer. Blood 2011, 117, 6999–7006. [Google Scholar] [CrossRef] [Green Version]
  219. Mazzocca, A.; Fransvea, E.; Lavezzari, G.; Antonaci, S.; Giannelli, G. Inhibition of transforming growth factor beta receptor I kinase blocks hepatocellular carcinoma growth through neo-angiogenesis regulation. Hepatology 2009, 50, 1140–1151. [Google Scholar] [CrossRef]
  220. Zhang, M.; Kleber, S.; Röhrich, M.; Timke, C.; Han, N.; Tuettenberg, J.; Martin-Villalba, A.; Debus, J.; Peschke, P.; Wirkner, U.; et al. Blockade of TGF-β signaling by the TGFβR-I kinase inhibitor LY2109761 enhances radiation response and prolongs survival in glioblastoma. Cancer Res. 2011, 71, 7155–7167. [Google Scholar] [CrossRef] [Green Version]
  221. Akbari, A.; Amanpour, S.; Muhammadnejad, S.; Ghahremani, M.H.; Ghaffari, S.H.; Dehpour, A.R.; Mobini, G.R.; Shidfar, F.; Abastabar, M.; Khoshzaban, A.; et al. Evaluation of antitumor activity of a TGF-beta receptor I inhibitor (SD-208) on human colon adenocarcinoma. DARU J. Pharm. Sci. 2014, 22, 47. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Mariathasan, S.; Turley, S.J.; Nickles, D.; Castiglioni, A.; Yuen, K.; Wang, Y.; Kadel, E.E.; Koeppen, H.; Astarita, J.L.; Cubas, R.; et al. TGFβ attenuates tumour response to PD-L1 blockade by contributing to exclusion of T cells. Nature 2018, 554, 544–548. [Google Scholar] [CrossRef] [PubMed]
  223. Nakanishi, Y.; Duran, A.; L’Hermitte, A.; Shelton, P.M.; Nakanishi, N.; Reina-Campos, M.; Huang, J.; Soldevila, F.; Baaten, B.J.G.; Tauriello, D.V.F.; et al. Simultaneous Loss of Both Atypical Protein Kinase C Genes in the Intestinal Epithelium Drives Serrated Intestinal Cancer by Impairing Immunosurveillance. Immunity 2018, 49, 1132–1147.e7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Chen, J.; Jiang, C.C.; Jin, L.; Zhang, X.D. Regulation of PD-L1: A novel role of pro-survival signalling in cancer. Ann. Oncol. 2016, 27, 409–416. [Google Scholar] [CrossRef]
  225. Bachman, K.E.; Argani, P.; Samuels, Y.; Silliman, N.; Ptak, J.; Szabo, S.; Konishi, H.; Karakas, B.; Blair, B.G.; Lin, C.; et al. The PIK3CA gene is mutated with high frequency in human breast cancers. Cancer Biol. Ther. 2004, 3, 772–775. [Google Scholar] [CrossRef] [Green Version]
  226. Lastwika, K.J.; Wilson, W.; Li, Q.K.; Norris, J.; Xu, H.; Ghazarian, S.R.; Kitagawa, H.; Kawabata, S.; Taube, J.M.; Yao, S.; et al. Control of PD-L1 Expression by Oncogenic Activation of the AKT-mTOR Pathway in Non-Small Cell Lung Cancer. Cancer Res. 2016, 76, 227–238. [Google Scholar] [CrossRef] [Green Version]
  227. Chen, M.; Sharma, A.; Lin, Y.; Wu, Y.; He, Q.; Gu, Y.; Xu, Z.P.; Monteiro, M.; Gu, W. Insluin and epithelial growth factor (EGF) promote programmed death ligand 1(PD-L1) production and transport in colon cancer stem cells. BMC Cancer 2019, 19, 153. [Google Scholar] [CrossRef]
  228. Chida, K.; Kawazoe, A.; Kawazu, M.; Suzuki, T.; Nakamura, Y.; Nakatsura, T.; Kuwata, T.; Ueno, T.; Kuboki, Y.; Kotani, D.; et al. A Low Tumor Mutational Burden and Mutations Are Predictors of a Negative Response to PD-1 Blockade in MSI-H/dMMR Gastrointestinal Tumors. Clin. Cancer Res. 2021, 27, 3714–3724. [Google Scholar] [CrossRef]
  229. Song, M.; Chen, D.; Lu, B.; Wang, C.; Zhang, J.; Huang, L.; Wang, X.; Timmons, C.L.; Hu, J.; Liu, B.; et al. PTEN loss increases PD-L1 protein expression and affects the correlation between PD-L1 expression and clinical parameters in colorectal cancer. PLoS ONE 2013, 8, e65821. [Google Scholar] [CrossRef]
  230. Peng, W.; Chen, J.Q.; Liu, C.; Malu, S.; Creasy, C.; Tetzlaff, M.T.; Xu, C.; McKenzie, J.A.; Zhang, C.; Liang, X.; et al. Loss of PTEN Promotes Resistance to T Cell-Mediated Immunotherapy. Cancer Discov. 2016, 6, 202–216. [Google Scholar] [CrossRef] [Green Version]
  231. Zhao, J.; Chen, A.X.; Gartrell, R.D.; Silverman, A.M.; Aparicio, L.; Chu, T.; Bordbar, D.; Shan, D.; Samanamud, J.; Mahajan, A.; et al. Immune and genomic correlates of response to anti-PD-1 immunotherapy in glioblastoma. Nat. Med. 2019, 25, 462–469. [Google Scholar] [CrossRef] [PubMed]
  232. Coelho, M.A.; de Carné Trécesson, S.; Rana, S.; Zecchin, D.; Moore, C.; Molina-Arcas, M.; East, P.; Spencer-Dene, B.; Nye, E.; Barnouin, K.; et al. Oncogenic RAS Signaling Promotes Tumor Immunoresistance by Stabilizing PD-L1 mRNA. Immunity 2017, 47, 1083–1099.e6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Napolitano, S.; Matrone, N.; Muddassir, A.L.; Martini, G.; Sorokin, A.; De Falco, V.; Giunta, E.F.; Ciardiello, D.; Martinelli, E.; Belli, V.; et al. Triple blockade of EGFR, MEK and PD-L1 has antitumor activity in colorectal cancer models with constitutive activation of MAPK signaling and PD-L1 overexpression. J. Exp. Clin. Cancer Res. 2019, 38, 492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Weiss, A.; Littman, D.R. Signal transduction by lymphocyte antigen receptors. Cell 1994, 76, 263–274. [Google Scholar] [CrossRef]
  235. Terheyden, P.; Schrama, D.; Pedersen, L.Ø.; Andersen, M.H.; Kämpgen, E.; thor Straten, P.; Becker, J.C. Longitudinal analysis of MART-1/HLA-A2-reactive T cells over the course of melanoma progression. Scand. J. Immunol. 2003, 58, 566–571. [Google Scholar] [CrossRef] [Green Version]
  236. Chen, D.; Heath, V.; O’Garra, A.; Johnston, J.; McMahon, M. Sustained activation of the raf-MEK-ERK pathway elicits cytokine unresponsiveness in T cells. J. Immunol. 1999, 163, 5796–5805. [Google Scholar]
  237. Liu, L.; Mayes, P.A.; Eastman, S.; Shi, H.; Yadavilli, S.; Zhang, T.; Yang, J.; Seestaller-Wehr, L.; Zhang, S.-Y.; Hopson, C.; et al. The BRAF and MEK Inhibitors Dabrafenib and Trametinib: Effects on Immune Function and in Combination with Immunomodulatory Antibodies Targeting PD-1, PD-L1, and CTLA-4. Clin. Cancer Res. 2015, 21, 1639–1651. [Google Scholar] [CrossRef] [Green Version]
  238. Vella, L.J.; Pasam, A.; Dimopoulos, N.; Andrews, M.; Knights, A.; Puaux, A.-L.; Louahed, J.; Chen, W.; Woods, K.; Cebon, J.S. MEK inhibition, alone or in combination with BRAF inhibition, affects multiple functions of isolated normal human lymphocytes and dendritic cells. Cancer Immunol. Res. 2014, 2, 351–360. [Google Scholar] [CrossRef] [Green Version]
  239. Boni, A.; Cogdill, A.P.; Dang, P.; Udayakumar, D.; Njauw, C.-N.J.; Sloss, C.M.; Ferrone, C.R.; Flaherty, K.T.; Lawrence, D.P.; Fisher, D.E.; et al. Selective BRAFV600E inhibition enhances T-cell recognition of melanoma without affecting lymphocyte function. Cancer Res. 2010, 70, 5213–5219. [Google Scholar] [CrossRef] [Green Version]
  240. Kist, M.; Vucic, D. Cell death pathways: Intricate connections and disease implications. EMBO J. 2021, 40, e106700. [Google Scholar] [CrossRef]
  241. Scaffidi, C.; Fulda, S.; Srinivasan, A.; Friesen, C.; Li, F.; Tomaselli, K.J.; Debatin, K.M.; Krammer, P.H.; Peter, M.E. Two CD95 (APO-1/Fas) signaling pathways. EMBO J. 1998, 17, 1675–1687. [Google Scholar] [CrossRef] [PubMed]
  242. Warren, C.F.A.; Wong-Brown, M.W.; Bowden, N.A. BCL-2 family isoforms in apoptosis and cancer. Cell Death Dis. 2019, 10, 177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Adams, J.M.; Cory, S. The Bcl-2 apoptotic switch in cancer development and therapy. Oncogene 2007, 26, 1324–1337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Zhang, L.; Yu, J.; Park, B.H.; Kinzler, K.W.; Vogelstein, B. Role of BAX in the apoptotic response to anticancer agents. Science 2000, 290, 989–992. [Google Scholar] [CrossRef]
  245. van Delft, M.F.; Wei, A.H.; Mason, K.D.; Vandenberg, C.J.; Chen, L.; Czabotar, P.E.; Willis, S.N.; Scott, C.L.; Day, C.L.; Cory, S.; et al. The BH3 mimetic ABT-737 targets selective Bcl-2 proteins and efficiently induces apoptosis via Bak/Bax if Mcl-1 is neutralized. Cancer Cell 2006, 10, 389–399. [Google Scholar] [CrossRef] [Green Version]
  246. Konopleva, M.; Contractor, R.; Tsao, T.; Samudio, I.; Ruvolo, P.P.; Kitada, S.; Deng, X.; Zhai, D.; Shi, Y.-X.; Sneed, T.; et al. Mechanisms of apoptosis sensitivity and resistance to the BH3 mimetic ABT-737 in acute myeloid leukemia. Cancer Cell 2006, 10, 375–388. [Google Scholar] [CrossRef] [Green Version]
  247. Joudeh, J.; Claxton, D. Obatoclax mesylate: Pharmacology and potential for therapy of hematological neoplasms. Expert Opin. Investig. Drugs 2012, 21, 363–373. [Google Scholar] [CrossRef] [PubMed]
  248. Timme, C.R.; Gruidl, M.; Yeatman, T.J. Gamma-secretase inhibition attenuates oxaliplatin-induced apoptosis through increased Mcl-1 and/or Bcl-xL in human colon cancer cells. Apoptosis 2013, 18, 1163–1174. [Google Scholar] [CrossRef]
  249. Hunter, A.M.; LaCasse, E.C.; Korneluk, R.G. The inhibitors of apoptosis (IAPs) as cancer targets. Apoptosis 2007, 12, 1543–1568. [Google Scholar] [CrossRef] [PubMed]
  250. Fulda, S.; Vucic, D. Targeting IAP proteins for therapeutic intervention in cancer. Nat. Rev. Drug Discov. 2012, 11, 109–124. [Google Scholar] [CrossRef]
  251. Connolly, K.; Mitter, R.; Muir, M.; Jodrell, D.; Guichard, S. Stable XIAP knockdown clones of HCT116 colon cancer cells are more sensitive to TRAIL, taxanes and irradiation in vitro. Cancer Chemother. Pharmacol. 2009, 64, 307–316. [Google Scholar] [CrossRef] [PubMed]
  252. Lippa, M.S.; Strockbine, L.D.; Le, T.T.; Branstetter, D.G.; Strathdee, C.A.; Holland, P.M. Expression of anti-apoptotic factors modulates Apo2L/TRAIL resistance in colon carcinoma cells. Apoptosis 2007, 12, 1465–1478. [Google Scholar] [CrossRef] [PubMed]
  253. Xiong, Z.; Fu, Z.; Shi, J.; Jiang, X.; Wan, H. HtrA1 Down-regulation Induces Cisplatin Resistance in Colon Cancer by Increasing XIAP and Activating PI3K/Akt Pathway. Ann. Clin. Lab. Sci. 2017, 47, 264–270. [Google Scholar] [PubMed]
  254. Zhang, Y.; Talmon, G.; Wang, J. MicroRNA-587 antagonizes 5-FU-induced apoptosis and confers drug resistance by regulating PPP2R1B expression in colorectal cancer. Cell Death Dis. 2015, 6, e1845. [Google Scholar] [CrossRef] [Green Version]
  255. Gyrd-Hansen, M.; Meier, P. IAPs: From caspase inhibitors to modulators of NF-kappaB, inflammation and cancer. Nat. Rev. Cancer 2010, 10, 561–574. [Google Scholar] [CrossRef]
  256. Enoch, T.; Norbury, C. Cellular responses to DNA damage: Cell-cycle checkpoints, apoptosis and the roles of p53 and ATM. Trends Biochem. Sci. 1995, 20, 426–430. [Google Scholar] [CrossRef]
  257. Yao, J.; Huang, A.; Zheng, X.; Liu, T.; Lin, Z.; Zhang, S.; Yang, Q.; Zhang, T.; Ma, H. 53BP1 loss induces chemoresistance of colorectal cancer cells to 5-fluorouracil by inhibiting the ATM-CHK2-P53 pathway. J. Cancer Res. Clin. Oncol. 2017, 143, 419–431. [Google Scholar] [CrossRef]
  258. Elsaleh, H.; Powell, B.; McCaul, K.; Grieu, F.; Grant, R.; Joseph, D.; Iacopetta, B. P53 alteration and microsatellite instability have predictive value for survival benefit from chemotherapy in stage III colorectal carcinoma. Clin. Cancer Res. 2001, 7, 1343–1349. [Google Scholar]
  259. Bunz, F.; Hwang, P.M.; Torrance, C.; Waldman, T.; Zhang, Y.; Dillehay, L.; Williams, J.; Lengauer, C.; Kinzler, K.W.; Vogelstein, B. Disruption of p53 in human cancer cells alters the responses to therapeutic agents. J. Clin. Investig. 1999, 104, 263–269. [Google Scholar] [CrossRef] [Green Version]
  260. Russo, A.; Bazan, V.; Iacopetta, B.; Kerr, D.; Soussi, T.; Gebbia, N. The TP53 colorectal cancer international collaborative study on the prognostic and predictive significance of p53 mutation: Influence of tumor site, type of mutation, and adjuvant treatment. J. Clin. Oncol. 2005, 23, 7518–7528. [Google Scholar] [CrossRef]
  261. Oliner, J.D.; Pietenpol, J.A.; Thiagalingam, S.; Gyuris, J.; Kinzler, K.W.; Vogelstein, B. Oncoprotein MDM2 conceals the activation domain of tumour suppressor p53. Nature 1993, 362, 857–860. [Google Scholar] [CrossRef] [PubMed]
  262. Honda, R.; Tanaka, H.; Yasuda, H. Oncoprotein MDM2 is a ubiquitin ligase E3 for tumor suppressor p53. FEBS Lett. 1997, 420, 25–27. [Google Scholar] [CrossRef] [Green Version]
  263. Vassilev, L.T.; Vu, B.T.; Graves, B.; Carvajal, D.; Podlaski, F.; Filipovic, Z.; Kong, N.; Kammlott, U.; Lukacs, C.; Klein, C.; et al. In vivo activation of the p53 pathway by small-molecule antagonists of MDM2. Science 2004, 303, 844–848. [Google Scholar] [CrossRef] [Green Version]
  264. Nadler-Milbauer, M.; Apter, L.; Haupt, Y.; Haupt, S.; Barenholz, Y.; Minko, T.; Rubinstein, A. Synchronized release of Doxil and Nutlin-3 by remote degradation of polysaccharide matrices and its possible use in the local treatment of colorectal cancer. J. Drug Target. 2011, 19, 859–873. [Google Scholar] [CrossRef] [PubMed]
  265. Wang, S.; Zhang, Y.; Huang, J.; Wong, C.C.; Zhai, J.; Li, C.; Wei, G.; Zhao, L.; Wang, G.; Wei, H.; et al. TRIM67 Activates p53 to Suppress Colorectal Cancer Initiation and Progression. Cancer Res. 2019, 79, 4086–4098. [Google Scholar] [CrossRef] [Green Version]
  266. Glick, D.; Barth, S.; Macleod, K.F. Autophagy: Cellular and molecular mechanisms. J. Pathol. 2010, 221, 3–12. [Google Scholar] [CrossRef] [Green Version]
  267. Levy, J.M.M.; Towers, C.G.; Thorburn, A. Targeting autophagy in cancer. Nat. Rev. Cancer 2017, 17, 528–542. [Google Scholar] [CrossRef]
  268. Aredia, F.; Scovassi, A.I. Manipulation of autophagy in cancer cells: An innovative strategy to fight drug resistance. Future Med. Chem. 2013, 5, 1009–1021. [Google Scholar] [CrossRef] [Green Version]
  269. Kumar, A.; Singh, U.K.; Chaudhary, A. Targeting autophagy to overcome drug resistance in cancer therapy. Future Med. Chem. 2015, 7, 1535–1542. [Google Scholar] [CrossRef]
  270. Koustas, E.; Sarantis, P.; Theoharis, S.; Saetta, A.A.; Chatziandreou, I.; Kyriakopoulou, G.; Giannopoulou, I.; Michelli, M.; Schizas, D.; Papavassiliou, A.G.; et al. Autophagy-related Proteins as a Prognostic Factor of Patients With Colorectal Cancer. Am. J. Clin. Oncol. 2019, 42, 767–776. [Google Scholar] [CrossRef] [Green Version]
  271. Yu, T.; Guo, F.; Yu, Y.; Sun, T.; Ma, D.; Han, J.; Qian, Y.; Kryczek, I.; Sun, D.; Nagarsheth, N.; et al. Fusobacterium nucleatum Promotes Chemoresistance to Colorectal Cancer by Modulating Autophagy. Cell 2017, 170, 548–563.e16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  272. Hu, F.; Song, D.; Yan, Y.; Huang, C.; Shen, C.; Lan, J.; Chen, Y.; Liu, A.; Wu, Q.; Sun, L.; et al. IL-6 regulates autophagy and chemotherapy resistance by promoting BECN1 phosphorylation. Nat. Commun. 2021, 12, 3651. [Google Scholar] [CrossRef]
  273. Zhang, Y.; Li, C.; Liu, X.; Wang, Y.; Zhao, R.; Yang, Y.; Zheng, X.; Zhang, Y.; Zhang, X. circHIPK3 promotes oxaliplatin-resistance in colorectal cancer through autophagy by sponging miR-637. EBioMedicine 2019, 48, 277–288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  274. Wang, X.; Lan, Z.; He, J.; Lai, Q.; Yao, X.; Li, Q.; Liu, Y.; Lai, H.; Gu, C.; Yan, Q.; et al. LncRNA SNHG6 promotes chemoresistance through ULK1-induced autophagy by sponging miR-26a-5p in colorectal cancer cells. Cancer Cell Int. 2019, 19, 234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  275. Wang, M.; Han, D.; Yuan, Z.; Hu, H.; Zhao, Z.; Yang, R.; Jin, Y.; Zou, C.; Chen, Y.; Wang, G.; et al. Long non-coding RNA H19 confers 5-Fu resistance in colorectal cancer by promoting SIRT1-mediated autophagy. Cell Death Dis. 2018, 9, 1149. [Google Scholar] [CrossRef] [PubMed]
  276. Liu, F.; Ai, F.-Y.; Zhang, D.-C.; Tian, L.; Yang, Z.-Y.; Liu, S.-J. LncRNA NEAT1 knockdown attenuates autophagy to elevate 5-FU sensitivity in colorectal cancer via targeting miR-34a. Cancer Med. 2020, 9, 1079–1091. [Google Scholar] [CrossRef] [Green Version]
  277. Han, Y.; Zhou, S.; Wang, X.; Mao, E.; Huang, L. SNHG14 stimulates cell autophagy to facilitate cisplatin resistance of colorectal cancer by regulating miR-186/ATG14 axis. Biomed. Pharmacother. 2020, 121, 109580. [Google Scholar] [CrossRef]
  278. Guo, G.-F.; Jiang, W.-Q.; Zhang, B.; Cai, Y.-C.; Xu, R.-H.; Chen, X.-X.; Wang, F.; Xia, L.-P. Autophagy-related proteins Beclin-1 and LC3 predict cetuximab efficacy in advanced colorectal cancer. World J. Gastroenterol. 2011, 17, 4779–4786. [Google Scholar] [CrossRef]
  279. Clarke, A.J.; Simon, A.K. Autophagy in the renewal, differentiation and homeostasis of immune cells. Nat. Rev. Immunol. 2019, 19, 170–183. [Google Scholar] [CrossRef]
  280. Michaud, M.; Martins, I.; Sukkurwala, A.Q.; Adjemian, S.; Ma, Y.; Pellegatti, P.; Shen, S.; Kepp, O.; Scoazec, M.; Mignot, G.; et al. Autophagy-dependent anticancer immune responses induced by chemotherapeutic agents in mice. Science 2011, 334, 1573–1577. [Google Scholar] [CrossRef]
  281. Dixon, S.J.; Lemberg, K.M.; Lamprecht, M.R.; Skouta, R.; Zaitsev, E.M.; Gleason, C.E.; Patel, D.N.; Bauer, A.J.; Cantley, A.M.; Yang, W.S.; et al. Ferroptosis: An iron-dependent form of nonapoptotic cell death. Cell 2012, 149, 1060–1072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  282. Yang, C.; Zhang, Y.; Lin, S.; Liu, Y.; Li, W. Suppressing the KIF20A/NUAK1/Nrf2/GPX4 signaling pathway induces ferroptosis and enhances the sensitivity of colorectal cancer to oxaliplatin. Aging 2021, 13, 13515–13534. [Google Scholar] [CrossRef] [PubMed]
  283. Chaudhary, N.; Choudhary, B.S.; Shah, S.G.; Khapare, N.; Dwivedi, N.; Gaikwad, A.; Joshi, N.; Raichanna, J.; Basu, S.; Gurjar, M.; et al. Lipocalin 2 expression promotes tumor progression and therapy resistance by inhibiting ferroptosis in colorectal cancer. Int. J. Cancer 2021, 149, 1495–1511. [Google Scholar] [CrossRef]
  284. Mou, Y.; Wang, J.; Wu, J.; He, D.; Zhang, C.; Duan, C.; Li, B. Ferroptosis, a new form of cell death: Opportunities and challenges in cancer. J. Hematol. Oncol. 2019, 12, 34. [Google Scholar] [CrossRef] [Green Version]
  285. Kang, K.A.; Piao, M.J.; Kim, K.C.; Kang, H.K.; Chang, W.Y.; Park, I.C.; Keum, Y.S.; Surh, Y.J.; Hyun, J.W. Epigenetic modification of Nrf2 in 5-fluorouracil-resistant colon cancer cells: Involvement of TET-dependent DNA demethylation. Cell Death Dis. 2014, 5, e1183. [Google Scholar] [CrossRef]
  286. Yang, W.S.; Stockwell, B.R. Synthetic lethal screening identifies compounds activating iron-dependent, nonapoptotic cell death in oncogenic-RAS-harboring cancer cells. Chem. Biol. 2008, 15, 234–245. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Yang, J.; Mo, J.; Dai, J.; Ye, C.; Cen, W.; Zheng, X.; Jiang, L.; Ye, L. Cetuximab promotes RSL3-induced ferroptosis by suppressing the Nrf2/HO-1 signalling pathway in KRAS mutant colorectal cancer. Cell Death Dis. 2021, 12, 1079. [Google Scholar] [CrossRef]
  288. Wang, W.; Green, M.; Choi, J.E.; Gijón, M.; Kennedy, P.D.; Johnson, J.K.; Liao, P.; Lang, X.; Kryczek, I.; Sell, A.; et al. CD8 T cells regulate tumour ferroptosis during cancer immunotherapy. Nature 2019, 569, 270–274. [Google Scholar] [CrossRef] [PubMed]
  289. Duan, X.; Chan, C.; Han, W.; Guo, N.; Weichselbaum, R.R.; Lin, W. Immunostimulatory nanomedicines synergize with checkpoint blockade immunotherapy to eradicate colorectal tumors. Nat. Commun. 2019, 10, 1899. [Google Scholar] [CrossRef] [Green Version]
  290. Han, W.; Duan, X.; Ni, K.; Li, Y.; Chan, C.; Lin, W. Co-delivery of dihydroartemisinin and pyropheophorbide-iron elicits ferroptosis to potentiate cancer immunotherapy. Biomaterials 2021, 280, 121315. [Google Scholar] [CrossRef]
  291. Katz, O.B.; Shaked, Y. Host effects contributing to cancer therapy resistance. Drug Resist. Updates 2015, 19, 33–42. [Google Scholar] [CrossRef] [PubMed]
  292. Schiavoni, G.; Gabriele, L.; Mattei, F. The tumor microenvironment: A pitch for multiple players. Front. Oncol. 2013, 3, 90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Nurmik, M.; Ullmann, P.; Rodriguez, F.; Haan, S.; Letellier, E. In search of definitions: Cancer-associated fibroblasts and their markers. Int. J. Cancer 2020, 146, 895–905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  294. Liu, R.; Li, J.; Xie, K.; Zhang, T.; Lei, Y.; Chen, Y.; Zhang, L.; Huang, K.; Wang, K.; Wu, H.; et al. FGFR4 promotes stroma-induced epithelial-to-mesenchymal transition in colorectal cancer. Cancer Res. 2013, 73, 5926–5935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  295. Guillén Díaz-Maroto, N.; Sanz-Pamplona, R.; Berdiel-Acer, M.; Cimas, F.J.; García, E.; Gonçalves-Ribeiro, S.; Albert, N.; Garcia-Vicién, G.; Capella, G.; Moreno, V.; et al. Noncanonical TGFβ Pathway Relieves the Blockade of IL1β/TGFβ-Mediated Crosstalk between Tumor and Stroma: TGFBR1 and TAK1 Inhibition in Colorectal Cancer. Clin. Cancer Res. 2019, 25, 4466–4479. [Google Scholar] [CrossRef] [Green Version]
  296. Tang, Y.-A.; Chen, Y.-F.; Bao, Y.; Mahara, S.; Yatim, S.M.J.M.; Oguz, G.; Lee, P.L.; Feng, M.; Cai, Y.; Tan, E.Y.; et al. Hypoxic tumor microenvironment activates GLI2 via HIF-1α and TGF-β2 to promote chemoresistance in colorectal cancer. Proc. Natl. Acad. Sci. USA 2018, 115, E5990–E5999. [Google Scholar] [CrossRef] [Green Version]
  297. Li, Z.; Chan, K.; Qi, Y.; Lu, L.; Ning, F.; Wu, M.; Wang, H.; Wang, Y.; Cai, S.; Du, J. Participation of CCL1 in Snail-Positive Fibroblasts in Colorectal Cancer Contribute to 5-Fluorouracil/Paclitaxel Chemoresistance. Cancer Res. Treat. 2018, 50, 894–907. [Google Scholar] [CrossRef] [Green Version]
  298. Lotti, F.; Jarrar, A.M.; Pai, R.K.; Hitomi, M.; Lathia, J.; Mace, A.; Gantt, G.A.; Sukhdeo, K.; DeVecchio, J.; Vasanji, A.; et al. Chemotherapy activates cancer-associated fibroblasts to maintain colorectal cancer-initiating cells by IL-17A. J. Exp. Med. 2013, 210, 2851–2872. [Google Scholar] [CrossRef] [Green Version]
  299. Hu, J.L.; Wang, W.; Lan, X.L.; Zeng, Z.C.; Liang, Y.S.; Yan, Y.R.; Song, F.Y.; Wang, F.F.; Zhu, X.H.; Liao, W.J.; et al. CAFs secreted exosomes promote metastasis and chemotherapy resistance by enhancing cell stemness and epithelial-mesenchymal transition in colorectal cancer. Mol. Cancer 2019, 18, 91. [Google Scholar] [CrossRef] [Green Version]
  300. Ren, J.; Ding, L.; Zhang, D.; Shi, G.; Xu, Q.; Shen, S.; Wang, Y.; Wang, T.; Hou, Y. Carcinoma-associated fibroblasts promote the stemness and chemoresistance of colorectal cancer by transferring exosomal lncRNA H19. Theranostics 2018, 8, 3932–3948. [Google Scholar] [CrossRef]
  301. Deng, X.; Ruan, H.; Zhang, X.; Xu, X.; Zhu, Y.; Peng, H.; Zhang, X.; Kong, F.; Guan, M. Long noncoding RNA CCAL transferred from fibroblasts by exosomes promotes chemoresistance of colorectal cancer cells. Int. J. Cancer 2020, 146, 1700–1716. [Google Scholar] [CrossRef] [PubMed]
  302. Woolston, A.; Khan, K.; Spain, G.; Barber, L.J.; Griffiths, B.; Gonzalez-Exposito, R.; Hornsteiner, L.; Punta, M.; Patil, Y.; Newey, A.; et al. Genomic and Transcriptomic Determinants of Therapy Resistance and Immune Landscape Evolution during Anti-EGFR Treatment in Colorectal Cancer. Cancer Cell 2019, 36, 35–50.e9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Luraghi, P.; Reato, G.; Cipriano, E.; Sassi, F.; Orzan, F.; Bigatto, V.; De Bacco, F.; Menietti, E.; Han, M.; Rideout, W.M.; et al. MET signaling in colon cancer stem-like cells blunts the therapeutic response to EGFR inhibitors. Cancer Res. 2014, 74, 1857–1869. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Chakravarthy, A.; Furness, A.; Joshi, K.; Ghorani, E.; Ford, K.; Ward, M.J.; King, E.V.; Lechner, M.; Marafioti, T.; Quezada, S.A.; et al. Pan-cancer deconvolution of tumour composition using DNA methylation. Nat. Commun. 2018, 9, 3220. [Google Scholar] [CrossRef] [Green Version]
  305. Kieffer, Y.; Hocine, H.R.; Gentric, G.; Pelon, F.; Bernard, C.; Bourachot, B.; Lameiras, S.; Albergante, L.; Bonneau, C.; Guyard, A.; et al. Single-Cell Analysis Reveals Fibroblast Clusters Linked to Immunotherapy Resistance in Cancer. Cancer Discov. 2020, 10, 1330–1351. [Google Scholar] [CrossRef] [PubMed]
  306. Ford, K.; Hanley, C.J.; Mellone, M.; Szyndralewiez, C.; Heitz, F.; Wiesel, P.; Wood, O.; Machado, M.; Lopez, M.-A.; Ganesan, A.-P.; et al. NOX4 Inhibition Potentiates Immunotherapy by Overcoming Cancer-Associated Fibroblast-Mediated CD8 T-cell Exclusion from Tumors. Cancer Res. 2020, 80, 1846–1860. [Google Scholar] [CrossRef] [Green Version]
  307. Gabrilovich, D.I.; Nagaraj, S. Myeloid-derived suppressor cells as regulators of the immune system. Nat. Rev. Immunol. 2009, 9, 162–174. [Google Scholar] [CrossRef] [Green Version]
  308. Katoh, H.; Wang, D.; Daikoku, T.; Sun, H.; Dey, S.K.; Dubois, R.N. CXCR2-expressing myeloid-derived suppressor cells are essential to promote colitis-associated tumorigenesis. Cancer Cell 2013, 24, 631–644. [Google Scholar] [CrossRef] [Green Version]
  309. Gabrilovich, D.I.; Ostrand-Rosenberg, S.; Bronte, V. Coordinated regulation of myeloid cells by tumours. Nat. Rev. Immunol. 2012, 12, 253–268. [Google Scholar] [CrossRef] [Green Version]
  310. Bronte, V.; Brandau, S.; Chen, S.-H.; Colombo, M.P.; Frey, A.B.; Greten, T.F.; Mandruzzato, S.; Murray, P.J.; Ochoa, A.; Ostrand-Rosenberg, S.; et al. Recommendations for myeloid-derived suppressor cell nomenclature and characterization standards. Nat. Commun. 2016, 7, 12150. [Google Scholar] [CrossRef] [Green Version]
  311. Chun, E.; Lavoie, S.; Michaud, M.; Gallini, C.A.; Kim, J.; Soucy, G.; Odze, R.; Glickman, J.N.; Garrett, W.S. CCL2 Promotes Colorectal Carcinogenesis by Enhancing Polymorphonuclear Myeloid-Derived Suppressor Cell Population and Function. Cell Rep. 2015, 12, 244–257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  312. De Cicco, P.; Sanders, T.; Cirino, G.; Maloy, K.J.; Ianaro, A. Hydrogen Sulfide Reduces Myeloid-Derived Suppressor Cell-Mediated Inflammatory Response in a Model of -Induced Colitis. Front. Immunol. 2018, 9, 499. [Google Scholar] [CrossRef] [Green Version]
  313. Toor, S.M.; Syed Khaja, A.S.; El Salhat, H.; Bekdache, O.; Kanbar, J.; Jaloudi, M.; Elkord, E. Increased Levels of Circulating and Tumor-Infiltrating Granulocytic Myeloid Cells in Colorectal Cancer Patients. Front. Immunol. 2016, 7, 560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  314. Rodriguez, P.C.; Quiceno, D.G.; Zabaleta, J.; Ortiz, B.; Zea, A.H.; Piazuelo, M.B.; Delgado, A.; Correa, P.; Brayer, J.; Sotomayor, E.M.; et al. Arginase I production in the tumor microenvironment by mature myeloid cells inhibits T-cell receptor expression and antigen-specific T-cell responses. Cancer Res. 2004, 64, 5839–5849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  315. Rodriguez, P.C.; Quiceno, D.G.; Ochoa, A.C. L-arginine availability regulates T-lymphocyte cell-cycle progression. Blood 2007, 109, 1568–1573. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  316. Stiff, A.; Trikha, P.; Mundy-Bosse, B.; McMichael, E.; Mace, T.A.; Benner, B.; Kendra, K.; Campbell, A.; Gautam, S.; Abood, D.; et al. Nitric Oxide Production by Myeloid-Derived Suppressor Cells Plays a Role in Impairing Fc Receptor-Mediated Natural Killer Cell Function. Clin. Cancer Res. 2018, 24, 1891–1904. [Google Scholar] [CrossRef] [Green Version]
  317. Delgado-Ramirez, Y.; Baltazar-Perez, I.; Martinez, Y.; Callejas, B.E.; Medina-Andrade, I.; Olguín, J.E.; Delgado-Buenrostro, N.L.; Chirino, Y.I.; Terrazas, L.I.; Leon-Cabrera, S. STAT1 Is Required for Decreasing Accumulation of Granulocytic Cells via IL-17 during Initial Steps of Colitis-Associated Cancer. Int. J. Mol. Sci. 2021, 22, 7695. [Google Scholar] [CrossRef]
  318. Wu, P.; Wu, D.; Ni, C.; Ye, J.; Chen, W.; Hu, G.; Wang, Z.; Wang, C.; Zhang, Z.; Xia, W.; et al. γδT17 cells promote the accumulation and expansion of myeloid-derived suppressor cells in human colorectal cancer. Immunity 2014, 40, 785–800. [Google Scholar] [CrossRef] [Green Version]
  319. Yang, R.; Cai, T.-T.; Wu, X.-J.; Liu, Y.-N.; He, J.; Zhang, X.-S.; Ma, G.; Li, J. Tumour YAP1 and PTEN expression correlates with tumour-associated myeloid suppressor cell expansion and reduced survival in colorectal cancer. Immunology 2018, 155, 263–272. [Google Scholar] [CrossRef]
  320. Liao, W.; Overman, M.J.; Boutin, A.T.; Shang, X.; Zhao, D.; Dey, P.; Li, J.; Wang, G.; Lan, Z.; Li, J.; et al. KRAS-IRF2 Axis Drives Immune Suppression and Immune Therapy Resistance in Colorectal Cancer. Cancer Cell 2019, 35, 559–572.e7. [Google Scholar] [CrossRef] [Green Version]
  321. Condamine, T.; Kumar, V.; Ramachandran, I.R.; Youn, J.-I.; Celis, E.; Finnberg, N.; El-Deiry, W.S.; Winograd, R.; Vonderheide, R.H.; English, N.R.; et al. ER stress regulates myeloid-derived suppressor cell fate through TRAIL-R-mediated apoptosis. J. Clin. Investig. 2014, 124, 2626–2639. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  322. Dominguez, G.A.; Condamine, T.; Mony, S.; Hashimoto, A.; Wang, F.; Liu, Q.; Forero, A.; Bendell, J.; Witt, R.; Hockstein, N.; et al. Selective Targeting of Myeloid-Derived Suppressor Cells in Cancer Patients Using DS-8273a, an Agonistic TRAIL-R2 Antibody. Clin. Cancer Res. 2017, 23, 2942–2950. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  323. Chen, D.; Zhang, X.; Li, Z.; Zhu, B. Metabolic regulatory crosstalk between tumor microenvironment and tumor-associated macrophages. Theranostics 2021, 11, 1016–1030. [Google Scholar] [CrossRef] [PubMed]
  324. Hao, N.-B.; Lü, M.-H.; Fan, Y.-H.; Cao, Y.-L.; Zhang, Z.-R.; Yang, S.-M. Macrophages in tumor microenvironments and the progression of tumors. Clin. Dev. Immunol. 2012, 2012, 948098. [Google Scholar] [CrossRef] [Green Version]
  325. Jinushi, M.; Chiba, S.; Yoshiyama, H.; Masutomi, K.; Kinoshita, I.; Dosaka-Akita, H.; Yagita, H.; Takaoka, A.; Tahara, H. Tumor-associated macrophages regulate tumorigenicity and anticancer drug responses of cancer stem/initiating cells. Proc. Natl. Acad. Sci. USA 2011, 108, 12425–12430. [Google Scholar] [CrossRef] [Green Version]
  326. Yin, Y.; Yao, S.; Hu, Y.; Feng, Y.; Li, M.; Bian, Z.; Zhang, J.; Qin, Y.; Qi, X.; Zhou, L.; et al. The Immune-microenvironment Confers Chemoresistance of Colorectal Cancer through Macrophage-Derived IL6. Clin. Cancer Res. 2017, 23, 7375–7387. [Google Scholar] [CrossRef] [Green Version]
  327. Malesci, A.; Bianchi, P.; Celesti, G.; Basso, G.; Marchesi, F.; Grizzi, F.; Di Caro, G.; Cavalleri, T.; Rimassa, L.; Palmqvist, R.; et al. Tumor-associated macrophages and response to 5-fluorouracil adjuvant therapy in stage III colorectal cancer. Oncoimmunology 2017, 6, e1342918. [Google Scholar] [CrossRef]
  328. Feng, Q.; Chang, W.; Mao, Y.; He, G.; Zheng, P.; Tang, W.; Wei, Y.; Ren, L.; Zhu, D.; Ji, M.; et al. Tumor-associated Macrophages as Prognostic and Predictive Biomarkers for Postoperative Adjuvant Chemotherapy in Patients with Stage II Colon Cancer. Clin. Cancer Res. 2019, 25, 3896–3907. [Google Scholar] [CrossRef] [Green Version]
  329. Gordon, S.R.; Maute, R.L.; Dulken, B.W.; Hutter, G.; George, B.M.; McCracken, M.N.; Gupta, R.; Tsai, J.M.; Sinha, R.; Corey, D.; et al. PD-1 expression by tumour-associated macrophages inhibits phagocytosis and tumour immunity. Nature 2017, 545, 495–499. [Google Scholar] [CrossRef]
  330. Stanley, E.R.; Chitu, V. CSF-1 receptor signaling in myeloid cells. Cold Spring Harb. Perspect. Biol. 2014, 6, a021857. [Google Scholar] [CrossRef] [Green Version]
  331. Doleschel, D.; Hoff, S.; Koletnik, S.; Rix, A.; Zopf, D.; Kiessling, F.; Lederle, W. Regorafenib enhances anti-PD1 immunotherapy efficacy in murine colorectal cancers and their combination prevents tumor regrowth. J. Exp. Clin. Cancer Res. 2021, 40, 288. [Google Scholar] [CrossRef] [PubMed]
  332. Fukuoka, S.; Hara, H.; Takahashi, N.; Kojima, T.; Kawazoe, A.; Asayama, M.; Yoshii, T.; Kotani, D.; Tamura, H.; Mikamoto, Y.; 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, 2053–2061. [Google Scholar] [CrossRef] [PubMed]
  333. Fakih, M.; Raghav, K.P.S.; Chang, D.Z.; Bendell, J.C.; Larson, T.; Cohn, A.L.; Huyck, T.K.; Cosgrove, D.; Fiorillo, J.A.; Garbo, L.E.; et al. Single-arm, phase 2 study of regorafenib plus nivolumab in patients with mismatch repair-proficient (pMMR)/microsatellite stable (MSS) colorectal cancer (CRC). J. Clin. Oncol. 2021, 39, 3560. [Google Scholar] [CrossRef]
  334. Baer, C.; Squadrito, M.L.; Laoui, D.; Thompson, D.; Hansen, S.K.; Kiialainen, A.; Hoves, S.; Ries, C.H.; Ooi, C.-H.; De Palma, M. Suppression of microRNA activity amplifies IFN-γ-induced macrophage activation and promotes anti-tumour immunity. Nat. Cell Biol. 2016, 18, 790–802. [Google Scholar] [CrossRef] [PubMed]
  335. Wu, M.-F.; Lin, C.-A.; Yuan, T.-H.; Yeh, H.-Y.; Su, S.-F.; Guo, C.-L.; Chang, G.-C.; Li, K.-C.; Ho, C.-C.; Chen, H.-W. The M1/M2 spectrum and plasticity of malignant pleural effusion-macrophage in advanced lung cancer. Cancer Immunol. Immunother. 2021, 70, 1435–1450. [Google Scholar] [CrossRef]
  336. Cassetta, L.; Pollard, J.W. Tumor-associated macrophages. Curr. Biol. 2020, 30, R246–R248. [Google Scholar] [CrossRef]
  337. Song, M.; Li, Z.-H.; Gu, H.-S.; Tang, R.-Y.; Zhang, R.; Zhu, Y.-L.; Liu, J.-L.; Zhang, J.-J.; Wang, L.-Y. Spore Polysaccharide Inhibits the Growth of Hepatocellular Carcinoma Cells by Altering Macrophage Polarity and Induction of Apoptosis. J. Immunol. Res. 2021, 2021, 6696606. [Google Scholar] [CrossRef]
  338. Bi, S.; Huang, W.; Chen, S.; Huang, C.; Li, C.; Guo, Z.; Yang, J.; Zhu, J.; Song, L.; Yu, R. Cordyceps militaris polysaccharide converts immunosuppressive macrophages into M1-like phenotype and activates T lymphocytes by inhibiting the PD-L1/PD-1 axis between TAMs and T lymphocytes. Int. J. Biol. Macromol. 2020, 150, 261–280. [Google Scholar] [CrossRef]
  339. Nie, W.; Wu, G.; Zhang, J.; Huang, L.-L.; Ding, J.; Jiang, A.; Zhang, Y.; Liu, Y.; Li, J.; Pu, K.; et al. Responsive Exosome Nano-bioconjugates for Synergistic Cancer Therapy. Angew. Chem. Int. Ed. Engl. 2020, 59, 2018–2022. [Google Scholar] [CrossRef]
  340. Simons, M.; Gordon, E.; Claesson-Welsh, L. Mechanisms and regulation of endothelial VEGF receptor signalling. Nat. Rev. Mol. Cell Biol. 2016, 17, 611–625. [Google Scholar] [CrossRef]
  341. Hack, S.P.; Zhu, A.X.; Wang, Y. Augmenting Anticancer Immunity Through Combined Targeting of Angiogenic and PD-1/PD-L1 Pathways: Challenges and Opportunities. Front. Immunol. 2020, 11, 598877. [Google Scholar] [CrossRef]
  342. Garcia, J.; Hurwitz, H.I.; Sandler, A.B.; Miles, D.; Coleman, R.L.; Deurloo, R.; Chinot, O.L. Bevacizumab (Avastin®) in cancer treatment: A review of 15 years of clinical experience and future outlook. Cancer Treat Rev. 2020, 86, 102017. [Google Scholar] [CrossRef] [PubMed]
  343. Pogue-Geile, K.; Yothers, G.; Taniyama, Y.; Tanaka, N.; Gavin, P.; Colangelo, L.; Blackmon, N.; Lipchik, C.; Kim, S.R.; Sharif, S.; et al. Defective mismatch repair and benefit from bevacizumab for colon cancer: Findings from NSABP C-08. J. Natl. Cancer Inst. 2013, 105, 989–992. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  344. Lenz, H.-J.; Parikh, A.R.; Spigel, D.R.; Cohn, A.L.; Yoshino, T.; Kochenderfer, M.D.; Elez, E.; Shao, S.H.; Deming, D.A.; Holdridge, R.C.; et al. Nivolumab (NIVO) + 5-fluorouracil/leucovorin/oxaliplatin (mFOLFOX6)/bevacizumab (BEV) versus mFOLFOX6/BEV for first-line (1L) treatment of metastatic colorectal cancer (mCRC): Phase 2 results from CheckMate 9X8. J. Clin. Oncol. 2022, 40, 8. [Google Scholar] [CrossRef]
  345. Vallböhmer, D.; Zhang, W.; Gordon, M.; Yang, D.Y.; Yun, J.; Press, O.A.; Rhodes, K.E.; Sherrod, A.E.; Iqbal, S.; Danenberg, K.D.; et al. Molecular determinants of cetuximab efficacy. J. Clin. Oncol. 2005, 23, 3536–3544. [Google Scholar] [CrossRef]
Figure 1. Currently FDA-approved (red) or promising (black) anti-CRC drugs. HGF: hepatocyte growth factor; c-MET: mesenchymal-epithelial transition factor; VEGF: vascular endothelial growth factor; EGFR: epidermal growth factor receptor; PI3K: phosphoinositide 3-kinase; AKT: protein kinase B, also known as PKB; mTOR: mammalian target of rapamycin; MEK: mitogen-activated protein kinase; ERK: extracellular signal regulated kinase; CTLA-4: cytotoxic T-lymphocyte-associated antigen 4; B7: B7 ligand; PD-1: programmed death-1; PD-L1: programmed death ligand 1; 5-FU: 5-fluorouracil; OXA: oxaliplatin.
Figure 1. Currently FDA-approved (red) or promising (black) anti-CRC drugs. HGF: hepatocyte growth factor; c-MET: mesenchymal-epithelial transition factor; VEGF: vascular endothelial growth factor; EGFR: epidermal growth factor receptor; PI3K: phosphoinositide 3-kinase; AKT: protein kinase B, also known as PKB; mTOR: mammalian target of rapamycin; MEK: mitogen-activated protein kinase; ERK: extracellular signal regulated kinase; CTLA-4: cytotoxic T-lymphocyte-associated antigen 4; B7: B7 ligand; PD-1: programmed death-1; PD-L1: programmed death ligand 1; 5-FU: 5-fluorouracil; OXA: oxaliplatin.
Cancers 14 02928 g001
Figure 2. The changes of drug targets and signaling pathways lead to the development of CRC resistance. IFN-γ: interferon-γ; JAK/STAT: janus kinases/signal transducer and activator of transcription; PTEN: phosphatase and tensin homolog; APC: adenomatous polyposis coli; TP: thymine phosphorylase; OPRT: orotate phosphoribosyl-transferase; UMPS: uridine monophosphate synthetase; UMPK: UMP kinase; CES2: carboxylesterase 2; ABCB1: multi-drug resistance protein 1, MDR1, P-gp; ABCG2: breast cancer resistance protein, BCRP; ABCC1: multi-drug-resistance-associated protein 1, MRP1; UGT1A1: uridine diphosphate glucuronosyltransferase 1A1; CYP3A4: cytochrome P450-3A4.
Figure 2. The changes of drug targets and signaling pathways lead to the development of CRC resistance. IFN-γ: interferon-γ; JAK/STAT: janus kinases/signal transducer and activator of transcription; PTEN: phosphatase and tensin homolog; APC: adenomatous polyposis coli; TP: thymine phosphorylase; OPRT: orotate phosphoribosyl-transferase; UMPS: uridine monophosphate synthetase; UMPK: UMP kinase; CES2: carboxylesterase 2; ABCB1: multi-drug resistance protein 1, MDR1, P-gp; ABCG2: breast cancer resistance protein, BCRP; ABCC1: multi-drug-resistance-associated protein 1, MRP1; UGT1A1: uridine diphosphate glucuronosyltransferase 1A1; CYP3A4: cytochrome P450-3A4.
Cancers 14 02928 g002
Figure 3. The changes of the TME lead to the development of CRC resistance. TAM: tumor-associated macrophages; MDSC: myeloid-derived suppressor cell; CAF: tumor-associated fibroblasts; ODC: ornithine decarboxylase; CSF1: colony-stimulating factor 1; NK cell: natural killer cell; TGF-β: transforming growth factor-β; IL-6: interleukin-6; FGF: fibroblast growth factor; ARG1: arginase I; iNOS: inducible nitric oxide synthase; NO: nitric oxide.
Figure 3. The changes of the TME lead to the development of CRC resistance. TAM: tumor-associated macrophages; MDSC: myeloid-derived suppressor cell; CAF: tumor-associated fibroblasts; ODC: ornithine decarboxylase; CSF1: colony-stimulating factor 1; NK cell: natural killer cell; TGF-β: transforming growth factor-β; IL-6: interleukin-6; FGF: fibroblast growth factor; ARG1: arginase I; iNOS: inducible nitric oxide synthase; NO: nitric oxide.
Cancers 14 02928 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, Q.; Shen, X.; Chen, G.; Du, J. Drug Resistance in Colorectal Cancer: From Mechanism to Clinic. Cancers 2022, 14, 2928. https://doi.org/10.3390/cancers14122928

AMA Style

Wang Q, Shen X, Chen G, Du J. Drug Resistance in Colorectal Cancer: From Mechanism to Clinic. Cancers. 2022; 14(12):2928. https://doi.org/10.3390/cancers14122928

Chicago/Turabian Style

Wang, Qianyu, Xiaofei Shen, Gang Chen, and Junfeng Du. 2022. "Drug Resistance in Colorectal Cancer: From Mechanism to Clinic" Cancers 14, no. 12: 2928. https://doi.org/10.3390/cancers14122928

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

Article Metrics

Back to TopTop