874 Full

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/350586085

Overcoming Resistance to Tumor-Targeted and Immune-Targeted Therapies

Article in Cancer Discovery · April 2021


DOI: 10.1158/2159-8290.CD-20-1638

CITATIONS READS

106 562

6 authors, including:

Mihaela Aldea Fabrice Andre


Institut de Cancérologie Gustave Roussy Institut de Cancérologie Gustave Roussy
143 PUBLICATIONS 1,129 CITATIONS 924 PUBLICATIONS 61,027 CITATIONS

SEE PROFILE SEE PROFILE

Semih Dogan Jean-Charles Soria


Institut de Cancérologie Gustave Roussy MedImmune
26 PUBLICATIONS 1,003 CITATIONS 192 PUBLICATIONS 5,961 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Mihaela Aldea on 04 April 2021.

The user has requested enhancement of the downloaded file.


Review

Overcoming Resistance to Tumor-Targeted


and Immune-Targeted Therapies
Mihaela Aldea1, Fabrice Andre1,2,3, Aurelien Marabelle3,4, Semih Dogan2, Fabrice Barlesi1,5,
and Jean-Charles Soria3,4

aBstRact Resistance to anticancer therapies includes primary resistance, usually related to


lack of target dependency or presence of additional targets, and secondary resist-
ance, mostly driven by adaptation of the cancer cell to the selection pressure of treatment. Resistance
to targeted therapy is frequently acquired, driven by on-target, bypass alterations, or cellular plasticity.
Resistance to immunotherapy is often primary, orchestrated by sophisticated tumor–host–microen-
vironment interactions, but could also occur after initial efficacy, mostly when only partial responses
are obtained. Here, we provide an overview of resistance to tumor and immune-targeted therapies and
discuss challenges of overcoming resistance, and current and future directions of development.

Significance: A better and earlier identification of cancer-resistance mechanisms could avoid the use
of ineffective drugs in patients not responding to therapy and provide the rationale for the admin-
istration of personalized drug associations. A clear description of the molecular interplayers is a
prerequisite to the development of novel and dedicated anticancer drugs. Finally, the implementation
of such cancer molecular and immunologic explorations in prospective clinical trials could de-risk the
demonstration of more effective anticancer strategies in randomized registration trials, and bring us
closer to the promise of cure.

intRoduction the two types of resistance can occur with both tumor and
immune-targeted drugs, it is more common to see acquired
The therapeutic success of tumor- and immune-targeted
resistance to tumor-targeted therapies and primary resistance
therapies has revolutionized the way we understand the biology
to immune-targeted therapies, where there is an important
of tumors and radically changed the therapeutic landscape of
proportion of patients without response to therapy in many
cancers. Spectacular tumor responses were seen with targeted
cancers (3–6). For targeted agents and notably EGFR tyros-
therapy in oncogene-addicted cancers, and unprecedented long-
ine kinase inhibitors (TKI), the Jackman criteria have been
term remissions in some patients treated with immunotherapy.
accepted by the medical community for defining acquired
Despite this immense progress, advanced cancer is ultimately
resistance (7). No such definition exists for immunotherapy,
lethal for most patients due to treatment resistance.
despite ongoing efforts from the Society for Immunotherapy
Cancer resistance is classified into two broad categories:
of Cancer (SITC; ref. 8).
primary resistance, with an early tumor progression, without
Finding predictive biomarkers of resistance has important
prior tumor response, and secondary (acquired) resistance,
implications for cancer care. It may avoid giving an ineffective
which occurs after initial tumor responses (1–3). Although
treatment to patients whose tumors do not respond (9–11) or
guide therapeutic choices to overcome resistance. In the case
of targeted therapy, several biomarkers of resistance are vali-
1
Department of Medical Oncology, Gustave Roussy, Villejuif, France. dated and routinely used. For immunotherapy, the focus has
2
INSERM U981, PRISM Institute, Gustave Roussy, Villejuif, France. 3Paris
been primarily to find biomarkers associated with a positive
Saclay University, Saint-Aubin, France. 4Drug Development Department,
Gustave Roussy, Villejuif, France. 5Aix Marseille University, CNRS, INSERM, predictive value of response [PD-L1, microsatellite instabil-
CRCM, Marseille, France. ity, tumor mutational burden (TMB), etc.]. Therefore, data
Note: Supplementary data for this article are available at Cancer Discovery are still sparse, and there is no approved, clinically validated
Online (http://cancerdiscovery.aacrjournals.org/). resistance biomarker to guide treatment selection (although
Corresponding Author: Jean-Charles Soria, Gustave Roussy Cancer Cam- the absence of PD-L1 expression is de facto rendering some
pus, Villejuif, 94805, France. Phone: 331-4211-4016; E-mail: Jean-Charles. patients with cancer not eligible for anti–PD-L1 therapies).
Soria@gustaveroussy.fr Here, we provide an overview of the main concepts of
Cancer Discov 2021;11:874–99 tumor resistance to cancer cell–, stromal cell–, and immune
doi: 10.1158/2159-8290.CD-20-1638 cell–targeted therapies, applicable across tumor types. We
©2021 American Association for Cancer Research. provide examples of the clinical and preclinical evidence of

874 | CANCER DISCOVERY 10TH ANNIVERSARY ISSUE April 2021 AACRJournals.org

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
Overcoming Resistance to Targeted Therapy and Immunotherapy REVIEW

tailoring treatment beyond progression, the challenges faced Short-Term Adaptation to Targeted Therapies
today, and potential directions of research. The inhibition of feedback loops by targeted therapies can
lead to unintended compensatory overactivation of upstream
pathways. In PIK3CA-mutant estrogen receptor (ER)–positive
Mechanisms of Resistance to Cancer
breast cancer, increased ER activity occurs early upon PI3K
Cell–Targeted Therapy inhibition and limits the antitumor activity of single-agent
A targeted therapy may fail because of drug resistance or PI3K inhibitors (18). Also, mTOR and dual PI3K/mTOR inhib-
lack of drug target, or following inadequate drug exposure. itors suppress mTORC1 and S6K, which normally send inhibi-
Here we will focus on mechanisms of drug resistance when tory signals via the insulin/IGFR and other receptor tyrosine
the target is expressed and validated. An overview of mecha- kinases (RTK). Once these feedback loops are released, resist-
nisms of resistance to targeted therapy is reported in Fig. 1. ance may occur via upstream activation of PI3K, AKT, and
ERK, which circumvents the antitumoral effects of PI3K/AKT/
Primary Resistance to Targeted Therapy mTOR inhibitors. Treatment-induced hyperglycemia induces
Primary resistance is mostly related to the lack of target systemic hyperinsulinemia, which may activate PI3K signal-
dependency or the presence of additional targets, in the con- ing despite continued suppression with PI3K inhibitors. The
text of tumor heterogeneity. Biomarkers of primary resistance ketogenic diet or SGLT2 inhibitors that lower blood glycemia
to a certain drug are (i) insensitive variants of the target [e.g., were shown to inhibit the insulin feedback, decrease mTORC1
EGFR insertions in exon 20 and the majority of EGFR inhibi- signaling in the tumor, and enhance the efficacy of PI3K inhib-
tors (EGFRi) in EGFR-mutated non–small cell lung cancer itors (19). Another example is anti-BRAF monotherapy that
(NSCLC)]; (ii) aberrations of oncogenic pathways connecting fails in BRAFV600E-mutated colorectal cancer due to feedback
to the target (e.g., RAS mutations and EGFR inhibition in activation of EGFR, as opposed to melanoma and NSCLC (20).
metastatic colorectal cancer; refs. 9, 10); (iii) activating muta- Similarly, the modest benefit of sotorasib in KRASG12C-mutated
tions downstream to the target (RB1 mutations and CDK4/6 colorectal cancer as compared with NSCLC is explained by
inhibitors in estrogen receptor–positive/HER2-negative breast EGFR dependency and signaling rebound kinetics (21). MEK
cancer; ref. 12); or (iv) activation of parallel oncogenic path- inhibitors also induce PI3K/AKT activation via EGFR, due
ways (e.g., de novo ALK rearrangement and EGFR mutation in to the cross-talk between the PI3K and RAS pathways (22),
NSCLC and single-agent EGFR/ALKi; Fig. 1A). The latter is and HER2 TKIs induce HER3 upregulation as a compensa-
a very rare situation where it is critical to establish the domi- tory mechanism of PI3K/AKT inhibition (23). The transla-
nant driver if single-agent inhibition is planned. In NSCLC, tional control of gene expression may also mediate resistance,
de novo ALK and EGFR aberrations were found to colocalize exemplified by the formation of the eukaryotic translation
in the same tumor cell population and possibly in the same initiation complex eIF4F, which regulates translation of many
cellular clone in 1.3% of NSCLC cases (13). The dominant oncogenic pathways, including RAS–MAPK and PI3K–mTOR.
driver receptor seems to be more frequently the EGFR muta- The abnormal expression of eIF4F has been linked to resist-
tion, as patients tend to benefit more from EGFRi than from ance to HER2, BRAF, and MEK inhibitors (24, 25). Another
ALKi. The sensitivity to TKI appears to be related to the dif- example of short-term adaptation is nicely illustrated by a pre-
ferential phosphorylation of EGFR and ALK, but the clinical clinical study investigating KRASG12C inhibitors. First-in-class
validity and utility of testing phosphorylation levels of these KRASG12C inhibitors bind only to the inactive guanosine-5′-
proteins is currently unknown (13, 14). Another example is triphosphate state of the oncoprotein. Upon MAPK suppres-
the co-occurrence of EGFR exon 19 deletions with nondisrup- sion by KRASG12C inhibition, some quiescent cells can produce
tive TP53 exon 8 mutations, which have been associated with new, active/­­drug-insensitive KRAS, in an EGFR–SHP2- and
primary resistance to EGFRi in NSCLC (15). AURK-dependent manner (26).

Gene Silencing
Acquired Resistance to Targeted Therapy
The main mechanisms of acquired resistance include: (i)
Despite the identification of a targetable genomic alteration,
on-target resistance alterations; (ii) bypass alterations in the
drugs may be ineffective in the case of gene silencing, when
same pathway or connecting pathways; (iii) alterations that
the target lacks its protein expression. An analysis focused on
cause a phenotypic transformation of the tumor [epithelial-
50 putative drivers comparing whole-exome tumor (somatic)/
to-mesenchymal transition (EMT), small cell transformation
normal (germline) sequencing with whole-transcriptome
of NSCLC, squamous transformation of adenocarcinomas];
sequencing in 1,417 primary tumors revealed that nearly
and (iv) loss of target and target dependency (Fig. 1B; refs.
13% of the somatic single-nucleotide variants (SNV) were
4, 6, 27–29).
­unexpectedly not transcribed as RNA. SNVs with high tran-
scription rates included TP53, PIK3CA, and KRAS, whereas
those with lower transcription rates included ALK, CSF1R, On-Target Mechanisms of Resistance
ERBB4, FLT3, GNAS, HNF1A, KDR, PDGFRA, RET, and SMO On-target mechanisms of resistance include secondary
alterations. Interestingly, the higher the mutational load, the mutations, target amplifications, or alternative splicing of the
higher the number of silenced variants (16). Another example protein kinase mRNA. Following TKIs, “second-site muta-
of gene silencing is posttranscriptional editing with RNA tions” are often seen in the intracellular kinase domain of
silencing by miRNA, RNA interference, and small interfering the protein. They reduce drug affinity for the target by weak-
RNA (17). ening the chemical bonds between the kinase and the drug

April 2021 CANCER DISCOVERY 10TH ANNIVERSARY ISSUE | 875

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
REVIEW Aldea et al.

A Primary resistance to targeted therapy

a. b. Anti-EGFR
EGFR antibody
Wild-type EGFR EGFR exon 20
insertion

C-helix RAS PI3K


Insertion
RAF AKT

MEK mTOR
Inactive Active
ERK

c. EGFRmut ALK fusion d. CDK4/6 inhibitor

P P CDK4/6
P
Cyclin D
RAS PI3K JAK
P G2
RB1 S
RAF AKT STAT
RB1
MEK mTOR E2F M
G1
Free
Cell-cycle
ERK E2F
progression

B Acquired resistance to targeted therapy


EGFRmut
i. Activating Activating + second-site ii.
mutation mutation
Drug STOP
Target ALK fusion

Drug
RAS PI3K JAK
ATP
RAF AKT STAT
Amplification Alternative splicing MEK mTOR

ERK

iii. iv.
Epithelial Mesenchymal Loss of target
EMT

Adenocarcinoma Squamous Small cell


Inhibitory
signal

876 | CANCER DISCOVERY 10TH ANNIVERSARY ISSUE April 2021 AACRJournals.org

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
Overcoming Resistance to Targeted Therapy and Immunotherapy REVIEW

(e.g., EGFRC797S), induce conformational changes of the glioblastoma, resistance to EGFR TKI occurred through the
kinase (gatekeeper mutations, e.g., BCR–ABLT315I, EGFRT790M, loss of extrachromosomal (ec) EGFRvIII DNA, whereas drug
ALKL1196M, ROS1L2026M, RETV804M/L, and TRKAF589L) or cause withdrawal was followed by reemergence of clonal EGFR on
direct steric hindrance to drug binding (solvent-front muta- ecDNA. This unexpected twist illustrates how EGFRvIII levels
tions, e.g., EGFRG796, ALKG1202R, ROS1G2032R, ROS1D2033N, RETG810, are modulated by ecDNA according to the tumor’s needs (54).
TRKAG595R, TRKBG639R, TRKCG623R; xDFG mutations, for exam-
ple, ROS1G2101A/C, TRKAG667S, or TRKCG696A; refs. 30–41). As a Loss or Activation of Mirror Protein. An illustrative example
consequence, some drugs with similar structure are susceptible of “mirror proteins” is PI3K and PTEN proteins, with their
to cross-resistance. Second-site mutations have been prepon- divergent role on PIP3 formation. In PI3K signaling, the trans-
derantly described with TKIs targeting membrane receptors formation of PIP2 to PIP3 is a necessary step for signal trans-
(RTKi; EGFR, cKIT, FGFR, MET, NTRK1–3, and RET) or cyto- duction, which is promoted by PI3K and negatively regulated
solic fusion proteins (ALK, ROS1, and BCR–ABL), whereas only by PTEN. In PIK3CA-mutated breast tumors treated with
exceptionally in case of intracellular protein kinases derived PI3Kα inhibitors, resistance may emerge via a progressive loss
from oncogenic mutations. A single case of a secondary BRAF of PTEN expression, which abrogates the pathway inhibition.
mutation (BRAFL514V) has yet been reported (42). Interestingly, Preclinical studies point out that concomitant use of PI3Kα
the ALK fusion variant seems to affect acquired resistance, and PI3K p110β blockade might reverse resistance (55).
with more ALK resistance mutations in patients with variant
3 than those with variant 1, particularly ALKG1202R (43). Treat- Bypass Resistance Mechanisms
ment with mAbs may result in mutations of the extracellular There are three major oncogenic signaling pathways that
domain of the protein, which impair antibody binding (e.g., drive cell growth and proliferation: the PI3K/AKT/mTOR
EGFR extracellular domain variants after cetuximab and pani- (PI3K pathway), RAS/RAF/ERK (MAPK pathway), and STAT/
tumumab; refs. 44, 45). Mutations affecting the ligand binding JAK pathways. Oncogenic drivers frequently signal through
domain may occur after endocrine therapy, leading to consti- the same pathways: EGFR, IGFR1, FGRF2, and HER2 signal
tutive receptor activation (e.g., ESR1-activating mutations in via the PI3K and MAPK pathways and ALK, ROS1, MET, and
endocrine treatment–resistant breast cancer, or androgen recep- cKIT via all three oncogenic pathways, whereas oncogenic
tor splice variants, such as AR-V7, in enzalutamide/abiraterone- BRAF signals exclusively through the MAPK pathway. Under
resistant prostate cancer; refs. 46, 47). targeted therapy, bypass tracks could eventually lead to the
abnormal activation of the downstream pathway or connect-
Alternative Splicing. Aberrantly spliced proteins can mediate ing signaling pathways. This will ensure a sustained abnormal
resistance via enhanced dimerization (e.g., splice variants of signaling despite continuous target inhibition. Notably, the
BRAFV600E that occur in 13% to 30% of patients with mela- more potent target inhibitors are used, the more frequently
noma failing BRAF inhibitors (BRAFi; ref. 48) or by constitu- bypass tracks are likely to develop (56). Also, despite their
tive protein activation in the absence of ligand (e.g., androgen apparent diversity, bypass tracks follow similar patterns due
receptor splice variants in prostate cancer; ref. 47). to the common signaling of targets. In the case of RTK inhi-
bition, one common event among drivers is the activation of
Target Amplification. Exemplified by EGFR amplification parallel RTK. For instance, in NSCLC, EGFR inhibition may
emerging after EGFR inhibitors, BRAF amplification after result in the emergence of MET, ERBB amplification, FGF2–
BRAFi, or androgen receptor amplification in hormone- FGFR1 loop mutations, IGF1R activation, and fusion events
resistant prostate cancer, target amplification may limit the (28, 57, 58); ALK inhibition could lead to increased EGFR acti-
effectiveness of the drug by exceeding its inhibition capacity vation or KIT amplification (52, 56); MET exon 14 inhibition
(49–51). In addition to the amplified mutant allele, high-level could generate EGFR, HER2 amplifications (59). Examples of
gene amplification of the wild-type allele can also induce downstream pathway reactivation include MAPK pathway
resistance (51, 52). Another process that drives resistance is activation via BRAF, NRAS, and KRAS alterations following
oncogene amplification on extrachromosomal DNA (ecDNA). RTK inhibition in NSCLC (60–63), MAP2K1/2 and MITF alter-
These circular structures lacking centromers undergo an un- ations upon BRAFi in melanoma (64–66), NF1-inactivating
equal segregation during mitosis, which causes increased copy- mutations associated with endocrine therapy resistance in ER+
number amplification with enhanced oncogene expression. breast cancer (67) or PI3K activation via PTEN loss in breast
This results in aggressive tumor behavior, poor prognosis, cancer, and PIK3CA or AKT alterations as acquired resistance
and drug resistance, but the exact underpinning mechanisms to RTK inhibitors in NSCLC, to anti-EGFR mAbs in colorectal
remain largely unknown (53). In experimental models of cancer, or to BRAFi in melanoma (51, 64, 68–70).

Figure 1. A, Examples of primary resistance. (a) Insensitive variants of the target: The EGFR exon 20 insertion promotes a rigid structure that “pushes”
the C-helix inward and prevents it from adopting the outward, inactive conformation. Subsequently, the permanently active EGFR protein preserves
a good affinity for ATP and does not enhance affinity for first-generation EGFR inhibitors, which are inactive in this case. (b) Aberrations of pathways
connecting to the target: In colorectal cancer, the activation of the RAS pathway, located downstream of the transmembrane EGFR receptor, limits the
activity of anti-EGFR mAbs. (c) Concomitant activity of parallel RTK: ALK and EGFR may rarely coexist (1.3% in NSCLC) in the same tumor biopsy and
limit benefit from single-agent treatment. The dominant activity seems to be harbored by the protein with the highest phosphorylation level. (d) Activat-
ing mutations downstream to the target: Deleterious mutations of RB1 encoding the RB1 protein located downstream of the intracellular CDK4/6 protein
impair the activity of CDK4/6 inhibitors and lead to cell-cycle progression through the free form of E2F. B, Examples of acquired resistance. (i) On-target
resistance alterations; (ii) bypass alterations by the activation of parallel RTK; (iii) phenotypic transformation of the tumor, including EMT, small-cell
transformation, and squamous transformation of adenocarcinomas; (iv) loss of target.

April 2021 CANCER DISCOVERY 10TH ANNIVERSARY ISSUE | 877

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
REVIEW Aldea et al.

Among bypass tracks, acquired fusions were recently dis- the original open reading frame and thus regain protein func-
covered at EGFRi failure in EGFR-mutated NSCLC (71). The tion (85, 86). Reversion mutations emerge as a mechanism of
most frequently reported are the RET fusion (46%), followed resistance to platinum agents as well, being more frequently
by ALK (26%), NTRK1 (16%), and FGFR3 (11%; refs. 71–73). found in platinum-refractory/resistant than in platinum-
They were identified at higher rates after third-generation sensitive ovarian tumors. In patients with high-grade ovarian
EGFRi than with first- or second-generation EGFRi (16% vs. carcinoma, the detection of reversion BRCA mutations in the
3%, respectively), when assessed in tissue biopsies by whole- liquid biopsy before the start of rucaparib has been correlated
exome sequencing (WES)/RNA sequencing (RNA-seq; ref. with a lower progression-free survival (PFS; ref. 87). Another
72). As opposed to the classic oncogenic drivers, they fre- example of repairing tumor vulnerabilities is the loss of tar-
quently harbor uncommon 5′ partners, such as STRN for the get. For instance, the loss of PARP1 function by mutations
ALK fusion and CCDC6 and NCOA4 for the RET fusion (71, in the DNA binding domain of PARP or by increased PARyla-
74). Fusions have been reported in other tumors where EGFR tion of PARP, which prevent PARP trapping, drives resistance
inhibition is routinely used, such as colorectal and head and to PARPi (29). Also, primarily ER+ breast tumors may lose
neck cancers (75–77). Acquired fusions are not limited to the estrogen receptor at relapse, which is predictive for poor
EGFR inhibition, as they have also been recently described in response to subsequent endocrine therapy (88).
metastatic ER+ breast cancer as a putative resistance mecha-
nism to endocrine therapy (78).
Other Mechanisms of Resistance
Lineage Plasticity Upregulation of genes encoding P-glycoprotein efflux
Phenotypic switching, or cell plasticity, is a tumor-escape pumps could result in increased efflux of drugs outside the
mechanism that allows cells with the same genotype to tumor cell, if drugs are substrates of the P-glycoprotein. Some
acquire diverse phenotypes in response to adverse tumor examples are PARPi or ALKi (with the exception of alectinib;
microenvironment (TME), such as hypoxia, inflammation, refs. 29, 89).
or exposure to targeted therapy. In the latter, this enables Transient adaptive mutability has recently been described
cells to proliferate independently of initial oncogenic driv- in colorectal cancer cell lines. In response to EGFRi and BRAFi,
ers and promotes tumor resistance. For instance, resistant cancer cells experienced a downregulation of m ­ ismatch-repair
drug-exposed cells have been shown to regain sensitivity to proteins, decreased homologous recombination (HR) profi-
the same drug after drug holidays, which suggests that non- ciency, and increased oxidative stress production. As a result,
genomic mechanisms of resistance are involved (79, 80). Cells cells transiently increased their mutational load as a stress-
could lose their epithelial phenotype and acquire mesenchy- response mechanism. This process was reversed once the
mal characteristics, a process called EMT. This is induced by tumor regained its abnormal growth capacity (90). Currently,
epigenetic modifications, such as upregulation of the histone there is no clinical evidence for this process.
methyltransferase enhancer of zeste homologue 2 (EZH2) or
the RE1-silencing transcription factor (REST). Transformed Epigenetic Changes
cells feature increased migration and invasive potential, being Aberrant epigenetic modifications to the genome occur
highly refractory to targeted therapy. Proposed mechanisms independently of the DNA sequence and could mediate drug
are decreased levels of proapoptotic proteins and increased resistance in addition to mutational processes. Epigenetic
drug efflux due to upregulated ABC-binding cassette trans- changes have been incriminated as key players in the induc-
porters. EMT has been described at failure of EGFRi in tion of dormant, quiescent cells, which are mostly nondivid-
EGFR-mutated NSCLC (81). Lineage plasticity was also ing cells. These cells, called persister cells, constitute a small
shown in breast cancer failing fulvestrant, where ARID1A- fraction of the tumor bulk that acquire stem cell features and
inactivating mutations promoted a phenotypic switch from survive despite high drug exposure or unfavorable micro-
ER-dependent luminal cells to ER-independent basal-like/ environmental conditions. They display a high expression
stem-like cells (82). Cell transdifferentiation, a phenotypic of the KDM5A gene, which encodes a histone demethylase,
transformation from adenocarcinoma to squamous or neu- resulting in reduced H3K4 methylation. The knockdown of
roendocrine carcinoma, has also been reported in 3% to 14% KDM5A was shown to reverse the resistance phenotype and
of patients with EGFR-mutated NSCLC after EGFRi and in diminished the development of drug-tolerant persisters and
around 17% of prostate cancer failing abiraterone/enzalu- their expansion in cell cultures (79).
tamide. Neuroendocrine transformation is preceded by the
concomitant inactivation of TP53 and RB1, but this is not Secretion of Soluble Growth Factors
enough to induce lung cell transformation. Additional fac- Cancer-associated stromal cells may promote drug resistance
tors are necessary, such as MYC, BCL2 overexpression, and by secretion of soluble factors, such as hepatocyte growth fac-
AKT overactivation (83, 84). tor (HGF; ref. 83). HGF has been shown to confer innate resist-
ance to BRAFi in BRAF-mutant melanoma, colorectal cancer, or
Loss of Target or Target Dependencies glioblastoma cell lines (91), and to promote resistance to alec-
Repairing Tumor Vulnerabilities. In tumors with DNA dam- tinib in ALK-positive NSCLC (92). HGF was able to activate
age repair alterations, cells rely on PARP-mediated DNA the MET receptor, even in the absence of MET amplifications
repair to survive. In tumors with frameshift or nonsense or activating mutations. As opposed to the selective ALKi
mutations of BRCA1/2, PALB2, and RAD51C/D treated with alectinib, this phenomenon was not observed with crizotinib,
PARP inhibitors (PARPi), reversion mutations could restore because of its dual inhibition of ALK and MET (92).

878 | CANCER DISCOVERY 10TH ANNIVERSARY ISSUE April 2021 AACRJournals.org

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
Overcoming Resistance to Targeted Therapy and Immunotherapy REVIEW

Increased Mutagenesis tinib (81). There are currently no approved treatments for these
The apolipoprotein B mRNA-editing catalytic polypeptide- tertiary mutations, but very encouraging results were reported
like (APOBEC) enzymes, responsible for DNA cytosine deami- with amivantamab (JNJ-61186372) in the phase I CHRYSA-
nation, are an important source of mutagenesis that fuel LIS trial, where patients with NSCLC failing third-generation
cancer diversity and subclonal evolution (93–95). APOBEC3B EGFR inhibitors achieved partial responses in 10 of 47 cases,
family member has been found to be upregulated in more than including 4 with EGFRC797S mutation (113).
half of all cancers, and the APOBEC mutational signature has In NTRK fusion–positive tumors, the second-generation
been shown to be the most prevalent in cancer after aging sig- TRK inhibitors selitrectinib and repotrectinib are able to over-
natures (96, 97). Moreover, abnormal APOBEC function can come resistance to first-generation TRK inhibitors (entrectinib
generate late driver mutations as subclonal events (94). and larotrectinib) induced by solvent-front substitutions (39,
108). In preclinical models, TRK xDFG mutations (TRKAG667
and TRKCG696) were shown to mediate resistance to both first-
Standard and Investigational and second-generation TRK inhibitors (type I inhibitors that
Drugs for Overcoming Resistance bind to the active conformation state of the kinase), by stabi-
to Targeted Therapy lizing the kinase in an inactive DFG-out conformation. How-
ever, type II inhibitors (cabozantinib, ponatinib, and foretinib)
Overcoming Primary Resistance were able to overcome this type of resistance, as the inactive
Promising results from phase I/II studies are currently avail­ state of the kinase facilitates their access to the allosteric “back
able in the case of EGFR exon 20 insertions, with favor­ pocket” of the DFG motif (115).
able objective response rate (ORR) and duration of response Another strategy is inducing degradation of target pro-
with mobocertinib (TAK-788), a selective inhibitor of EGFR teins, exemplified by selective ER degraders (SERD) for ER
and HER2 exon 20 insertion mutations (98), and with ami- downregulation, or proteolysis-targeting chimeras (PROTAC)
vantamab (JNJ-61186372), a bispecific anti-EGFR–cMET that act in conjunction with the ubiquitin-proteasome sys-
antibody (99). Bispecific antibodies reduce the expression tem (116, 117). Although the technology of PROTACs is still
of EGFR and MET on the cell surface by promoting their maturing, this is highly promising for undruggable, resistant
internalization with subsequent downregulation by lysoso- targets (117).
mal degradation. They also induce tumor cell apoptosis in
a BIM- and caspase-dependent fashion and show antibody- Overcoming Bypass Resistance
dependent cell-mediated cytotoxicity, with natural killer (NK) Bypass tracks warrant a dual blockade of the founder
cells as effectors (100). mutation and the acquired resistance alteration. This strategy
To overcome short-term adaptation mechanisms to tar- has proved to be safe and effective in the phase I TATTON
geted therapies, there is a need for concomitant blockade of trial evaluating savolitinib, a MET inhibitor, in combina-
feedback loops, such as blocking ER in addition to PIK3CA tion with osimertinib in EGFR-mutated NSCLC (118). In
in PI3KCA-mutated ER+ breast cancer (101) or EGFR in individual case reports, acquired fusions have also benefited
addition to BRAF and probably KRAS in BRAFV600E- and from a dual blockade (refs. 56, 61, 63, 71, 73, 119–124; Sup-
KRASG12C-mutated colorectal cancer, respectively (21, 102). plementary Table S2). Also, many drug combinations have
proved to reverse resistance in in vitro studies, but clinical
Overcoming On-Target Resistance validation is pending (Supplementary Table S2). The ongo-
Second-site mutations should be treated with drugs that ing phase II ORCHARD study (NCT03944772) evaluates
have a different structure or a different binding and that are combination strategies for bypass tracks in patients with
not subject to the same resistance alteration as the failing NSCLC failing first-line osimertinib. Bispecific antibodies
drug. In the case of TKIs, therapeutic strategies include the are also promising treatment strategies. In preclinical stud-
switch between drugs of different generations that differ in ies on wild-type and mutant EGFR NSCLC with c-MET
size and binding affinity, the switch between type I and II pathway activation, the bispecific antibody targeting EGFR
ATP-competitive inhibitors, an alternance between allosteric and c-MET (JNJ-61186372) demonstrated antitumor activ-
inhibitors and ATP-competitive inhibitors (Supplementary ity through target inhibition, enhanced a­ ntibody-dependent
Table S1; refs. 32–38, 59, 103–112), and, more recently, the cell-mediated cytotoxicity, and FcγRIIIa binding (125). A
use of bispecific antibodies (113). phase I clinical trial is currently ongoing (NCT02609776).
For example, in EGFR-mutated NSCLC, the EGFRT790M- The prevention of acquired resistance is another valuable
resistant mutation located in the ATP-binding cleft of the strategy. In BRAFV600E-mutated melanoma or NSCLC, MEKi
kinase domain is able to restore the ATP affinity of the kinase. are combined with BRAFi to prevent MAPK pathway reac-
As a result, direct ATP competitors, first- and second-­generation tivation, which frequently occurs under single-agent BRAFi
EGFRi, become ineffective (30). Osimertinib, a third-genera- (64, 126). Interestingly, aside from increasing efficacy and
tion EGFRi, covalently binds to the C797 residue of the ATP- preventing resistance, the combination of MEKi and BRAFi
binding site and bypasses the change of the T790 residue (114). also decreases cutaneous toxicities observed with paradoxical
Nevertheless, tertiary resistance mutations may emerge under MAPK activation under BRAFi monotherapy (127, 128).
osimertinib, mainly at the C797 binding site, and further pre-
vent the covalent binding of osimertinib. Other third-genera- Lineage Plasticity
tion EGFRi with similar structures (e.g., rociletinib, olmutinib, Histologic transformation is treated with chemotherapy or
and narzatinib) are likely to show cross-resistance with osimer- chemoimmunotherapy. Potential novel therapies suggested

April 2021 CANCER DISCOVERY 10TH ANNIVERSARY ISSUE | 879

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
REVIEW Aldea et al.

Hypoxia

Bevacizumab
VEGFA
VEGFR
TKI
T helper HIF1α
17

IL17
Proangiogenic factors CAFs, M2-TAM IL6/STAT3 Acquisition High apelin
VEGF recruitment activation of stemness
APLNR
Angiopoietins
PDGF G-CSF Proangiogenic
FGF1/2 factors IL6 MAPK signaling
HGF blockade
EPC EC
MDSC VEGFA
MAPKi
POLR1D1 amplification High IL8 expression Vasculogenic mimicry Normal vessel co-option

Bevacizumab Sunitinib

VEGFA STOP

IL8
VEGFA

Chr 13 CRC cell


IL8 blockade
proliferation

Figure 2. Potential mechanisms of resistance to antiangiogenic drugs and proposed ways of overcoming resistance. Chr, chromosome; CRC, colorectal
cancer; EC, endothelial cell; EPC, endothelial progenitor cell; FGF, fibroblast growth factor; HGF, hepatocyte growth factor; IL, interleukin; MAPKi, MAPK
inhibitors; M2-TAM, M2 tumor–associated macrophages; MDSC, myeloid-derived suppressor cells; TKI, tyrosine kinase inhibitors.

by preclinical data are targeting factors involved in lineage drugs targeting eIF4E complex formation might overcome
plasticity, such as blocking epigenetic factors (e.g., EZH2 most of the resistance mechanisms arising from BRAF/MEK
inhibitors), antiapoptotic proteins (e.g., navitoclax, a BCL2 inhibition (25).
inhibitor), or stem cell markers (e.g., SOX2 inhibitors; refs.
83, 84).
Mechanisms of Resistance to
Other Antiangiogenic Targeted Therapies
Overcoming resistance to PARPi is a novel field where only Resistance to antiangiogenic drugs is still poorly under-
preclinical data are available so far. In tumors with restored stood. Important interplayers are both the tumor and the
HR or restored replication fork protection, the combina- TME, with treatment-induced hypoxia influencing both
tion of PARPi with ATR inhibitors (ATRi) might overcome compartments. Tumor cells adapt to VEGF blockade and
PARPi resistance by hindering ATR-dependent mechanisms hypoxia by redundancy in angiogenic signaling and activa-
of proper HR/replication fork functionality. Moreover, the tion of compensatory signaling pathways of angiogenesis
association between PARPi and ATRi might resensitize cells (Fig. 2; ref. 129). This explains why antiangiogenic drugs
with combined disruption of BRCA1 and TP53BP1 or REV7 targeting different tyrosine kinases are effective even in
to PARPi (85). sequential use (130). Hypoxia promotes the recruitment of
As multiple oncogenic pathways may use the same key ­cancer-associated fibroblasts (CAF) or M2–tumor-associated
regulators of mRNA translation, the inhibition of these macrophages (M2-TAM), which can drive angiogenesis by
regulators may overcome resistance emerging from the inhi- activation of endothelial cells and their bone marrow–derived
bition of their upstream pathways. An example is the case precursors (131). Stromal cell–mediated resistance to anti-
of eIF4E for RAS/RAF and PI3K pathways, where resistance VEGF therapy includes an increase of T helper (TH) type 17
to BRAF, MEK inhibitors, or combinations, either ERK- tumor infiltration with release of IL17. In murine models,
dependent or ERK-independent, results in increased eIF4E this resulted in higher G-CSF secretion by CAFs and G-CSF–
complex formation, with subsequent signal transduction. It dependent myeloid-derived suppressor cell (MDSC) activa-
is hypothesized that combinations of BRAF signaling with tion and recruitment in the TME. Consequently, there was

880 | CANCER DISCOVERY 10TH ANNIVERSARY ISSUE April 2021 AACRJournals.org

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
Overcoming Resistance to Targeted Therapy and Immunotherapy REVIEW

a secretion of G-CSF–inducible angiogenic factors indepen­ which relate to the cancer cell–intrinsic biology, the impact of
dent of VEGF. Pharmacologic inhibition of TH17 rendered this cancer cell on the phenotype of the TME, and the biology
treatment with VEGF antibodies effective, suggesting that of the host (refs. 2, 3, 5, 143; Fig. 3).
immunomodulation might have a role in overcoming resist-
ance of antiangiogenic drugs (132). Hypoxia also drives Cancer Cell–Intrinsic Mechanisms of Resistance
cancer cells to undergo metabolic changes to face nutrient Neoantigen Loss
starvation and poor oxygenation. It may even promote the The proof that neoantigen presentation drives immuno-
acquisition of stem-like properties, with increased VEGF therapy sensitivity lies in cancers with secondary resistance to
production, enhanced migratory properties, and aggressive- immunotherapy, which lose their initial antigen presentation
ness, thus limiting treatment efficacy (133–135). Secondary (Fig. 3A). Following adoptive T-cell transfer therapy, it was
resistance with anti-VEGF treatment was associated with shown that melanoma may undergo dedifferentiation with loss
increased expression of apelin (APLN) in preclinical xenograft of the melanocyte differentiation antigen (144). Of note, neo-
models of ovarian cancer. Patients with ovarian cancer and antigen loss has been described even in early-stage, untreated
high APLN expression had a significantly shorter disease- melanoma and NSCLC, caused by promoter hypermethylation
free survival than those with a low APLN expression (136). of genes coding neoantigens, or by copy-number loss or chro-
APLN binds to APLN receptor and activates MAPK signaling, mosomal deletions of truncal alterations (145–147).
thus resulting in endothelial progenitor cell proliferation
and new vessel formation (137). Also, in vitro experiments Defective Neoantigen Presentation
showed that resistance to VEGF inhibitors may be medi- Disrupted antigen presentation may affect ICB efficacy.
ated by IL6/STAT3 activation and overcome by IL6 blockade For instance, a comparison between squamous lung car-
(138). Bevacizumab resistance has been associated in two cinoma and adenocarcinoma with similar predicted neo-
cases with a focal amplification on chromosome 13q12.2, antigens showed that squamous lung carcinoma is more
reported in 8.7% of patients with metastatic colorectal cancer. prone to immune escape by downregulating HLA class I
This is explained by the amplification of POLR1D1, which genes (148). Other examples are loss of function (LOF) of
acts as a potential driver and upregulates VEGFA, an impor- β2-microglobulin, another component of MHC class I; dys-
tant promoter of angiogenesis (139). In the case of TKI, regulation of proteins involved in the proteolytic degradation
patients with renal cancer refractory to sunitinib were found of antigens (PSMB5); of transporters that pump the anti-
to have increased IL8 expression. This was confirmed in xeno- genic fragments across the endoplasmic reticulum (TAP1,
graft models, where sunitinib-resistant tumors had high IL8 TAP2, and TAPBP); or disruption of IFN–JAK–STAT signal-
tumor secretion, whereas IL8 blockade resensitized tumors ing, which normally promotes MHC I expression on the cell
to sunitinib (140). surface (refs. 147, 149, 150; Fig. 3B and C). Interestingly,
Sustained tumor angiogenesis may occur via VEGF-inde- only B2M LOF mutations were found to be significantly
pendent pathways, including normal vessel co-option or enriched in nonresponders versus responders in two inde-
vasculogenic mimicry. In glioblastoma, hypoxia induced by pendent cohorts, in contrast to mutations in IFNGR1, JAK1,
bevacizumab may rapidly trigger adaptive mechanisms and JAK2, STAT2, and TAP1/2 found in both nonresponders and
drug resistance via vasculogenic mimicry, a process where responders (150). Nevertheless, B2M mutations were reported
tumor cells acquire endothelial-like properties and develop in 24% of microsatellite instability–high (MSI-H) colorectal
vessel-like structures (141). In response to antiangiogenic cancers, where they did not preclude response to ICB (151).
drugs, primary resistance occurs most likely in tumors with B2M LOF or downregulation has also been shown to cause
alternative neovascularization, whereas in tumors with high acquired resistance to ICB in melanoma, NSCLC, and MSI-H
VEGF dependency, such as in clear cell renal cell carcinoma, colorectal cancer (152–154).
resistance is often acquired (142).
Neoantigen Intratumoral Heterogeneity
Mechanisms of Resistance to T-cell responses are effectively elicited by clonal neoantigens,
as opposed to subclonal neoantigens and intratumoral hetero-
Immune-Targeted Therapies geneity (ITH; ref. 148). This may explain why TMB, which does
Approved immune-targeted therapies for cancer include not distinguish between clonal and subclonal neoantigens,
antagonistic mAbs targeting immune checkpoints such as is only imperfectly correlated with ICB responses. Notably,
CTLA4, PD-1, and PD-L1 with the aim of modulating the patients lacking a durable benefit under anti–PD-1 were found
antitumor T-cell immune response. The mechanisms of to have higher ITH scores than patients with durable clinical
resistance of such T-cell immune modulatory drugs can, to benefits. The acquisition of subclonal neoantigens is induced
some extent, speak to the mechanisms of resistance to bispe- by abnormal APOBEC activity (mutational signature 2) and
cific T-cell engagers and chimeric antigen receptor T (CAR-T) favored by chemotherapy and radiation. This might contribute
cells approved for CD19+ B-cell leukemias and lymphomas. to immune escape by the outgrowth of T cells that act against
Resistance to immune-checkpoint targeted immunotherapies only a limited tumor cell fraction (94, 148, 155).
[immune checkpoint blockade (ICB)] has been reported to be
driven by genomic and nongenomic mechanisms, where the IFN Signaling Defects
tumor–host–microenvironment relationship has an essential Copy-number alterations and mutations within the IFNγ
role. Defects or alterations in the processes orchestrating this pathway have been identified in 75% of melanomas with
relationship could be divided into three broad categories, primary resistance to anti-CTLA4 inhibitors (156, 157).

April 2021 CANCER DISCOVERY 10TH ANNIVERSARY ISSUE | 881

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
REVIEW Aldea et al.

Cancer cell Cancer cell and/or Tumor stroma Host


stroma

Primary Primary Primary Primary


PD-L1 = 0 TGFβ Organ location (visceral) LDH
TMBlo IL6 CD3/CD8 (d)NLR
MSS (CRC, ...) IL8 PD-L1 Eosinophils
EGFRmut (NSCLC) Microbiome
ALK fusion (NSCLC)
STK11mut Secondary
WNT/β-catenin TIM3 upregulated on
upregulation CD4/8
PTEN loss

Secondary
JAK1/2mut (melanoma)
IFNGR1/2 (melanoma)
IRF1mut
B2Mmut
Loss of heterozygosity
Subclonal elimination

A Neoantigen loss B Defects of the antigen processing C Abnormal IFNγ signaling


machinery

Gene silencing
T cell T cell
MHC-I
INFγ
MHC-I with receptor
PD-L1
antigen STOP STOP

MHC-I JAK STOP


B2M JAK
Tumor cell Antigen Proteasome
+
Tumor STAT
cell IFN genes
TAP
A T G C GA C T
Dedifferentiation

D Aberrant genomic signaling E Coinhibitory checkpoints F Immunosuppressive TME

Immunosuppressive TIGIT Treg


cytokines
WNT WNT BTLA
Frizzled
MDSC
LAG3

TIM3

M2-TAM TGFβ
CTLA4
β-catenin
PD-1 IL10 IL35

Figure 3. Most commonly described mechanisms driving resistance to immunotherapy. A, Neoantigen loss. B, Defects of the antigen processing
machinery. C, Abnormal IFNγ signaling. D, Aberrant genomic signaling. E, Coinhibitory checkpoints. F, Immunosuppressive TME. CRC, colorectal cancer;
IL, interleukin; IFNy, interferon gamma; MDSC, myeloid-derived suppressor cells; MHC I, major histocompatibility complex I; M2-TAM, tumor-associated
macrophages type 2; MSS, microsatellite-stable; mut, mutation; Treg, regulatory T cells.

882 | CANCER DISCOVERY 10TH ANNIVERSARY ISSUE April 2021 AACRJournals.org

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
Overcoming Resistance to Targeted Therapy and Immunotherapy REVIEW

Less frequently, LOF of JAK1 and JAK2 has also been shown the tumor (156). In a retrospective study of 551 patients with
to abrogate IFNγ signaling and IFN-induced inhibition of oncogene-addicted NSCLC receiving ICB monotherapy, some
growth, described in both primary and acquired resistance patients experienced tumor regression, but generally the effi-
to ICB (152, 156, 158, 159). Deleterious JAK1 and JAK2 cacy was lower as compared with the KRAS group. Notably,
mutations are enriched in tumors with MSI, notably in patients with ALK-positive NSCLC did not obtain any objec-
endometrial cancer (35% of immune-naïve patients had JAK1 tive response (178). Immunotherapy responses seem to differ
mutations; refs. 158, 160, 161). Interestingly, JAK2 and type between EGFR mutation subtypes when compared with EGFR
I IFN genes are frequently codeleted with CDKN2A, due to wild-type patients, with lower ORR and overall survival for
their relative proximity on chromosome 9p, suggesting that EGFR exon 19 deletion and similar outcomes for EGFRL858R
CDKN2A loss may be a biomarker indirectly associated with substitution. A possible explanation is the higher TMB
immunotherapy resistance (162, 163). Patel and colleagues observed among EGFRL858R-mutated tumors as compared
identified 13 IFNγ-induced genes and three TNFα-induced with EGFR exon 19–mutated tumors (179). Also, oncogenic
genes that could foster resistance to ICB (149). Also, the LOF signaling, notably the RAS–MAPK and PI3K–mTOR path-
of the APLNR gene, which normally encodes a rhodopsin-like ways, stimulates the eIF4F eukaryotic translation initiation
receptor with roles in blood pressure regulation, has recently complex, which regulates the translation of STAT1 mRNA
been shown to reduce immune response by IFNγ modulation and mediates the induction of PD-L1 expression (180).
via JAK1 (149). IFN secretion might also be impaired when Tumors with increased levels of chromosomal instability
STING and/or cGAS expression is silenced through epige- activate the cGAS–STING pathway through chronic release
netic hypermethylation processes (164). of cytosolic DNA, but they may escape the immune system
by suppressing downstream type I IFN signaling and instead
Oncogenic Signaling with T-cell Exclusion upregulating the alternative NF-κB pathway. This leads to
Tumors lacking CD8+ T cells are termed “cold tumors” paradoxical protumorigenic effects and the recruitment of
and have been associated with a lack of response to ICB. Such immunosuppressive cells (164, 181). A similar immune eva-
tumors do not display an inflammatory, chemokine signature sion may occur in HR-deficient tumors, where high levels
and generally lack negative regulators, such as PD-L1, regula- of DNA damage may activate the ATM–TRAF6 alternative
tory T cells (Treg), or indoleamine-2,3-dioxygenase (IDO; ref. STING pathway, resulting in IL6 and TGFβ production, with
165). The lack of effector T-cell infiltration could be mediated M2-TAM and Treg recruitment (182, 183).
by specific oncogenic signals, such as tumor-intrinsic active An interesting biomarker is polybromo-1 (PBRM1) gene
β-catenin signaling, loss of PTEN, STK11/LKB1, or KEAP1 aber- LOF mutations, clinically validated as a predictor of ICB
rations, or because of IFN signaling defects (JAK1/2 and B2M response in clear cell renal cell carcinoma in an independent
LOF mutations; refs. 166, 167). Active β-catenin signaling was cohort of patients treated with nivolumab within a rand-
shown to promote T-cell exclusion by failure of T-cell priming, omized clinical trial (184). Nevertheless, in a retrospective
with resistance to anti–PD-L1/anti-CTLA4 mAb therapy in a analysis of 2,764 patients with NSCLC who received ICB as
melanoma mouse model (166), as well as resistance to T-cell a second or later line of therapy, patients with PBRM1 muta-
adoptive transfer and vaccination (168). The activators of the tions (n = 84 patients, 3%) had worse survival than patients
β-catenin pathway could be activating mutations of CTNNB1, with wild-type PBRM1, despite associating with a significantly
overexpression of specific WNT ligands or Frizzled receptors, higher TMB (185).
or inactivating mutations of pathway inhibitors, such as the
AXIN1 gene (Fig. 3D; ref. 166). Biallelic loss of PTEN was Resistance to Lysis of Metastatic Tumor Cells
reported to mediate resistance to ICB (169, 170), possibly In a prostate cancer model, both early and metastatic
by inducing an increased expression of immunosuppressive tumors induced cytotoxic T cells after irradiation, but only
cytokines and the recruitment of suppressive immune cells via primary cancer cells were preponderantly eliminated by
VEGFR (171–173). STK11/LKB1 and KEAP1-mutant or -defi- immunotherapy, whereas metastatic cells were resistant to
cient NSCLC has been associated with impaired IFN signaling cytotoxic T cell–induced cell lysis. Despite IFN-induced MHC I
(167). Patients with NSCLC presenting STK11/KEAP1 aberra- expression on metastatic cells, these cells maintained resist-
tions had inferior PFS with combined chemoimmunotherapy ance to lysis (186).
compared with those without, and there has been no observed
benefit from the addition of ICB to chemotherapy (174). The
Stromal Mechanisms of Resistance
prognostic effect of STK11–KEAP1 was further confirmed in a
large real-world data cohort of patients with advanced NSCLC, Location of Metastasis
but not its predictive value (175). Notably, various KEAP1- In patients with melanoma, the presence of visceral metas-
driven comutations (STK11, SMARCA4, and PBRM1) have been tasis was shown to be associated with primary resistance
associated with a lack of response to ICB despite a high TMB to anti–PD-L1, as opposed to lung, lymph node, and soft-
in lung adenocarcinoma (176). In KRAS-mutant NSCLC cell tissue metastases (187). In patients with both melanoma
lines with STK11/LKB1 loss, STING expression was markedly and NSCLC, patients with liver metastases had worse clinical
silenced by hyperactivation of DNMT1 and EZH2, thus facili- outcomes with ICB, compared with patients without liver
tating immune escape and explaining why such tumors are metastases (188, 189). Liver metastases exhibit an immu-
often refractory to ICB (177). nosuppressive TME, and it has been suggested that they
Certain clonal driver alterations associate with an increased may even promote systemic immunosuppression (188). In
immunosuppressive TME and low CD8+ T cell/Treg ratios in melanoma, patients with liver metastases had higher levels of

April 2021 CANCER DISCOVERY 10TH ANNIVERSARY ISSUE | 883

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
REVIEW Aldea et al.

eotaxin 2, interferon gamma-induced protein 10 (IP10) and results in increased release of H+ ions into the TME. The
IL8, MMP8 and HIF1α, as compared with patients without abnormal tumor vasculature is unable to effectively clear H+
liver metastases (188). Brain metastases also display an immu- ions, which promotes an acidic tumor milieu. Consequently,
nosuppressive TME, with low tumor-infiltrating lympho- TILs and cytokine production are inhibited, while the accu-
cytes (TIL), increased M2-TAM and neutrophils, inhibition of mulation of immune-suppressive cells is favored, thus lead-
dendritic cell maturation, TH1, and leukocyte extravasation ing to immune resistance (193, 200). Another mechanism
signaling pathways, as well as a significant reduction in the is the production of the tryptophan-metabolizing IDO by
proinflammatory cell adhesion molecule vascular cell adhe- tumor cells in response to inflammation. IDO has immuno-
sion protein 1 (VCAM1), responsible for leukocyte adhesion suppressive functions that limit effector T-cell activity and
to inflammatory sites (190–192). engage mechanisms of immune tolerance (165).

Expression of Multiple Inhibitory Regulators Immune Heterogeneity


In addition to PD-1 and CTLA4 expression, infiltrating In the TRACERx and LATTICe-A cohorts of primary lung
CD8+ T cells may have increased expression of T-cell immu- adenocarcinoma, the presence of immune cold regions within
noglobulin mucin-3 (TIM3), lymphocyte-activation gene 3 a tumor was shown to be an independent predictor of cancer
(LAG3), and B and T lymphocyte attenuator (BTLA; Fig. relapse, even in the case of an average increased immune infil-
3E). These regulators are also expressed by other immune tration of the tumor. The comparison between cold and hot
cells, such as Tregs or cells of innate immunity (193). The regions reflected different cancer subclones, possibly due to
upregulation of coinhibitory molecules has been described immunoediting (201). An increased heterogeneity was found
as a mechanism of acquired resistance to single-agent anti– in the immune microenvironment of different regions of the
PD-1/PD-L1, causing an impairment of T cells to respond to same tumor in nearly one third of early NSCLC tumors (145).
polyclonal activation. It has been shown that multiple inhibi-
tory checkpoints were acquired gradually, with PD-1 as an Characteristics of the Host
early event and LAG3/BTLA expression as a late event. This Age and Sex
mechanism could drive resistance to ICB even in PD-1 high- Growing evidence points out that younger and female
expressing tumors (194, 195). patients respond less to immunotherapy than older and
male patients. It has been shown that younger and female
Exhaustion of Cytotoxic T Cells
patients are prone to stronger immunoediting early in their
The chronic exposure of cytotoxic T cells to high antigen disease evolution, with depletion of potentially immunogenic
load might severely impair their inflammatory and cytotoxic mutant peptides and MHC presentation of less immuno-
function, from a gradual loss of effector function to an genic, more invisible peptides to the immune system (202).
exhausted phenotype (194). This exhausted phenotype can
present abnormal DNA methylation (196), high PD-1, and Microbiota
LAG3 expression (148).
Specific gut bacteria are associated with ICB response and/
or toxicity. For instance, in patients with melanoma treated
Immunosuppressive Microenvironment
with ICB, baseline gut microbiota enriched with Faecalibacte-
Tregs have important roles in preventing autoimmune rium and other Firmicutes was associated with enhanced anti-
reactions by secretion of immunosuppressive cytokines, such tumor responses (203, 204). Also, a molecular mimicry was
as TGFβ, IL10, and IL35. When present in the TME, they recently observed between tumor MHC class I antigens and
restrict T-cell responses (197). TGFβ may also be released by an enterococcal prophage, which correlated with long-term
cancer cells or stromal fibroblasts (198). It has been found responses under anti–PD-1 therapy in renal and lung cancers
to be increased in the bone marrow from bone metastasis (205). The composition of gut microbiota may affect immune
compared with the normal bone, which might explain the responses of distant lesions, especially via the blood circula-
low benefit of ICB in metastatic castration-resistant prostate tion of microbial metabolites produced in the colon through
cancer (mCRPC) with bone-predominant disease. Moreover, bacterial fermentation of dietary fibers. High blood levels of
in mice with bone mCRPC, the association of anti-CTLA4 short-chain fatty acids, such as butyrate and propionate, have
and anti-TGFβ increased TH1 cells and decreased Tregs in been shown to associate with CTLA4 blockade resistance and
the tumor-infiltrating T-cell population. The combination with an increased proportion of Tregs. In patients treated
of drugs resulted in an improved tumor response and over- with ipilimumab, high blood levels of butyrate limited the
all survival of mice in comparison with either drug alone accumulation of memory and ICOS+ CD4+ T cells, as well as
(199). Innate immune cells, MDSCs, and M2-TAM have also IL2 impregnation (206).
emerged as major regulators of immune responses in cancer,
and they have been shown to negatively affect the activity Host–Tumor Interactions
of ICB, adoptive T-cell therapy, or dendritic cell vaccination Blood Immune Cells
(Fig. 3F; ref. 3). Relative eosinophil count (REC) <1.5% was associated with
a worse survival in patients with melanoma treated with pem-
Metabolic Pathways brolizumab, as compared with REC ≥ 1.5% (187). Interestingly,
Under aerobic conditions, tumor cells rapidly process glu- lymphopenia per se is not associated with primary resist-
cose into lactic acid, a process called the Warburg effect. This ance to anti–PD-L1 but rather the proportion among other

884 | CANCER DISCOVERY 10TH ANNIVERSARY ISSUE April 2021 AACRJournals.org

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
Overcoming Resistance to Targeted Therapy and Immunotherapy REVIEW

cells (187, 207), notably neutrophils, within a neutrophil- Increase Tumor Visibility to the Immune System
to-lymphocyte ratio (NLR; refs. 208–210). The negative Making the tumor visible to the immune system is essential
impact of NLR seems to be limited to patients with stable for T-cell activation and migration into the TME. One major
disease (SD) and not to patients with progression or objective strategy is the induction of “immunogenic cell death” by
response as best response (211). Also, the proportion of neu- tumor cytotoxic therapies, including chemotherapy, irradia-
trophils among total leukocytes minus neutrophils (dNLR), tion, targeted therapies, epigenetic drugs, or oncolytic viruses
combined with lactate dehydrogenase (LDH) in the Lung or peptides (193, 224). Following cell death, tumor antigens,
Immune Prognostic Index (LIPI) score, has been associated proinflammatory cytokines, damage-associated molecular
with primary resistance to anti–PD-L1 (212). patterns (DAMP), calreticulin, and ATP are released in the
TME (Fig. 4A). This may favor antitumor immune responses
Blood Cytokines by recruiting antigen-presenting cells (e.g., dendritic cells),
IL6 has prognostic value, with high IL6 levels being cor- and inducing their maturation (e.g., HLA-I, CD80/86 upregu-
related with a worse survival in patients with metastatic mela- lation) and subsequent T-cell activation (e.g., upregulation of
noma treated with ICB or chemotherapy (213). IL8 has been costimulatory checkpoints; refs. 224, 225). DNA lesions may
related to the level of infiltrating myeloperoxidase-positive activate the cGAS–STING pathway, which can increase the
(MPO+) and/or CD15+ monocytes and neutrophils in tumors immunogenicity of the tumor by activating the type I IFN
(214). IL8-high patients, as defined by serum IL8 values above pathway (164). In addition, some chemotherapies can deplete
23 pg/mL, represent around 25% to 30% of patients with some suppressive immune cells such as Tregs or MDSCs
advanced cancer. Interestingly, IL8 serum levels are independ- (193). Chemoimmunotherapy improved survival outcomes
ent from PD-L1 tumor expression by IHC and TMB. in comparison with chemotherapy alone in advanced lung
cancer, regardless of PD-L1 expression (226). Interestingly, it
LDH seems to reduce the risk of early progression seen with ICB
Serum LDH value above normal threshold is a strong pre- alone. It is unknown if a shorter administration of chemo-
dictor of primary resistance to ICB (187, 209, 215, 216). LDH therapy would be more effective when combined with ICB in
leads to accumulation of lactates in the TME, which leads order to prevent lymphopenia or the generation of subclonal
to modifications of immune cells toward a tolerogenic phe- antigens. Possibly, the next step will be the administration
notype (217) and negatively affects the type I IFN pathway of antibody–drug conjugates (ADC) in combination with
by inhibition of retinoic acid–inducible gene 1 (RIG1; refs. ICBs, as they would spare lymphocytes. Successful examples
218, 219). of combining antiangiogenic drugs with immunotherapy
Of note, patterns of primary resistance differ among are especially seen in renal cell carcinoma (e.g., axitinib plus
tumor types even for the same immunotherapy (Supple- pembrolizumab, axitinib with avelumab), where in 90% of
mentary Table S3; refs. 145, 157, 158, 167, 174, 178, 185, cases there is a deregulation of the VHL gene, which regulates
220, 221). Tumor types with a low likelihood of response HIF1α and HIF2α (227). Also, antiangiogenic drugs have
to ICB, for instance, sarcoma or prostate cancer, are usually been shown to modulate the expression of inhibitory check-
cold tumors with a low TMB. Responses may be exception- points on CD8+ T cells and the levels of Treg-infiltrating
ally seen, for instance in sarcomas with B cell–rich tertiary tumors (228, 229). Radiotherapy has been shown to induce
lymphoid structures (221), or MSI-H or CDK12 biallelic an abscopal effect, first described in a case report in 2012
alterations in prostate cancer (222, 223). In these cases, pre- (230), but this effect has not been reproduced in large studies.
dictors of response should be searched. Conversely, in tumor Concerns were raised on the effectiveness of single-site irra-
types with a high likelihood of response to ICB, such as diation, as it fails to address intertumoral and immune het-
melanoma, Hodgkin lymphoma, and MSI-H tumors, predic- erogeneity (225, 231). New strategies of combining ICB with
tors of resistance should be interrogated, and these usually multiple site irradiation are being tested (225). It has been
consist of defects of antigen presentation and IFN signaling. suggested that nodal irradiation should be avoided when
Concerning acquired resistance, defects of antigen presenta- used to enhance ICB effects, as it may compromise T-cell
tion, loss of antigen, IFN signaling defects, and overexpres- proliferation and activation (232). Furthermore, radiation-
sion of coinhibitory checkpoints seem to be the central induced STING activation may cause resistance by recruiting
players in progression, without evident differences among MDSCs via the CCR2 pathway, for which a possible abrogat-
tumor types at the actual state of knowledge (144, 146, 147, ing strategy might be the use of anti-CCR2 antibodies (233).
152, 153, 159, 170, 195). Oncolytic viruses are used to preferentially infect cancer cells,
where they trigger immune responses by viral replication,
cancer cell release of DAMPs, pathogen-associated molecular
Overcoming Resistance patterns, and type I IFNs, with subsequent increase of TILs
to Immunotherapy and even induction of abscopal effects (234–236). Oncolytic
peptides may act by mitochondrial membrane permeabiliza-
Overcoming resistance to immunotherapy is guided by sev- tion, which triggers a caspase-independent necrosis (237).
eral main pillars: (i) increase tumor visibility to the immune
system, (ii) enhance T-cell infiltration, (iii) remove barriers
from T-cell intratumoral trafficking, (iv) enhance immune Enhance T-cell Infiltration
system function, (v) address genomic and epigenetic abnor- In addition to poor tumor immunogenicity, the lack
malities, and (vi) adapt to the host (Fig. 4). of T-cell infiltration could be caused by inefficient T-cell

April 2021 CANCER DISCOVERY 10TH ANNIVERSARY ISSUE | 885

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
REVIEW Aldea et al.

A Increase tumor visibility B Enhance T-cell infiltration C Remove TME barriers


Cytotoxic therapies
Chemotherapy Antigens Abnormal
Radiotherapy vasculature,
Targeted therapy Cytokines hypoxia
Oncolytic virus DAMPs

CAR-T cell
Antiangiogenic drugs

Cancer cell Calreticulin


ATP Vasculature
normalization
Immunogenic cell death
TCR
Immune cell

D Enhance T/NK cell function E Unfavorable genomics/epigenetics F Adapt to the host

+ + STOP
A T G C G A C T
Manipulate microbiota

Anti-TIGIT TLR9 agonist Anti-TGFβ


Anti-LAG3 proinflammatory Anti CSF1R Abnormal IFN Abnormal β-catenin
Anti-VISTA cytokines Anti-CCR5 signaling NK agonists?
Anti-TIM3 Anti-CXR2/CXCR Older
INFγ Biallelic PTEN loss male
STING agonists PI3K inhibitors? Younger
Specific oncogenes female
?

Effector T cell NK cell M2-TAM


Epigenetic changes
Epigenetic drugs Drug combinations? Monotherapy?
OX40 agonist

Figure 4. Overcoming resistance to immunotherapy. A, Increase tumor visibility. B, Enhance T-cell infiltration. C, Remove TME barriers. D, Enhance
T-cell/NK-cell function. E, Unfavorable genomics/epigenetics. F, Adapt to the host.

priming or the lack of T-cell attraction. STING agonists, mCRPC, ipilimumab failed to induce durable responses in
which activate the STING pathway, promote tumor secre- most patients despite being able to increase T-cell infiltra-
tion of type I IFN and proinflammatory cytokines (164). tion in the tumors. This was caused by the overexpression of
The role of chemokines on ICB to attract intratumor T cells compensatory inhibitory pathways, PD-L1 and VISTA (244).
is being investigated in early-phase clinical trials, as certain A similar escape mechanism was observed in patients treated
chemokines (e.g., CXCL13) may mediate the recruitment of T with dendritic vaccines, which led to a significant increase
cells into the TME (238). The induction of tertiary lymphoid in PD-L1 expression (245). This provides the rationale that
structure formation in tumors is being investigated using bringing T cells into the tumor must be accompanied by a
cytokines, chemokines, antibodies, antigen-presenting cells, blockade of emerging inhibitory pathways.
or synthetic scaffolds (239). Other strategies include vaccines
and adoptive cellular therapies (Fig. 4B). Engineered T cells Remove Barriers from Intratumoral
with CAR or transgenic T-cell receptors (TCR) have shown T-cell Trafficking
modest activity in solid tumors so far, possibly explained Leaky blood vessels coupled with defective lymphatic drain-
by tumor heterogeneity, which limits the effectiveness of age cause an elevated interstitial pressure in the TME, which
strategies targeting one single antigen, the lack of cancer- may impair drug and immune cell entry. Moreover, TME cells
specific targets for CARs, and the difficulty in matching HLA exposed to hypoxia produce various cytokines with immu-
antigen for TCRs (240–242). Promising results were seen in nosuppressive effects (VEGF, TGFβ, and prostaglandin E),
patients with neuroblastoma treated with CAR-T cells target- whereas tumor cells might acquire highly aggressive stemness
ing di­sialoganglioside GD2, showing the potential of CAR-T features (193, 246). In addition to promoting immunogenic
cells in solid tumors, when the target is adequate (243). cell death, antiangiogenic drugs may normalize tumor vascu-
lature and increase T-cell infiltration in the TME (Fig. 4C).
Why Combination with ICB Is Important The spatial distance between T cells and tumor cells might
Once non–T cell–inflamed tumors have T-cell infiltrates, be minimized by bispecific antibodies that bind T cells with
IFNγ signaling will be activated, with subsequent expression tumor cells. For instance, an early success was seen with the
of inhibitory mechanisms, such as PD-L1. For instance, in bispecific antibody blinatumomab, approved in the t­ reatment

886 | CANCER DISCOVERY 10TH ANNIVERSARY ISSUE April 2021 AACRJournals.org

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
Overcoming Resistance to Targeted Therapy and Immunotherapy REVIEW

of B-cell leukemia, which acts by binding CD3 on T cells and Conversely, PD-1 blockade resistance caused by JAK1/2 and
CD19 on B cells (247). B2M inactivating mutations has been successfully reversed in
murine models by TLR9 agonists and CD122-preferential IL2
Enhance Immune System Function pathway agonists, respectively (255).
Enhancing immune functions focuses on promoting the The potential synergy between epigenetic drugs and ICB is
infiltration and function of effector T cells and antigen- also being investigated by addressing the epigenetic modulation
presenting cells on inhibiting immunosuppressive cells such of tumor neoantigens, chemokines, or inhibiting activation-
as Tregs and MDSCs, or modifying cytokine production or induced cell death of T cells (193). For example, promising results
cellular metabolism (Fig. 4D; ref. 193). were seen in the ENCORE-601 phase Ib/II trial in patients with
Boosting effector T cells can be achieved by concomitant melanoma, NSCLC, or mismatch repair–proficient (pMMR)
blockade of inhibitory regulatory checkpoints or by activat- colorectal cancer treated with pembrolizumab in association
ing stimulatory pathways (193–195). The combination of with a histone deacetylase inhibitor who obtained tumor
anti–PD-1/PD-L1 agents with anti-CTLA4 mAbs was the first responses even after prior immunotherapy failure (257).
proof of increased efficacy compared with monotherapy in
several cancer types, albeit with the cost of increased toxic- Adapt to the Host
ity (5). Ongoing clinical trials investigate the combination Younger and female patients may be better candidates for
of anti–PD-1/PD-L1 or anti-CTLA4 with inhibitors of other drug combinations, as opposed to older and male patients
immune-checkpoint regulators, such as LAG3 or TIGIT (5, with higher chances of response to ICB (202). Manipulat-
193). Recently, the phase II CITYSCAPE trial evaluating ing microbiota is undergoing investigation for enhancing
the anti-TIGIT mAb tiragolumab plus the PD-L1 inhibitor immune responses or managing immune-related adverse
atezolizumab in patients with NSCLC with positive PD-L1 events (Fig. 4F; ref. 258). Studies investigating ICB combined
showed an improved ORR and PFS for the combination with LDH inhibitors might be of interest for patients with
therapy, as compared with atezolizumab alone. Efficacy was high LDH levels, as experimental data showed antitumor
higher in the high–PD-L1 subgroup (248). Phase III trials are effects of LDHA inhibitors by inducing reactive oxygen spe-
ongoing (SKYSCRAPER-01 and SKYSCRAPER-02). cies and apoptosis (259).
T-cell function can also be enhanced by removing immu-
nosuppressive signals. In case of immunosuppressive TME,
inhibitors of M2-TAM or ICB combinations with anti-TGFβ Current Challenges with Tumor-
therapies may be used (156). and Immune-Targeted Therapies
Both innate immunity and adaptive immunity may be stim- and Future Perspectives
ulated by proinflammatory cytokines (249–251), as nicely illus-
trated in a phase I trial investigating an engineered IL2 receptor Improving the Characterization of Resistance
agonist (bempegaldesleukin) in combination with nivolumab Tumors may contain cells that follow different evolu-
in immunotherapy-naïve patients with advanced solid tumors. tionary tracks, including drug-sensitive cells, drug-tolerant
Longitudinal biopsies revealed increased T-cell infiltration and persister cells, fully resistant cells, and partially sensitive or
clonality in patients responding to therapy, with no expansion weakly resistant cells (79, 260). Tumor response is dictated
of Tregs in the tumor (249). In preclinical models, IL15 agonist by the tumor fraction harbored by these cell subtypes, which
or recombinant human IL2 used in combination with ICB was makes a deeper tumor characterization essential for antici-
shown to be more effective than either monotherapy by stimu- pating resistance and improving treatment choices (Fig. 5A
lating both CD8+ T cells and NK cells (250, 251). and B). We are currently limited in our ability to distinguish
Innate immunity may be stimulated by intratumoral between the clonal expansion of preexistent resistant cells
administration of Toll-like receptor 9 (TLR9) agonists (CpG and acquired resistance (e.g., EGFRT790M might also be found
olidonucleotides, e.g., SD-101; refs. 252, 253), double-stranded in patients with treatment-naïve NSCLC; ref. 1). Treatment
RNA viruses triggering TLR3 and RNA helicases (254), or by effectiveness may differ in these cases, with earlier treatment
STING/cGAS activation (164). This might be a rational strat- failure when resistant cells are present at baseline. To improve
egy for tumors that lose MHC, as MHC I–deficient cells are the characterization of tumor heterogeneity, single-cell analy-
unable to present the antigen to effector T cells, but they are sis emerged as a technological breakthrough that may reveal
susceptible to NK cell–dependent cytotoxicity (255). clonal relationships with high precision (261, 262). However,
the use of single-cell analysis in clinical practice is still limited
Address Genomic and Epigenetic Abnormalities by its cost and duration of sample processing.
In mice bearing melanomas with PTEN loss, the addi- Liquid biopsy, increasingly used in routine practice, is able
tion of a PIK3β inhibitor to anti–PD-1 blockade was shown to depict tumor heterogeneity and to reveal the spatial hierar-
to increase intratumoral T-cell infiltration and to improve chy and temporal distribution of cancer alterations. It effec-
tumor control as compared with each treatment alone (171). tively identifies resistance alterations (71, 75), and it may
Abnormal IFN signaling is addressed by combinations perform better than tissue in the identification of complex,
of ICB with IFNγ or STING agonists (Fig. 4E; ref. 156). polyclonal alterations (263, 264). Also, liquid biopsy may
JAK inhibition was also tested in a phase I trial, but proved be used to monitor resistance during treatment or minimal
unsuccessful when combined with pembrolizumab, being residual disease after curative treatments. However, when cir-
associated with reduced peripheral T-cell activation and no culating tumor DNA shedding is low (e.g., low tumor burden,
significant changes in intratumoral Treg infiltration (256). isolated central nervous system progression, lung-only ­disease,

April 2021 CANCER DISCOVERY 10TH ANNIVERSARY ISSUE | 887

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
REVIEW Aldea et al.

A Tumor cell Primary


composition resistance Resistance

Sensitive cells

Partially sensitive/
weakly resistant cells Acquired
resistance
Mutational and
Fully resistant cells epigenetic changes
Drug-tolerant
persister cells
Drug Time
start

Natural evolution (mutagenesis, epigenetic changes)

Tumor composition in time

B
Early stage Advanced stage Drug resistance Complex resistance

Mutagenesis Mutagenesis Mutagenesis

Tumor mutational burden, subclonal diversity, intratumoral heterogeneity, genomic complexity

Block clonal Block multiple drivers; Block targetable Selective delivery of


drivers earlier prioritize clonal >> resistance alterations; unselective drugs
subclonal, rank alterations use drug combinations if
needed

Inhibit mutagenesis
Promote mutagenesis
Adapt treatment: clonal and subclonal
Eliminate persister cells
monitoring with sequential therapy
Monitor and treat
minimal residual disease

PREVENT Resistance TREAT

Drug selectivity

Figure 5. A, Evolutionary tracks of resistance. B, Proposed strategies to defer and overcome tumor resistance.

888 | CANCER DISCOVERY 10TH ANNIVERSARY ISSUE April 2021 AACRJournals.org

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
Overcoming Resistance to Targeted Therapy and Immunotherapy REVIEW

or certain histologic subtypes such as urological cancers; refs. in oncogene-addicted cancers with increased central nervous
265–267), more sensitive techniques are required (e.g., whole- system (CNS) tropism (275–277). Thus, it would be interest-
genome sequencing of circulating free DNA; ref. 268). ing to investigate the role played by deferring resistance when
In the case of immunotherapy, nongenomic mechanisms comparing drugs that have similar CNS activity.
of resistance need to be addressed in conjunction with the Other strategies to prevent resistance are the simultaneous
genomic profile, where the main limitation is the lack of inhibition of co-occurring drivers, proved to be effective and
validated tools to provide a comprehensive view (2). An safe in the I-PREDICT trial (278), or the concomitant block-
interesting approach is the investigation of genomic or tran- ade of two proteins of the same oncogenic pathway for which
scriptomic signatures. The tumor–host–microenvironment cancer cells show a strong dependency (279). To maximize the
interaction could leave specific genomic/transcriptomic expected outcomes, selected treatment should target the pre-
scars, which could be identified by DNA-seq and RNA-seq. dominant cell populations within the tumor, prioritize clonal
Examples include the T cell–inflamed TME signature, the rather than subclonal target inhibition, and prioritize accord-
transcriptional signatures of fibroblasts, endothelial cells, ing to the highest level of actionability of the targets. In the
TGFβ signaling, and MDSCs and regulatory Tregs (156). The case of sequential therapies, monitoring clonal and subclonal
comparison of genomic/transcriptomic signatures between alterations may guide subsequent treatment based on the
responsive and nonresponsive tumors might identify pre- dominant resistant cell population (Fig. 5B).
dictors of resistance. For instance, the innate anti–PD-1
resistance signature (IPRES) was found to be correlated Treating Multiple, Highly Resistant Alterations
with primary resistance to PD-1 blockade in patients with The treatment of complex, polyclonal aberrations is a major
melanoma. IPRES displays enhanced expression of genes challenge. The sequential use of TKIs may result in a stepwise
involved in mesenchymal transition, extracellular matrix accumulation of mutations that highly increase drug resist-
remodeling, angiogenesis, cell adhesion, and wound healing. ance (56, 263, 280). Compound on-target alterations may
Notably, the same signature could occur in melanomas with induce resistance even to third-generation TKIs. Also, multi-
activated MAPK signaling following treatment with TKIs, ple bypass tracks or on-target and bypass alterations might
which could explain in some cases the cross-resistance with co-occur in the same patient (59). In these cases, it is currently
anti–PD-1 therapy (269). Another challenge is capturing the unknown if targets should be ranked for treatment prior-
heterogeneity of TME immune infiltration, which might be itization, or if drug combinations should be used. With the
addressed by single-cell analysis and radiomic signatures increasing number of acquired alterations, multidrug resist-
(e.g., the radiomic signature of CD8-infiltrating T cells; refs. ance would probably need different strategies. One example
261, 262, 270). is the administration of nonselective cytotoxic therapies with
Nongenomic mechanisms of resistance may also have an preferential delivery in the tumor, such as ADCs. The selective
important role in tumors failing targeted therapies, especially in intratumoral accumulation of ADCs is based on their ability
the absence of clear bona fide genomic alterations responsible to target proteins overexpressed in the tumor (281). The novel
for resistance. In addition to genomic analyses, the integration HER2-targeted ADC trastuzumab deruxtecan recently showed
of a proteomic and epigenetic characterization may expand our increased efficacy in heavily pretreated HER2-expressing or
knowledge and further improve treatment tailoring. HER2-mutant solid tumors (282). As tumor heterogeneity is
likely to be increased in this stage, the use of ADCs that harbor
Defer Resistance and Treat Earlier a bystander effect is likely to have more efficacy.
A growing tumor develops novel alterations and exhibits
increasing genomic instability, clonal diversity, and tumor Manipulating the Mutational Process Might
heterogeneity, which enhance drug resistance even in the Be More Important than Treating the Genomic
absence of prior treatments (Fig. 5B; refs. 12, 181, 271). It is Alteration That Mediates Resistance
possible that an earlier drug administration might be more If preventing resistance is more effective than treating
effective than a later administration. In favor of this hypoth- resistance, targeting the mutational process that drives can-
esis is the efficacy of tamoxifen in the adjuvant therapy of cer progression might have a paramount importance. Block-
localized breast cancer, and osimertinib in the first-line set- ing mutagenesis by “therapy by hypomutation” (e.g., A3B
ting of advanced EGFR-mutated NSCLC or even in the adju- inhibition of the APOBEC complex) would be expected to
vant setting, where it showed an impressive reduction in the prevent subclone diversification and thus limit the aggres-
risk of relapse (272–274). The lesser subclonal neoantigens siveness of tumors and drug resistance (94, 283). The thera-
and lower tumoral burden in earlier stages may also favor an peutic targeting of ecDNA would share a similar purpose and
earlier use of immunotherapy. is worth further investigation (53). In advanced cancers, such
Novel-generation drug use in the up-front setting may strategies should be evaluated in conjunction with the block-
prevent the emergence of on-target resistance mechanisms ade of tumor drivers, where it may increase the treatment
seen with older-generation drugs, as successfully exemplified duration on targeted treatments, while in early cancer stages
in the FLAURA trial with osimertinib versus first-genera- potentially as maintenance therapy after local treatments.
tion EGFRi in patients with treatment-naïve EGFR-mutated A reverse strategy is the use of “therapy by hypermutation,”
advanced NSCLC (273). However, in this trial and others, by enhancing mutagenesis until cell destabilization and cell
deferring resistance alterations is not the only factor respon- death. This could be useful in advanced, pretreated APOBEC-
sible for improved clinical outcomes, as the increased intrac- high tumors that already achieved hard-to-treat genomic
ranial activity of newer-generation drugs has an essential role alterations. Strategies may use agonists of A3, PARP, and ATR

April 2021 CANCER DISCOVERY 10TH ANNIVERSARY ISSUE | 889

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
REVIEW Aldea et al.

inhibitors (94, 283). A similar strategy may be used in tumors resistance mechanism. Also, as opposed to targeted thera-
with high chromosomal instability, by using agents that pies, advancements in immunotherapeutic drug discovery
further increase genomic instability (e.g., WEE1 inhibitors in and research are slowed down by the intrinsic limitations of
conjunction with DNA-damaging agents). current preclinical models for testing, which need an intact
human TME in order to be useful (291).
Demonstrate the Missing Links between Finding better tumor-agnostic biomarkers to predict
Oncogenic Stresses and Immune Suppression immunotherapy response is essential to find patients
Active research is needed to deepen the current under- responding to therapy in tumor types not expected to ben-
standing of immune exclusion and suppression associated efit. Tertiary lymphoid structures, recently shown to highly
with certain oncogenic pathways, as to be able to reverse this correlate with immunotherapy responses even in tumors with
phenomenon and to integrate immunotherapy in the thera- low mutational burden, warrant further investigation in a
peutic armamentarium of oncogene-addicted tumors. In this pan-tumor setting (221). Resistance biomarkers need further
domain, research is still in its infancy. Examples of putative validation and ranking, based on their magnitude of predic-
players are the expression of CD73 in EGFR-mutated cancers tion. For example, resistance mechanisms found in nonre-
that might predict ICB responses (284), the PTEN-loss PI3K sponding tumors that were highly expected to respond, such
pathway activation that induces PD-L1 expression and pro- as MSI-high tumors, should have a high prediction value of
motes resistance to T-cell lysis (285), or the INFα/β-induced resistance in other tumor types as well.
p53 gene activation with subsequent cell apoptosis (286). Patient “immunograms,” first proposed by Blank and col-
Associations between ICB and epigenetic drugs should be leagues (292), should be further developed to integrate the
used in a personalized fashion, guided by epigenetic modi- current state of knowledge (Fig. 6A) and guide treatment
fications. Epigenetic alterations would need clinical studies choice (Fig. 6B). Trials investigating a personalized approach
to interrogate and validate a potential predictive role for ICB with combination therapies based on biomarkers are ongo-
response, as certain mutations have been correlated with ICB ing (e.g., ADVISE, NCT03335540; PIONeeR, NCT03833440).
response, especially the loss of SWI/SNF components (287, The preliminary results of the PIONeeR trial suggest a predic-
288). Also, genomic ITH might determine different immune tive value for PD-L1 density on all cell types, PD-L1 tumor
infiltration patterns. For instance, tumor subregions with expression, infiltrating cytotoxic T cells, and immunosup-
copy-number gain in chromosome 7 were associated with pressive cell density. Of note, nonresponding patients with
unfavorable immune composition, with a lack of leukocyte high PD-L1 tumors actually had low PD-L1 density on all cell
infiltration and presence of activated neutrophils (289). The types (tumor and stromal; ref. 293).
identification of the molecular-associated immune response
in the tumor might be crucial for anticipating resistance pat- Better Define Acquired Resistance
terns and personalizing treatment. to Immunotherapy
An accepted set of criteria to define acquired resistance to
Personalizing Treatment Selection immunotherapy is essential for future drug development.
with Immunotherapy Many new immunotherapy compounds have failed because
There is an urgent need to personalize the administration they enroll in clinical trials both patients with primary resist-
of immunotherapy, as the blinded use of ICB and investiga- ance and patients with acquired resistance to ICB. The lack
tional immunotherapies is most likely to fail in a significant of consensus for defining resistance means that patients
proportion of cases where the target is absent. The most illus- with acquired resistance are defined by some as only patients
trative example is the administration of PD-1/PD-L1 inhibi- with partial response/complete response versus others who
tors in tumors with no T-cell infiltration. Another example also account for stable disease (SD) for a certain period. It
is the combination of anti–PD-1 with anti-IDO that showed is totally unclear if SD for 3 or 6 months truly represents a
promising antitumor activity in phase I–II trials on bio- sensitive patient who then acquires resistance. In order to
marker-unselected patients, but the phase III trial showed no increase the likelihood of response to new immunotherapy
benefit of the combination in comparison with single-agent molecules, we should probably focus more on patients who
PD-1 inhibitor (290). Would the results be the same in selected truly responded to ICB initially and then developed acquired
patients with high IDO–expressing tumors? Also, strategies resistance. Such patients may still have inflamed tumors that
for enhancing the function of effector T cells are likely to can be amenable to a new immunomodulation.
work in “hot” tumors with T-cell infiltration, but are they
effective in tumors featuring immune exclusion or minimal Aim to Cure: The Role of Surgery or Local Ablative
immune cell infiltrates? MHC I–deficient cells are unable to Therapies in the Treatment of Metastatic Disease
present antigen to effector T cells, but they are susceptible to In oligoprogressive tumors, local therapies are used, as
NK cell–dependent cytotoxicity, which provides a rationale they allow the pursuit of an effective treatment for the rest of
for NK-cell stimulation in such cases (255). Such tumors are the lesions and improve survival outcomes (294–296). They
also good candidates for CAR-T cell therapy, as CAR-T cells are also discussed in tumor boards in the case of dissociated
are able to recognize antigens without MHC, whereas TCR response under ICB, but prospective studies are still pending
therapies could better address tumors with lower antigen to confirm benefit (297). The choice between surgery or local
densities within the TME (193). Similarly, investigational ablative therapies still needs refinement, especially in selected
agents aiming to overcome acquired resistance to immuno- oligoprogressive patients with a long disease control under
therapy are unlikely to succeed without targeting the specific treatment. Surgery remains the most effective treatment for

890 | CANCER DISCOVERY 10TH ANNIVERSARY ISSUE April 2021 AACRJournals.org

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
Overcoming Resistance to Targeted Therapy and Immunotherapy REVIEW

A B Predicted response to
anti–PD-1/PD-L1 therapy
1
Tumor foreignness 14 2
Tumor mutational burden
13 3

Clonal neoantigen 12 4
burden
Low neoantigen
Favorable genomics intratumoral heterogeneity 5
1 11
14 2
Favorable microbiota 6
General immune status
Favorable TME 13 3
Lymphocyte count
No hypoxia, no spacial 10 7
distance between tumor
12 4 Low immune 8
cells and TILs 9
heterogeneity

Tumor sensibility to 5 TIL Secondary resistance to anti–PD-1/PD-L1


immune effectors 11 therapy
MHC 6
IFN signaling Tertiary lymphoid
structures 1
14 2 1
14 2
Absence of inhibitory 10 7
13 3
tumor metabolism Absence of coinhibitory 13 3
LDH, glucose checkpoints
8 LAG3, VISTA, TIGIT, etc. 12 4 12
utilization, IDO 9 4

Absence of Absence of 5 Loss of 5


soluble inhibitors PD-L1 11 MHC 11
6 6
IL6, CRP
Loss of
immune
10 7
10 7 infiltration
Expression
8
9 of LAG3

Anti– Anti-LAG3 antibody


NK-cell
PD-1/PD-L1
stimulators
antibody

Figure 6. Cancer immunogram adapted with permission from Blank et al. (292). A, Ideal features of an “immunogenic” tumor. B, Immunograms cor-
responding to a theoretical patient illustrating features predictive of response to anti–PD-1/PD-L1 antibodies, followed by two examples of secondary
resistance with potential tailored treatments. Red text is used to illustrate the changes made to the initial cancer immunogram of Blank et al.

ITH, hypoxic tumors, and persister cells, and these might be non-financial support, and other from Pfizer, Merck Serono, and
the exact triggers behind progression. However, the selection Novartis outside the submitted work. F. Barlesi reports personal
of the ideal candidates is challenged by the difficulty of predict- fees from AstraZeneca, Bayer, Bristol-Myers Squibb, Boehringer-
ing the behavior of the rest of the lesions. Innovative methods Ingelheim, Eli Lilly Oncology, F. Hoffmann-La Roche Ltd, Novartis,
Merck, Mirati, MSD, Pierre Fabre, Pfizer, Seattle Genetics, and
predicting relapse, based on genomics and radiomics, would
Takeda outside the submitted work. J.-C. Soria reports personal fees
be desired. Also, radiotherapy could be modulated based on from Relay Therapeutics, and other from AstraZeneca, Daiichi San-
the tumor genomic and microenvironmental features, as well kyo, Gritstone, and Hookipa Pharmaceuticals during the conduct of
as its immune infiltration. Another direction of research is to the study. No other disclosures were reported.
investigate the role of a “consolidation” therapy for patients
responding to treatment in a heterogeneous manner. The less Acknowledgments
responsive lesions might exhibit genomic/immune heteroge- The figures are previously unpublished original works, created
neity and be the culprit of subsequent local or distant progres- by Mihaela Aldea and Semih Dogan with BioRender.com, for the
sion. In the case of ICB, the addition of a local treatment of express purpose of publication in this Cancer Discovery review.
liver metastases might also be beneficial, especially when con-
sidering recent hypotheses of systemic immunosuppression Received November 9, 2020; revised January 13, 2021; accepted
induced by liver metastases, which might impair ICB efficacy. February 1, 2021; published first April 2, 2021.

Conclusion References
Having all this in mind, it is clear that despite major pro- 1. Hata AN, Niederst MJ, Archibald HL, Gomez-Caraballo M,
gress in cancer care, drug resistance evolves dynamically and Siddiqui FM, Mulvey HE, et al. Tumor cells can follow distinct
remains the most important barrier to achieving cure. A deeper evolutionary paths to become resistant to epidermal growth factor
understanding of the processes underlying resistance and more receptor inhibition. Nat Med 2016;22:262–9.
2. Schoenfeld AJ, Hellmann MD. Acquired resistance to immune
potent drugs and innovative strategies are urgently needed. checkpoint inhibitors. Cancer Cell 2020;37:443–55.
3. Sharma P, Hu-Lieskovan S, Wargo JA, Ribas A. Primary, adaptive,
Authors’ Disclosures and acquired resistance to cancer immunotherapy. Cell 2017;168:
F. Andre reports grants from Novartis, Roche, Daiichi, Eli Lilly, 707–23.
Astra Zeneca, and Pfizer outside the submitted work. A. Marabelle 4. Lovly CM, Shaw AT. Molecular pathways: resistance to kinase inhib-
reports grants, personal fees, non-financial support, and other from itors and implications for therapeutic strategies. Clin Cancer Res
AstraZeneca, BMS, MSD, and Roche/Genentech and personal fees, 2014;20:2249–56.

April 2021 CANCER DISCOVERY 10TH ANNIVERSARY ISSUE | 891

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
REVIEW Aldea et al.

5. Kalbasi A, Ribas A. Tumour-intrinsic resistance to immune check- sensitivity to anticancer drugs targeting the EGFR and HER2 recep-
point blockade. Nat Rev Immunol 2020;20:25–39. tors. Cancer Res 2011;71:4068–73.
6. Neel DS, Bivona TG. Resistance is futile: overcoming resistance 25. Boussemart L, Malka-Mahieu H, Girault I, Allard D, Hemmingsson
to targeted therapies in lung adenocarcinoma. NPJ Precis Oncol O, Tomasic G, et al. eIF4F is a nexus of resistance to anti-BRAF and
2017;1:3. anti-MEK cancer therapies. Nature 2014;513:105–9.
7. Jackman D, Pao W, Riely GJ, Engelman JA, Kris MG, Janne PA, et al. 26. Xue JY, Zhao Y, Aronowitz J, Mai TT, Vides A, Qeriqi B, et al. Rapid
Clinical definition of acquired resistance to epidermal growth factor non-uniform adaptation to conformation-specific KRAS(G12C)
receptor tyrosine kinase inhibitors in non-small-cell lung cancer. inhibition. Nature 2020;577:421–5.
J Clin Oncol 2010;28:357–60. 27. Pros E, Saigi M, Alameda D, Gomez-Mariano G, Martinez-Delgado B,
8. Kluger HM, Tawbi HA, Ascierto ML, Bowden M, Callahan MK, Cha E, Alburquerque-Bejar JJ, et al. Genome-wide profiling of non-smoking-
et al. Defining tumor resistance to PD-1 pathway blockade: recom- related lung cancer cells reveals common RB1 rearrangements
mendations from the first meeting of the SITC immunotherapy associated with histopathologic transformation in EGFR-mutant
resistance taskforce. J Immunother Cancer 2020;8:e000398. tumors. Ann Oncol 2020;31:274–82.
9. Lievre A, Bachet JB, Le Corre D, Boige V, Landi B, Emile JF, et al. 28. Schoenfeld AJ, Chan JM, Kubota D, Sato H, Rizvi H, Daneshbod
KRAS mutation status is predictive of response to cetuximab ther- Y, et al. Tumor analyses reveal squamous transformation and off-
apy in colorectal cancer. Cancer Res 2006;66:3992–5. target alterations as early resistance mechanisms to first-line osimer-
10. Karapetis CS, Khambata-Ford S, Jonker DJ, O’Callaghan CJ, Tu D, tinib in EGFR-mutant lung cancer. Clin Cancer Res 2020;26:2654–63.
Tebbutt NC, et al. K-ras mutations and benefit from cetuximab in 29. Rottenberg S, Jaspers JE, Kersbergen A, van der Burg E, Nygren AO,
advanced colorectal cancer. N Engl J Med 2008;359:1757–65. Zander SA, et al. High sensitivity of BRCA1-deficient mammary
11. Nakadate Y, Kodera Y, Kitamura Y, Shirasawa S, Tachibana T, tumors to the PARP inhibitor AZD2281 alone and in combination
Tamura T, et al. KRAS mutation confers resistance to antibody- with platinum drugs. Proc Natl Acad Sci U S A 2008;105:17079–84.
dependent cellular cytotoxicity of cetuximab against human colo- 30. Ahnert JR, Gray N, Mok T, Gainor J. What it takes to improve a
rectal cancer cells. Int J Cancer 2014;134:2146–55. first-generation inhibitor to a second- or third-generation small
12. Bertucci F, Ng CKY, Patsouris A, Droin N, Piscuoglio S, Carbuccia N, molecule. Am Soc Clin Oncol Educ Book 2019:196–205.
et al. Genomic characterization of metastatic breast cancers. Nature 31. Ramalingam SS, Cheng Y, Zhou C, Ohe Y, Imamura F, Cho BC,
2019;569:560–4. et al. Mechanisms of acquired resistance to first-line osimertinib:
13. Yang JJ, Zhang XC, Su J, Xu CR, Zhou Q, Tian HX, et al. Lung can- preliminary data from the phase III FLAURA study. Ann Oncol 2018;
cers with concomitant EGFR mutations and ALK rearrangements: 29:viii740.
diverse responses to EGFR-TKI and crizotinib in relation to diverse 32. Mok TS, Wu YL, Ahn MJ, Garassino MC, Kim HR, Ramalingam SS,
receptors phosphorylation. Clin Cancer Res 2014;20:1383–92. et al. Osimertinib or platinum-pemetrexed in EGFR T790M-positive
14. Schmid S, Gautschi O, Rothschild S, Mark M, Froesch P, Klingbiel D, lung cancer. N Engl J Med 2017;376:629–40.
et al. Clinical outcome of ALK-positive non-small cell lung cancer 33. Recondo G, Facchinetti F, Olaussen KA, Besse B, Friboulet L. Mak-
(NSCLC) patients with de novo EGFR or KRAS co-mutations ing the first move in EGFR-driven or ALK-driven NSCLC: first-
receiving tyrosine kinase inhibitors (TKIs). J Thorac Oncol 2017;12: generation or next-generation TKI? Nat Rev Clin Oncol 2018;15:
681–8. 694–708.
15. Canale M, Petracci E, Delmonte A, Chiadini E, Dazzi C, Papi M, et al. 34. Shaw AT, Solomon BJ, Besse B, Bauer TM, Lin CC, Soo RA, et al.
Impact of TP53 mutations on outcome in EGFR-mutated patients ALK resistance mutations and efficacy of lorlatinib in advanced
treated with first-line tyrosine kinase inhibitors. Clin Cancer Res anaplastic lymphoma kinase-positive non-small-cell lung cancer.
2017;23:2195–202. J Clin Oncol 2019;37:1370–9.
16. Adashek JJ, Kato S, Parulkar R, Szeto CW, Sanborn JZ, Vaske CJ, et al. 35. Katayama R, Gong B, Togashi N, Miyamoto M, Kiga M, Iwasaki S,
Transcriptomic silencing as a potential mechanism of treatment et al. The new-generation selective ROS1/NTRK inhibitor DS-
resistance. JCI insight 2020;5:e134824. 6051b overcomes crizotinib resistant ROS1-G2032R mutation in
17. Agrawal N, Dasaradhi PV, Mohmmed A, Malhotra P, Bhatnagar RK, preclinical models. Nat Commun 2019;10:3604.
Mukherjee SK. RNA interference: biology, mechanism, and applica- 36. Lin JJ, Shaw AT. Recent advances in targeting ROS1 in lung cancer.
tions. Microbiol Mol Biol Rev 2003;67:657–85. J Thorac Oncol 2017;12:1611–25.
18. Bosch A, Li Z, Bergamaschi A, Ellis H, Toska E, Prat A, et al. 37. Yun MR, Kim DH, Kim SY, Joo HS, Lee YW, Choi HM, et al.
PI3K inhibition results in enhanced estrogen receptor function and Repotrectinib exhibits potent antitumor activity in treatment-naive
dependence in hormone receptor-positive breast cancer. Sci Transl and solvent-front-mutant ROS1-rearranged non-small cell lung
Med 2015;7:283ra51. cancer. Clin Cancer Res 2020;26:3287–95.
19. Hopkins BD, Pauli C, Du X, Wang DG, Li X, Wu D, et al. Suppres- 38. Solomon BJ, Tan L, Lin JJ, Wong SQ, Hollizeck S, Ebata K, et al.
sion of insulin feedback enhances the efficacy of PI3K inhibitors. RET solvent front mutations mediate acquired resistance to selec-
Nature 2018;560:499–503. tive RET inhibition in RET-driven malignancies. J Thorac Oncol
20. Ducreux M, Chamseddine A, Laurent-Puig P, Smolenschi C, 2020;15:541–9.
Hollebecque A, Dartigues P, et al. Molecular targeted therapy of 39. Drilon A, Ou SI, Cho BC, Kim DW, Lee J, Lin JJ, et al. Repotrectinib
BRAF-mutant colorectal cancer. Ther Adv Med Oncol 2019;11: (TPX-0005) is a next-generation ROS1/TRK/ALK inhibitor that
1758835919856494. potently inhibits ROS1/TRK/ALK solvent- front mutations. Cancer
21. Amodio V, Yaeger R, Arcella P, Cancelliere C, Lamba S, Lorenzato A, Discov 2018;8:1227–36.
et al. EGFR blockade reverts resistance to KRAS(G12C) inhibition 40. Drilon A, Laetsch TW, Kummar S, DuBois SG, Lassen UN, Demetri
in colorectal cancer. Cancer Discov 2020;10:1129–39. GD, et al. Efficacy of larotrectinib in TRK fusion-positive cancers in
22. Rozengurt E, Soares HP, Sinnet-Smith J. Suppression of feedback adults and children. N Engl J Med 2018;378:731–9.
loops mediated by PI3K/mTOR induces multiple overactivation 41. Mian AA, Schull M, Zhao Z, Oancea C, Hundertmark A, Beissert T,
of compensatory pathways: an unintended consequence leading to et al. The gatekeeper mutation T315I confers resistance against
drug resistance. Mol Cancer Ther 2014;13:2477–88. small molecules by increasing or restoring the ABL-kinase activ-
23. Garrett JT, Olivares MG, Rinehart C, Granja-Ingram ND, Sanchez V, ity accompanied by aberrant transphosphorylation of endogenous
Chakrabarty A, et al. Transcriptional and posttranslational up-reg- BCR, even in loss-of-function mutants of BCR/ABL. Leukemia
ulation of HER3 (ErbB3) compensates for inhibition of the HER2 2009;23:1614–21.
tyrosine kinase. Proc Natl Acad Sci U S A 2011;108:5021–6. 42. Wang J, Yao Z, Jonsson P, Allen AN, Qin ACR, Uddin S, et al. A sec-
24. Zindy P, Berge Y, Allal B, Filleron T, Pierredon S, Cammas A, et al. ondary mutation in BRAF confers resistance to RAF inhibition in a
Formation of the eIF4F translation-initiation complex determines BRAF(V600E)-mutant brain tumor. Cancer Discov 2018;8:1130–41.

892 | CANCER DISCOVERY 10TH ANNIVERSARY ISSUE April 2021 AACRJournals.org

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
Overcoming Resistance to Targeted Therapy and Immunotherapy REVIEW

43. Lin JJ, Zhu VW, Yoda S, Yeap BY, Schrock AB, Dagogo-Jack I, 63. Suzawa K, Offin M, Lu D, Kurzatkowski C, Vojnic M, Smith RS,
et al. Impact of EML4-ALK variant on resistance mechanisms and et al. Activation of KRAS mediates resistance to targeted therapy in
clinical outcomes in ALK-positive lung cancer. J Clin Oncol 2018;36: MET exon 14-mutant non-small cell lung cancer. Clin Cancer Res
1199–206. 2019;25:1248–60.
44. Van Emburgh BO, Arena S, Siravegna G, Lazzari L, Crisafulli G, 64. Van Allen EM, Wagle N, Sucker A, Treacy DJ, Johannessen CM,
Corti G, et al. Acquired RAS or EGFR mutations and duration of Goetz EM, et al. The genetic landscape of clinical resistance to RAF
response to EGFR blockade in colorectal cancer. Nat Commun inhibition in metastatic melanoma. Cancer Discov 2014;4:94–109.
2016;7:13665. 65. Trunzer K, Pavlick AC, Schuchter L, Gonzalez R, McArthur GA,
45. Arena S, Bellosillo B, Siravegna G, Martinez A, Canadas I, ­Lazzari L, Hutson TE, et al. Pharmacodynamic effects and mechanisms of
et al. Emergence of multiple EGFR extracellular mutations during resistance to vemurafenib in patients with metastatic melanoma.
cetuximab treatment in colorectal cancer. Clin Cancer Res 2015;21: J Clin Oncol 2013;31:1767–74.
2157–66. 66. Wagle N, Emery C, Berger MF, Davis MJ, Sawyer A, Pochanard P,
46. Jeselsohn R, Bergholz JS, Pun M, Cornwell M, Liu W, Nardone A, et al. Dissecting therapeutic resistance to RAF inhibition in mela-
et al. Allele-specific chromatin recruitment and therapeutic vulner- noma by tumor genomic profiling. J Clin Oncol 2011;29:3085–96.
abilities of ESR1 activating mutations. Cancer Cell 2018;33:173–86. 67. Pearson A, Proszek P, Pascual J, Fribbens C, Shamsher MK, Kingston B,
47. Antonarakis ES, Lu C, Wang H, Luber B, Nakazawa M, Roeser JC, et al. Inactivating NF1 mutations are enriched in advanced breast
et al. AR-V7 and resistance to enzalutamide and abiraterone in pros- cancer and contribute to endocrine therapy resistance. Clin Cancer
tate cancer. N Engl J Med 2014;371:1028–38. Res 2020;26:608–22.
48. Vido MJ, Le K, Hartsough EJ, Aplin AE. BRAF splice variant resist- 68. Sos ML, Koker M, Weir BA, Heynck S, Rabinovsky R, Zander T, et al.
ance to RAF inhibitor requires enhanced MEK association. Cell Rep PTEN loss contributes to erlotinib resistance in EGFR-mutant lung
2018;25:1501–10. cancer by activation of Akt and EGFR. Cancer Res 2009;69:3256–61.
49. Shi H, Moriceau G, Kong X, Lee MK, Lee H, Koya RC, et al. Mela- 69. Xu JM, Wang Y, Wang YL, Wang Y, Liu T, Ni M, et al. PIK3CA muta-
noma whole-exome sequencing identifies (V600E)B-RAF amplifica- tions contribute to acquired cetuximab resistance in patients with
tion-mediated acquired B-RAF inhibitor resistance. Nat Commun metastatic colorectal cancer. Clin Cancer Res 2017;23:4602–16.
2012;3:724. 70. Shi H, Hugo W, Kong X, Hong A, Koya RC, Moriceau G, et al.
50. Edwards J, Krishna NS, Grigor KM, Bartlett JM. Androgen receptor Acquired resistance and clonal evolution in melanoma during BRAF
gene amplification and protein expression in hormone refractory inhibitor therapy. Cancer Discov 2014;4:80–93.
prostate cancer. Br J Cancer 2003;89:552–6. 71. Zhu VW, Klempner SJ, Ou SI. Receptor tyrosine kinase fusions as an
51. Sequist LV, Waltman BA, Dias-Santagata D, Digumarthy S, Turke AB, actionable resistance mechanism to EGFR TKIs in EGFR-mutant
Fidias P, et al. Genotypic and histological evolution of lung can- non-small-cell lung cancer. Trends Cancer 2019;5:677–92.
cers acquiring resistance to EGFR inhibitors. Sci Transl Med 72. Enrico D, Lacroix L, Chen J, Rouleau E, Scoazec J-Y, Loriot Y, et al.
2011;3:75ra26. Oncogenic fusions may be frequently present at resistance of EGFR
52. Katayama R, Shaw AT, Khan TM, Mino-Kenudson M, Solomon BJ, tyrosine kinase inhibitors in patients with NSCLC: a brief report.
Halmos B, et al. Mechanisms of acquired crizotinib resistance in JTO Clin Res Reports 2020;1:100023.
ALK-rearranged lung cancers. Sci Transl Med 2012;4:120ra17. 73. Xu C, Li D, Duan W, Tao M. TPD52L1-ROS1 rearrangement as a
53. Bailey C, Shoura MJ, Mischel PS, Swanton C. Extrachromosomal new acquired resistance mechanism to osimertinib that responds
DNA-relieving heredity constraints, accelerating tumour evolution. to crizotinib in combination with osimertinib in lung adenocarci-
Ann Oncol 2020;31:884–93. noma. JTO Clin Res Reports 2020:100034.
54. Nathanson DA, Gini B, Mottahedeh J, Visnyei K, Koga T, Gomez G, 74. Rich TA, Reckamp KL, Chae YK, Doebele RC, Iams WT, Oh M,
et al. Targeted therapy resistance mediated by dynamic regulation of et al. Analysis of cell-free DNA from 32,989 advanced cancers reveals
extrachromosomal mutant EGFR DNA. Science 2014;343:72–6. novel co-occurring activating RET alterations and oncogenic signal-
55. Juric D, Castel P, Griffith M, Griffith OL, Won HH, Ellis H, et al. ing pathway aberrations. Clin Cancer Res 2019;25:5832–42.
Convergent loss of PTEN leads to clinical resistance to a PI(3)Kal- 75. Clifton K, Rich TA, Parseghian C, Raymond VM, Dasari A, Pereira
pha inhibitor. Nature 2015;518:240–4. AAL, et al. Identification of actionable fusions as an anti-EGFR
56. Redaelli S, Ceccon M, Zappa M, Sharma GG, Mastini C, Mauri M, resistance mechanism using a circulating tumor DNA assay. JCO
et al. Lorlatinib treatment elicits multiple on- and off-target mecha- Precis Oncol 2019;3:PO.19.00141.
nisms of resistance in ALK-driven cancer. Cancer Res 2018;78: 76. Daly C, Castanaro C, Zhang W, Zhang Q, Wei Y, Ni M, et al. FGFR3-
6866–80. TACC3 fusion proteins act as naturally occurring drivers of tumor
57. Liu Q, Yu S, Zhao W, Qin S, Chu Q, Wu K. EGFR-TKIs resistance via resistance by functionally substituting for EGFR/ERK signaling.
EGFR-independent signaling pathways. Mol Cancer 2018;17:53. Oncogene 2017;36:471–81.
58. Murtuza A, Bulbul A, Shen JP, Keshavarzian P, Woodward BD, 77. Ouyang X, Barling A, Lesch A, Tyner JW, Choonoo G, Zheng C, et al.
Lopez-Diaz FJ, et al. Novel third-generation EGFR tyrosine kinase Induction of anaplastic lymphoma kinase (ALK) as a novel mecha-
inhibitors and strategies to overcome therapeutic resistance in lung nism of EGFR inhibitor resistance in head and neck squamous
cancer. Cancer Res 2019;79:689–98. cell carcinoma patient-derived models. Cancer Biol Ther 2018;19:
59. Recondo G, Bahcall M, Spurr LF, Che J, Ricciuti B, Leonardi GC, 921–33.
et al. Molecular mechanisms of acquired resistance to MET tyrosine 78. Ross DS, Liu B, Schram AM, Razavi P, Lagana SM, Zhang Y, et al.
kinase inhibitors in patients with MET exon 14-mutant NSCLC. Enrichment of kinase fusions in ESR1 wild type, metastatic breast
Clin Cancer Res 2020;26:2615–25. cancer revealed by a systematic analysis of 4,854 patients. Ann
60. Vojnic M, Kubota D, Kurzatkowski C, Offin M, Suzawa K, Benayed Oncol 2020;31:991–1000.
R, et al. Acquired BRAF rearrangements induce secondary resistance 79. Sharma SV, Lee DY, Li B, Quinlan MP, Takahashi F, Maheswaran S,
to EGFR therapy in EGFR-mutated lung cancers. J Thorac Oncol et al. A chromatin-mediated reversible drug-tolerant state in cancer
2019;14:802–15. cell subpopulations. Cell 2010;141:69–80.
61. Cocco E, Schram AM, Kulick A, Misale S, Won HH, Yaeger R, et al. 80. Fujiwara K, Kiura K, Ueoka H, Tabata M, Hamasaki S, Tanimoto M.
Resistance to TRK inhibition mediated by convergent MAPK path- Dramatic effect of ZD1839 (‘Iressa’) in a patient with advanced non-
way activation. Nat Med 2019;25:1422–7. small-cell lung cancer and poor performance status. Lung Cancer
62. Nelson-Taylor SK, Le AT, Yoo M, Schubert L, Mishall KM, Doak A, 2003;40:73–6.
et al. Resistance to RET-inhibition in RET-rearranged NSCLC is 81. Leonetti A, Sharma S, Minari R, Perego P, Giovannetti E, Tiseo M.
mediated by reactivation of RAS/MAPK signaling. Mol Cancer Ther Resistance mechanisms to osimertinib in EGFR-mutated non-small
2017;16:1623–33. cell lung cancer. Br J Cancer 2019;121:725–37.

April 2021 CANCER DISCOVERY 10TH ANNIVERSARY ISSUE | 893

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
REVIEW Aldea et al.

82. Xu G, Chhangawala S, Cocco E, Razavi P, Cai Y, Otto JE, et al. ARID1A effective against vemurafenib-resistant melanoma harboring the
determines luminal identity and therapeutic response in estrogen- MEK1-C121S mutation in a preclinical model. Mol Cancer Ther
receptor-positive breast cancer. Nat Genet 2020;52:198–207. 2014;13:823–32.
83. Boumahdi S, de Sauvage FJ. The great escape: tumour cell plastic- 104. Chiba M, Togashi Y, Bannno E, Kobayashi Y, Nakamura Y, Hayashi H,
ity in resistance to targeted therapy. Nat Rev Drug Discov 2020;19: et al. Efficacy of irreversible EGFR-TKIs for the uncommon second-
39–56. ary resistant EGFR mutations L747S, D761Y, and T854A. BMC
84. Quintanal-Villalonga A, Chan JM, Yu HA, Pe’er D, Sawyers CL, Sen Cancer 2017;17:281.
T, et al. Lineage plasticity in cancer: a shared pathway of therapeutic 105. Okada K, Araki M, Sakashita T, Ma B, Kanada R, Yanagitani N,
resistance. Nat Rev Clin Oncol 2020;17:360–71. et al. Prediction of ALK mutations mediating ALK-TKIs resistance
85. Noordermeer SM, van Attikum H. PARP inhibitor resistance: a tug- and drug re-purposing to overcome the resistance. EBioMedicine
of-war in BRCA-mutated cells. Trends Cell Biol 2019;29:820–34. 2019;41:105–19.
86. Mateo J, Lord CJ, Serra V, Tutt A, Balmana J, Castroviejo-Bermejo M, 106. Facchinetti F, Loriot Y, Kuo MS, Mahjoubi L, Lacroix L, Planchard D,
et al. A decade of clinical development of PARP inhibitors in per- et al. Crizotinib-resistant ROS1 mutations reveal a predictive kinase
spective. Ann Oncol 2019;30:1437–47. inhibitor sensitivity model for ROS1- and ALK-rearranged lung
87. Lin KK, Harrell MI, Oza AM, Oaknin A, Ray-Coquard I, Tinker AV, cancers. Clin Cancer Res 2016;22:5983–91.
et al. BRCA reversion mutations in circulating tumor DNA predict 107. Drilon A, Somwar R, Wagner JP, Vellore NA, Eide CA, Zabriskie MS,
primary and acquired resistance to the PARP inhibitor rucaparib in et al. A novel crizotinib-resistant solvent-front mutation responsive
high-grade ovarian carcinoma. Cancer Discov 2019;9:210–9. to cabozantinib therapy in a patient with ROS1-rearranged lung
88. Kuukasjarvi T, Kononen J, Helin H, Holli K, Isola J. Loss of estrogen cancer. Clin Cancer Res 2016;22:2351–8.
receptor in recurrent breast cancer is associated with poor response 108. Drilon A, Nagasubramanian R, Blake JF, Ku N, Tuch BB, Ebata K,
to endocrine therapy. J Clin Oncol 1996;14:2584–9. et al. A next-generation TRK kinase inhibitor overcomes acquired
89. Katayama R, Sakashita T, Yanagitani N, Ninomiya H, Horiike A, resistance to prior TRK kinase inhibition in patients with TRK
Friboulet L, et al. P-glycoprotein mediates ceritinib resistance in fusion-positive solid tumors. Cancer Discov 2017;7:963–72.
anaplastic lymphoma kinase-rearranged non-small cell lung cancer. 109. Smith BD, Kaufman MD, Lu WP, Gupta A, Leary CB, Wise SC,
EBioMedicine 2016;3:54–66. et al. Ripretinib (DCC-2618) is a switch control kinase inhibitor of a
90. Russo M, Crisafulli G, Sogari A, Reilly NM, Arena S, Lamba S, et al. broad spectrum of oncogenic and drug-resistant KIT and PDGFRA
Adaptive mutability of colorectal cancers in response to targeted variants. Cancer Cell 2019;35:738–51.
therapies. Science 2019;366:1473–80. 110. Goyal L, Shi L, Liu LY, Fece de la Cruz F, Lennerz JK, Raghavan S,
91. Straussman R, Morikawa T, Shee K, Barzily-Rokni M, Qian ZR, Du J, et al. TAS-120 overcomes resistance to ATP-competitive FGFR
et al. Tumour micro-environment elicits innate resistance to RAF inhibitors in patients with FGFR2 fusion-positive intrahepatic chol-
inhibitors through HGF secretion. Nature 2012;487:500–4. angiocarcinoma. Cancer Discov 2019;9:1064–79.
92. Chen H, Lin C, Peng T, Hu C, Lu C, Li L, et al. Metformin reduces 111. Hatlen MA, Schmidt-Kittler O, Sherwin CA, Rozsahegyi E, Rubin N,
HGF-induced resistance to alectinib via the inhibition of Gab1. Cell Sheets MP, et al. Acquired on-target clinical resistance validates
Death Dis 2020;11:111. FGFR4 as a driver of hepatocellular carcinoma. Cancer Discov
93. Swanton C, McGranahan N, Starrett GJ, Harris RS. APOBEC 2019;9:1686–95.
enzymes: mutagenic fuel for cancer evolution and heterogeneity. 112. Katayama R, Kobayashi Y, Friboulet L, Lockerman EL, Koike S,
Cancer Discov 2015;5:704–12. Shaw AT, et al. Cabozantinib overcomes crizotinib resistance in
94. Venkatesan S, Rosenthal R, Kanu N, McGranahan N, Bartek J, ROS1 fusion-positive cancer. Clin Cancer Res 2015;21:166–74.
Quezada SA, et al. Perspective: APOBEC mutagenesis in drug resist- 113. Haura EB, Cho BC, Lee JS, Han J-Y, Lee KH, Sanborn RE, et al. JNJ-
ance and immune escape in HIV and cancer evolution. Ann Oncol 61186372 (JNJ-372), an EGFR-cMet bispecific antibody, in EGFR-
2018;29:563–72. driven advanced non-small cell lung cancer (NSCLC). J Clin Oncol
95. Jamal-Hanjani M, Wilson GA, McGranahan N, Birkbak NJ, Watkins 2019;37:9009.
TBK, Veeriah S, et al. Tracking the evolution of non-small-cell lung 114. Cross DA, Ashton SE, Ghiorghiu S, Eberlein C, Nebhan CA, Spitzler PJ,
cancer. N Engl J Med 2017;376:2109–21. et al. AZD9291, an irreversible EGFR TKI, overcomes T790M-medi-
96. Alexandrov LB, Nik-Zainal S, Wedge DC, Aparicio SA, Behjati S, ated resistance to EGFR inhibitors in lung cancer. Cancer Discov
Biankin AV, et al. Signatures of mutational processes in human 2014;4:1046–61.
cancer. Nature 2013;500:415–21. 115. Cocco E, Lee JE, Kannan S, Schram AM, Won HH, Shifman S, et al.
97. Burns MB, Temiz NA, Harris RS. Evidence for APOBEC3B mutagen- TRK xDFG mutations trigger a sensitivity switch from type I to II
esis in multiple human cancers. Nat Genet 2013;45:977–83. kinase inhibitors. Cancer Discov 2021;11:126–41.
98. Riely GJ, Neal JW, Camidge DR, Spira A, Piotrowska Z, Horn L, et al. 116. Cromm PM, Crews CM. Targeted protein degradation: from chemi-
1261MO Updated results from a phase I/II study of mobocertinib cal biology to drug discovery. Cell Chem Biol 2017;24:1181–90.
(TAK-788) in NSCLC with EGFR exon 20 insertions (exon20ins). 117. Schapira M, Calabrese MF, Bullock AN, Crews CM. Targeted protein
Ann Oncol 2020;31:S815–S6. degradation: expanding the toolbox. Nat Rev Drug Discov 2019;18:
99. Park K, John T, Kim S-W, Lee JS, Shu CA, Kim D-W, et al. Amivan- 949–63.
tamab (JNJ-61186372), an anti-EGFR-MET bispecific antibody, in 118. Sequist LV, Han JY, Ahn MJ, Cho BC, Yu H, Kim SW, et al. Osimer-
patients with EGFR exon 20 insertion (exon20ins)-mutated non- tinib plus savolitinib in patients with EGFR mutation-positive,
small cell lung cancer (NSCLC). J Clin Oncol 2020;38:9512. MET-amplified, non-small-cell lung cancer after progression on
100. Yun J, Lee SH, Kim SY, Jeong SY, Kim JH, Pyo KH, et al. Antitumor EGFR tyrosine kinase inhibitors: interim results from a multicentre,
activity of amivantamab (JNJ-61186372), an EGFR-MET bispecific open-label, phase 1b study. Lancet Oncol 2020;21:373–86.
antibody, in diverse models of EGFR exon 20 insertion-driven 119. Piotrowska Z, Isozaki H, Lennerz JK, Gainor JF, Lennes IT, Zhu VW,
NSCLC. Cancer Discov 2020;10:1194–209. et al. Landscape of acquired resistance to osimertinib in EGFR-
101. Andre F, Ciruelos E, Rubovszky G, Campone M, Loibl S, Rugo HS, mutant NSCLC and clinical validation of combined EGFR and RET
et al. Alpelisib for PIK3CA-mutated, hormone receptor-positive inhibition with osimertinib and BLU-667 for acquired RET fusion.
advanced breast cancer. N Engl J Med 2019;380:1929–40. Cancer Discov 2018;8:1529–39.
102. Kopetz S, Grothey A, Yaeger R, Van Cutsem E, Desai J, Yoshino T, 120. Park JH, Choi YJ, Kim SY, Lee JE, Sung KJ, Park S, et al. Activation
et al. Encorafenib, binimetinib, and cetuximab in BRAF V600E- of the IGF1R pathway potentially mediates acquired resistance
mutated colorectal cancer. N Engl J Med 2019;381:1632–43. to mutant-selective 3rd-generation EGF receptor tyrosine kinase
103. Narita Y, Okamoto K, Kawada MI, Takase K, Minoshima Y, inhibitors in advanced non-small cell lung cancer. Oncotarget
Kodama K, et al. Novel ATP-competitive MEK inhibitor E6201 is 2016;7:22005–15.

894 | CANCER DISCOVERY 10TH ANNIVERSARY ISSUE April 2021 AACRJournals.org

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
Overcoming Resistance to Targeted Therapy and Immunotherapy REVIEW

121. Huang Y, Gan J, Guo K, Deng Y, Fang W. Acquired BRAF V600E 140. Huang D, Ding Y, Zhou M, Rini BI, Petillo D, Qian CN, et al. Inter-
mutation mediated resistance to osimertinib and responded to osi- leukin-8 mediates resistance to antiangiogenic agent sunitinib in
mertinib, dabrafenib, and trametinib combination therapy. J Thorac renal cell carcinoma. Cancer Res 2010;70:1063–71.
Oncol 2019;14:e236–e7. 141. Xue W, Du X, Wu H, Liu H, Xie T, Tong H, et al. Aberrant glioblas-
122. Meng P, Koopman B, Kok K, Ter Elst A, Schuuring E, van Kem- toma neovascularization patterns and their correlation with DCE-
pen LC, et al. Combined osimertinib, dabrafenib and trametinib MRI-derived parameters following temozolomide and bevacizumab
treatment for advanced non-small-cell lung cancer patients with treatment. Sci Rep 2017;7:13894.
an osimertinib-induced BRAF V600E mutation. Lung Cancer 142. Moserle L, Jimenez-Valerio G, Casanovas O. Antiangiogenic thera-
2020;146:358–61. pies: going beyond their limits. Cancer Discov 2014;4:31–41.
123. Dagogo-Jack I, Piotrowska Z, Cobb R, Banwait M, Lennerz JK, 143. de Charette M, Marabelle A, Houot R. Turning tumour cells into
Hata AN, et al. Response to the combination of osimertinib and antigen presenting cells: the next step to improve cancer immuno-
trametinib in a patient with EGFR-mutant NSCLC harboring an therapy? Eur J Cancer 2016;68:134–47.
acquired BRAF fusion. J Thorac Oncol 2019;14:e226–e8. 144. Mehta A, Kim YJ, Robert L, Tsoi J, Comin-Anduix B, Berent-Maoz
124. Haura EB, Hicks JK, Boyle TA. Erdafitinib overcomes FGFR3- B, et al. Immunotherapy resistance by inflammation-induced dedif-
TACC3-mediated resistance to osimertinib. J Thorac Oncol 2020; ferentiation. Cancer Discov 2018;8:935–43.
15:e154–e6. 145. Rosenthal R, Cadieux EL, Salgado R, Bakir MA, Moore DA, Hiley
125. Grugan KD, Dorn K, Jarantow SW, Bushey BS, Pardinas JR, CT, et al. Neoantigen-directed immune escape in lung cancer evolu-
Laquerre S, et al. Fc-mediated activity of EGFR x c-Met bispecific tion. Nature 2019;567:479–85.
antibody JNJ-61186372 enhanced killing of lung cancer cells. mAbs 146. Anagnostou V, Smith KN, Forde PM, Niknafs N, Bhattacharya R,
2017;9:114–26. White J, et al. Evolution of neoantigen landscape during immune
126. Facchinetti F, Lacroix L, Mezquita L, Scoazec JY, Loriot Y, Tselikas L, checkpoint blockade in non-small cell lung cancer. Cancer Discov
et al. Molecular mechanisms of resistance to BRAF and MEK 2017;7:264–76.
inhibitors in BRAF(V600E) non-small cell lung cancer. Eur J Cancer 147. Lee JH, Shklovskaya E, Lim SY, Carlino MS, Menzies AM, Stewart A,
2020;132:211–23. et al. Transcriptional downregulation of MHC class I and melanoma
127. Su F, Viros A, Milagre C, Trunzer K, Bollag G, Spleiss O, et al. de-differentiation in resistance to PD-1 inhibition. Nat Commun
RAS mutations in cutaneous squamous-cell carcinomas in patients 2020;11:1897.
treated with BRAF inhibitors. N Engl J Med 2012;366:207–15. 148. McGranahan N, Furness AJ, Rosenthal R, Ramskov S, Lyngaa R,
128. Robert C, Karaszewska B, Schachter J, Rutkowski P, Mackiewicz A, Saini SK, et al. Clonal neoantigens elicit T cell immunoreactivity
Stroiakovski D, et al. Improved overall survival in melanoma with and sensitivity to immune checkpoint blockade. Science 2016;351:
combined dabrafenib and trametinib. N Engl J Med 2015;372:30–9. 1463–9.
129. Bueno MJ, Mouron S, Quintela-Fandino M. Personalising and 149. Patel SJ, Sanjana NE, Kishton RJ, Eidizadeh A, Vodnala SK, Cam M,
targeting antiangiogenic resistance: a complex and multifactorial et al. Identification of essential genes for cancer immunotherapy.
approach. Br J Cancer 2017;116:1119–25. Nature 2017;548:537–42.
130. Choueiri TK, Escudier B, Powles T, Tannir NM, Mainwaring PN, 150. Sade-Feldman M, Jiao YJ, Chen JH, Rooney MS, Barzily-Rokni M,
Rini BI, et al. Cabozantinib versus everolimus in advanced renal cell Eliane JP, et al. Resistance to checkpoint blockade therapy through
carcinoma (METEOR): final results from a randomised, open-label, inactivation of antigen presentation. Nat Commun 2017;8:1136.
phase 3 trial. Lancet Oncol 2016;17:917–27. 151. Middha S, Yaeger R, Shia J, Stadler ZK, King S, Guercio S, et al.
131. Comito G, Giannoni E, Segura CP, Barcellos-de-Souza P, Raspollini Majority of B2M-mutant and -deficient colorectal carcinomas
MR, Baroni G, et al. Cancer-associated fibroblasts and M2-polarized achieve clinical benefit from immune checkpoint inhibitor therapy
macrophages synergize during prostate carcinoma progression. and are microsatellite instability-high. JCO Precis Oncol 2019;3:
Oncogene 2014;33:2423–31. PO.18.00321.
132. Chung AS, Wu X, Zhuang G, Ngu H, Kasman I, Zhang J, et al. An 152. Zaretsky JM, Garcia-Diaz A, Shin DS, Escuin-Ordinas H, Hugo W,
interleukin-17-mediated paracrine network promotes tumor resist- Hu-Lieskovan S, et al. Mutations associated with acquired resistance
ance to anti-angiogenic therapy. Nat Med 2013;19:1114–23. to PD-1 blockade in melanoma. N Engl J Med 2016;375:819–29.
133. Conley SJ, Gheordunescu E, Kakarala P, Newman B, Korkaya H, 153. Gettinger S, Choi J, Hastings K, Truini A, Datar I, Sowell R, et al.
Heath AN, et al. Antiangiogenic agents increase breast cancer stem Impaired HLA class I antigen processing and presentation as a
cells via the generation of tumor hypoxia. Proc Natl Acad Sci U S A mechanism of acquired resistance to immune checkpoint inhibitors
2012;109:2784–9. in lung cancer. Cancer Discov 2017;7:1420–35.
134. Qiang L, Wu T, Zhang HW, Lu N, Hu R, Wang YJ, et al. HIF-1alpha 154. Le DT, Durham JN, Smith KN, Wang H, Bartlett BR, Aulakh LK,
is critical for hypoxia-mediated maintenance of glioblastoma stem et al. Mismatch repair deficiency predicts response of solid tumors
cells by activating Notch signaling pathway. Cell Death Differ to PD-1 blockade. Science 2017;357:409–13.
2012;19:284–94. 155. McGranahan N, Favero F, de Bruin EC, Birkbak NJ, Szallasi Z,
135. Folkins C, Shaked Y, Man S, Tang T, Lee CR, Zhu Z, et al. Glioma Swanton C. Clonal status of actionable driver events and the tim-
tumor stem-like cells promote tumor angiogenesis and vasculo- ing of mutational processes in cancer evolution. Sci Transl Med
genesis via vascular endothelial growth factor and stromal-derived 2015;7:283ra54.
factor 1. Cancer Res 2009;69:7243–51. 156. Keenan TE, Burke KP, Van Allen EM. Genomic correlates of response
136. Jaiprasart P, Dogra S, Neelakantan D, Devapatla B, Woo S. Identifi- to immune checkpoint blockade. Nat Med 2019;25:389–402.
cation of signature genes associated with therapeutic resistance to 157. Gao J, Shi LZ, Zhao H, Chen J, Xiong L, He Q, et al. Loss of IFN-
anti-VEGF therapy. Oncotarget 2020;11:99–114. gamma pathway genes in tumor cells as a mechanism of resistance
137. Zhang J, Liu Q, Fang Z, Hu X, Huang F, Tang L, et al. Hypoxia to anti-CTLA-4 therapy. Cell 2016;167:397–404.
induces the proliferation of endothelial progenitor cells via upregu- 158. Shin DS, Zaretsky JM, Escuin-Ordinas H, Garcia-Diaz A, Hu-Liesk-
lation of Apelin/APLNR/MAPK signaling. Mol Med Rep 2016;13: ovan S, Kalbasi A, et al. Primary resistance to PD-1 blockade medi-
1801–6. ated by JAK1/2 mutations. Cancer Discov 2017;7:188–201.
138. Eichten A, Su J, Adler AP, Zhang L, Ioffe E, Parveen AA, et al. Resist- 159. Sucker A, Zhao F, Pieper N, Heeke C, Maltaner R, Stadtler N, et al.
ance to anti-VEGF therapy mediated by autocrine IL6/STAT3 sign- Acquired IFNgamma resistance impairs anti-tumor immunity
aling and overcome by IL6 blockade. Cancer Res 2016;76:2327–39. and gives rise to T-cell-resistant melanoma lesions. Nat Commun
139. Zhou Q, Perakis SO, Ulz P, Mohan S, Riedl JM, Talakic E, et al. Cell- 2017;8:15440.
free DNA analysis reveals POLR1D-mediated resistance to bevaci- 160. Albacker LA, Wu J, Smith P, Warmuth M, Stephens PJ, Zhu P, et al.
zumab in colorectal cancer. Genome Med 2020;12:20. Loss of function JAK1 mutations occur at high frequency in cancers

April 2021 CANCER DISCOVERY 10TH ANNIVERSARY ISSUE | 895

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
REVIEW Aldea et al.

with microsatellite instability and are suggestive of immune eva- 180. Cerezo M, Guemiri R, Druillennec S, Girault I, Malka-Mahieu H,
sion. PLoS One 2017;12:e0176181. Shen S, et al. Translational control of tumor immune escape via the
161. Stelloo E, Versluis MA, Nijman HW, de Bruyn M, Plat A, Osse EM, eIF4F-STAT1-PD-L1 axis in melanoma. Nat Med 2018;24:1877–86.
et al. Microsatellite instability derived JAK1 frameshift mutations 181. Bakhoum SF, Ngo B, Laughney AM, Cavallo JA, Murphy CJ, Ly P,
are associated with tumor immune evasion in endometrioid endo- et al. Chromosomal instability drives metastasis through a cytosolic
metrial cancer. Oncotarget 2016;7:39885–93. DNA response. Nature 2018;553:467–72.
162. Horn S, Leonardelli S, Sucker A, Schadendorf D, Griewank KG, 182. Dunphy G, Flannery SM, Almine JF, Connolly DJ, Paulus C,
­Paschen A. Tumor CDKN2A-associated JAK2 loss and susceptibility ­Jonsson KL, et al. Non-canonical activation of the DNA sensing
to immunotherapy resistance. J Natl Cancer Inst 2018;110:677–81. adaptor STING by ATM and IFI16 mediates NF-kappaB signaling
163. Ye Z, Dong H, Li Y, Ma T, Huang H, Leong HS, et al. Prevalent after nuclear DNA damage. Mol Cell 2018;71:745–60.
homozygous deletions of type I interferon and defensin genes 183. Pellegrino B, Musolino A, Llop-Guevara A, Serra V, De Silva P,
in human cancers associate with immunotherapy resistance. Clin Hlavata Z, et al. Homologous recombination repair deficiency and
­Cancer Res 2018;24:3299–308. the immune response in breast cancer: a literature review. Transl
164. Kwon J, Bakhoum SF. The cytosolic DNA-sensing cGAS-STING Oncol 2020;13:410–22.
pathway in cancer. Cancer Discov 2020;10:26–39. 184. Braun DA, Ishii Y, Walsh AM, Van Allen EM, Wu CJ, Shukla SA,
165. Spranger S, Spaapen RM, Zha Y, Williams J, Meng Y, Ha TT, et al. et al. Clinical validation of PBRM1 alterations as a marker of immune
Up-regulation of PD-L1, IDO, and T(regs) in the melanoma tumor checkpoint inhibitor response in renal cell carcinoma. JAMA Oncol
microenvironment is driven by CD8(+) T cells. Sci Transl Med 2013; 2019;5:1631–3.
5:200ra116. 185. Zhou H, Liu J, Zhang Y, Huang Y, Shen J, Yang Y, et al. PBRM1
166. Spranger S, Bao R, Gajewski TF. Melanoma-intrinsic beta-catenin mutation and preliminary response to immune checkpoint block-
signalling prevents anti-tumour immunity. Nature 2015;523:231–5. ade treatment in non-small cell lung cancer. NPJ Precis Oncol
167. Skoulidis F, Goldberg ME, Greenawalt DM, Hellmann MD, Awad 2020;4:6.
MM, Gainor JF, et al. STK11/LKB1 mutations and PD-1 inhibitor 186. Lee HM, Timme TL, Thompson TC. Resistance to lysis by cytotoxic
resistance in KRAS-mutant lung adenocarcinoma. Cancer Discov T cells: a dominant effect in metastatic mouse prostate cancer cells.
2018;8:822–35. Cancer Res 2000;60:1927–33.
168. Spranger S, Dai D, Horton B, Gajewski TF. Tumor-residing Batf3 187. Weide B, Martens A, Hassel JC, Berking C, Postow MA, Bisschop K,
dendritic cells are required for effector T cell trafficking and adop- et al. Baseline biomarkers for outcome of melanoma patients treated
tive T cell therapy. Cancer Cell 2017;31:711–23. with pembrolizumab. Clin Cancer Res 2016;22:5487–96.
169. George S, Miao D, Demetri GD, Adeegbe D, Rodig SJ, Shukla S, et al. 188. Silva I, Tasker A, Quek C, Rawson R, Lim SY, Wang K, et al. Abstract
Loss of PTEN is associated with resistance to anti-PD-1 checkpoint 975: liver metastases (mets) induce systemic immunosuppression
blockade therapy in metastatic uterine leiomyosarcoma. Immunity and immunotherapy resistance in metastatic melanoma. Cancer Res
2017;46:197–204. 2019;79:975.
170. Trujillo JA, Luke JJ, Zha Y, Segal JP, Ritterhouse LL, Spranger S, 189. Tumeh PC, Hellmann MD, Hamid O, Tsai KK, Loo KL, Gubens MA,
et al. Secondary resistance to immunotherapy associated with beta- et al. Liver metastasis and treatment outcome with anti-PD-1 mono-
catenin pathway activation or PTEN loss in metastatic melanoma. clonal antibody in patients with melanoma and NSCLC. Cancer
J Immunother Cancer 2019;7:295. Immunol Res 2017;5:417–24.
171. Peng W, Chen JQ, Liu C, Malu S, Creasy C, Tetzlaff MT, et al. Loss 190. Kudo Y, Haymaker C, Zhang J, Reuben A, Duose DY, Fujimoto J,
of PTEN promotes resistance to T cell-mediated immunotherapy. et al. Suppressed immune microenvironment and repertoire in
Cancer Discov 2016;6:202–16. brain metastases from patients with resected non-small-cell lung
172. Vidotto T, Melo CM, Castelli E, Koti M, Dos Reis RB, Squire JA. cancer. Ann Oncol 2019;30:1521–30.
Emerging role of PTEN loss in evasion of the immune response to 191. Fischer GM, Jalali A, Kircher DA, Lee WC, McQuade JL, Haydu LE,
tumours. Br J Cancer 2020;122:1732–43. et al. Molecular profiling reveals unique immune and metabolic
173. Voron T, Marcheteau E, Pernot S, Colussi O, Tartour E, Taieb J, et al. features of melanoma brain metastases. Cancer Discov 2019;9:
Control of the immune response by pro-angiogenic factors. Front 628–45.
Oncol 2014;4:70. 192. Zhang L, Yao J, Wei Y, Zhou Z, Li P, Qu J, et al. Blocking immuno-
174. Skoulidis F, Arbour KC, Hellmann MD, Patil PD, Marmarelis ME, suppressive neutrophils deters pY696-EZH2-driven brain metasta-
Awad MM, et al. Association of STK11/LKB1 genomic alterations ses. Sci Transl Med 2020;12:eaaz5387.
with lack of benefit from the addition of pembrolizumab to plati- 193. Murciano-Goroff YR, Warner AB, Wolchok JD. The future of cancer
num doublet chemotherapy in non-squamous non-small cell lung immunotherapy: microenvironment-targeting combinations. Cell
cancer. J Clin Oncol 2019;37:102. Res 2020;30:507–19.
175. Papillon-Cavanagh S, Doshi P, Dobrin R, Szustakowski J, Walsh AM. 194. Thommen DS, Schreiner J, Muller P, Herzig P, Roller A, Belousov A,
STK11 and KEAP1 mutations as prognostic biomarkers in an et al. Progression of lung cancer is associated with increased dys-
observational real-world lung adenocarcinoma cohort. ESMO open function of T cells defined by coexpression of multiple inhibitory
2020;5:e000706. receptors. Cancer Immunol Res 2015;3:1344–55.
176. Marinelli D, Mazzotta M, Scalera S, Terrenato I, Sperati F, 195. Koyama S, Akbay EA, Li YY, Herter-Sprie GS, Buczkowski KA,
D’Ambrosio L, et al. KEAP1-driven co-mutations in lung adeno- ­Richards WG, et al. Adaptive resistance to therapeutic PD-1 block-
carcinoma unresponsive to immunotherapy despite high tumor ade is associated with upregulation of alternative immune check-
mutational burden. Ann Oncol 2020;31:1746–54. points. Nat Commun 2016;7:10501.
177. Kitajima S, Ivanova E, Guo S, Yoshida R, Campisi M, Sundararaman 196. Loo Yau H, Ettayebi I, De Carvalho DD. The cancer epigenome:
SK, et al. Suppression of STING associated with LKB1 loss in KRAS- exploiting its vulnerabilities for immunotherapy. Trends Cell Biol
driven lung cancer. Cancer Discov 2019;9:34–45. 2019;29:31–43.
178. Mazieres J, Drilon A, Lusque A, Mhanna L, Cortot AB, Mezquita L, 197. Tanaka A, Sakaguchi S. Targeting Treg cells in cancer immuno-
et al. Immune checkpoint inhibitors for patients with advanced therapy. Eur J Immunol 2019;49:1140–6.
lung cancer and oncogenic driver alterations: results from the 198. Derynck R, Turley SJ, Akhurst RJ. TGFbeta biology in cancer pro-
IMMUNOTARGET registry. Ann Oncol 2019;30:1321–8. gression and immunotherapy. Nat Rev Clin Oncol 2021;18:9–34.
179. Hastings K, Yu HA, Wei W, Sanchez-Vega F, DeVeaux M, Choi J, 199. Jiao S, Subudhi SK, Aparicio A, Ge Z, Guan B, Miura Y, et al.
et al. EGFR mutation subtypes and response to immune check- Differences in tumor microenvironment dictate T helper lineage
point blockade treatment in non-small-cell lung cancer. Ann Oncol polarization and response to immune checkpoint therapy. Cell
2019;30:1311–20. 2019;179:1177–90.

896 | CANCER DISCOVERY 10TH ANNIVERSARY ISSUE April 2021 AACRJournals.org

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
Overcoming Resistance to Targeted Therapy and Immunotherapy REVIEW

200. Chang CH, Qiu J, O’Sullivan D, Buck MD, Noguchi T, Curtis JD, 219. Pucino V, Bombardieri M, Pitzalis C, Mauro C. Lactate at the
et al. Metabolic competition in the tumor microenvironment is a crossroads of metabolism, inflammation, and autoimmunity. Eur J
driver of cancer progression. Cell 2015;162:1229–41. Immunol 2017;47:14–21.
201. AbdulJabbar K, Raza SEA, Rosenthal R, Jamal-Hanjani M, Veeriah S, 220. Labriola MK, Zhu J, Gupta R, McCall S, Jackson J, Kong EF,
Akarca A, et al. Geospatial immune variability illuminates differen- et al. Characterization of tumor mutation burden, PD-L1 and DNA
tial evolution of lung adenocarcinoma. Nat Med 2020;26:1054–62. repair genes to assess relationship to immune checkpoint inhibitors
202. Castro A, Pyke RM, Zhang X, Thompson WK, Day CP, Alexandrov LB, response in metastatic renal cell carcinoma. J Immunother Cancer
et al. Strength of immune selection in tumors varies with sex and 2020;8:e000319.
age. Nat Commun 2020;11:4128. 221. Petitprez F, de Reynies A, Keung EZ, Chen TW, Sun CM, Calderaro
203. Chaput N, Lepage P, Coutzac C, Soularue E, Le Roux K, Monot C, J, et al. B cells are associated with survival and immunotherapy
et al. Baseline gut microbiota predicts clinical response and colitis in response in sarcoma. Nature 2020;577:556–60.
metastatic melanoma patients treated with ipilimumab. Ann Oncol 222. Barata P, Agarwal N, Nussenzveig R, Gerendash B, Jaeger E, Hatton W,
2017;28:1368–79. et al. Clinical activity of pembrolizumab in metastatic prostate can-
204. Gopalakrishnan V, Spencer CN, Nezi L, Reuben A, Andrews MC, cer with microsatellite instability high (MSI-H) detected by circulat-
Karpinets TV, et al. Gut microbiome modulates response to anti- ing tumor DNA. J Immunother Cancer 2020;8:e001065.
PD-1 immunotherapy in melanoma patients. Science 2018;359: 223. Antonarakis ES, Isaacsson Velho P, Fu W, Wang H, Agarwal N,
97–103. Sacristan Santos V, et al. CDK12-altered prostate cancer: clinical
205. Fluckiger A, Daillere R, Sassi M, Sixt BS, Liu P, Loos F, et al. Cross- features and therapeutic outcomes to standard systemic therapies,
reactivity between tumor MHC class I-restricted antigens and an poly (ADP-Ribose) polymerase inhibitors, and PD-1 inhibitors. JCO
enterococcal bacteriophage. Science 2020;369:936–42. Precis Oncol 2020;4:370–81.
206. Coutzac C, Jouniaux JM, Paci A, Schmidt J, Mallardo D, Seck A, 224. Galluzzi L, Vitale I, Warren S, Adjemian S, Agostinis P, Martinez AB,
et al. Systemic short chain fatty acids limit antitumor effect of et al. Consensus guidelines for the definition, detection and inter-
CTLA-4 blockade in hosts with cancer. Nat Commun 2020;11: pretation of immunogenic cell death. J Immunother Cancer 2020;8:
2168. e000337.
207. Sun R, Champiat S, Dercle L, Aspeslagh S, Castanon E, Limkin EJ, 225. Brooks ED, Chang JY. Time to abandon single-site irradiation for
et al. Baseline lymphopenia should not be used as exclusion criteria inducing abscopal effects. Nat Rev Clin Oncol 2019;16:123–35.
in early clinical trials investigating immune checkpoint blockers 226. Gandhi L, Rodriguez-Abreu D, Gadgeel S, Esteban E, Felip E,
(PD-1/PD-L1 inhibitors). Eur J Cancer 2017;84:202–11. De Angelis F, et al. Pembrolizumab plus chemotherapy in metastatic
208. Cassidy MR, Wolchok RE, Zheng J, Panageas KS, Wolchok JD, Coit D, non-small-cell lung cancer. N Engl J Med 2018;378:2078–92.
et al. Neutrophil to lymphocyte ratio is associated with outcome 227. Rassy E, Flippot R, Albiges L. Tyrosine kinase inhibitors and immu-
during ipilimumab treatment. EBioMedicine 2017;18:56–61. notherapy combinations in renal cell carcinoma. Ther Adv Med
209. Bigot F, Castanon E, Baldini C, Hollebecque A, Carmona A, Postel- Oncol 2020;12:1758835920907504.
Vinay S, et al. Prospective validation of a prognostic score for 228. Voron T, Colussi O, Marcheteau E, Pernot S, Nizard M, Pointet AL,
patients in immunotherapy phase I trials: The Gustave Roussy et al. VEGF-A modulates expression of inhibitory checkpoints on
Immune Score (GRIm-Score). Eur J Cancer 2017;84:212–8. CD8+ T cells in tumors. J Exp Med 2015;212:139–48.
210. Ferrucci PF, Gandini S, Battaglia A, Alfieri S, Di Giacomo AM, Gian- 229. Terme M, Pernot S, Marcheteau E, Sandoval F, Benhamouda N,
narelli D, et al. Baseline neutrophil-to-lymphocyte ratio is associ- Colussi O, et al. VEGFA-VEGFR pathway blockade inhibits tumor-
ated with outcome of ipilimumab-treated metastatic melanoma induced regulatory T-cell proliferation in colorectal cancer. Cancer
patients. Br J Cancer 2015;112:1904–10. Res 2013;73:539–49.
211. Gauci ML, Lanoy E, Champiat S, Caramella C, Ammari S, Aspeslagh 230. Postow MA, Callahan MK, Barker CA, Yamada Y, Yuan J, Kitano S,
S, et al. Long-term survival in patients responding to anti-PD-1/ et al. Immunologic correlates of the abscopal effect in a patient with
PD-L1 therapy and disease outcome upon treatment discontinua- melanoma. N Engl J Med 2012;366:925–31.
tion. Clin Cancer Res 2019;25:946–56. 231. Theelen WSME, Peulen HMU, Lalezari F, van der Noort V, de Vries
212. Mezquita L, Auclin E, Ferrara R, Charrier M, Remon J, Planchard JF, Aerts JGJV, et al. Effect of pembrolizumab after stereotactic
D, et al. Association of the lung immune prognostic index with body radiotherapy vs pembrolizumab alone on tumor response
immune checkpoint inhibitor outcomes in patients with advanced in patients with advanced non–small cell lung cancer: results of
non-small cell lung cancer. JAMA Oncol 2018;4:351–7. the PEMBRO-RT phase 2 randomized clinical trial. JAMA Oncol
213. Laino AS, Woods D, Vassallo M, Qian X, Tang H, Wind-Rotolo M, 2019;5:1276–82.
et al. Serum interleukin-6 and C-reactive protein are associated with 232. Marciscano AE, Ghasemzadeh A, Nirschl TR, Theodros D, Kochel
survival in melanoma patients receiving immune checkpoint inhibi- CM, Francica BJ, et al. Elective nodal irradiation attenuates the
tion. J Immunother Cancer 2020;8:e000842. combinatorial efficacy of stereotactic radiation therapy and immu-
214. Schalper KA, Carleton M, Zhou M, Chen T, Feng Y, Huang SP, notherapy. Clin Cancer Res 2018;24:5058–71.
et al. Elevated serum interleukin-8 is associated with enhanced 233. Liang H, Deng L, Hou Y, Meng X, Huang X, Rao E, et al. Host
intratumor neutrophils and reduced clinical benefit of immune- STING-dependent MDSC mobilization drives extrinsic radiation
checkpoint inhibitors. Nat Med 2020;26:688–92. resistance. Nat Commun 2017;8:1736.
215. Kim JY, Lee KH, Kang J, Borcoman E, Saada-Bouzid E, Kronbichler 234. Chon HJ, Lee WS, Yang H, Kong SJ, Lee NK, Moon ES, et al. Tumor
A, et al. Hyperprogressive disease during anti-PD-1 (PDCD1)/PD-L1 microenvironment remodeling by intratumoral oncolytic vaccinia
(CD274) therapy: a systematic review and meta-analysis. Cancers virus enhances the efficacy of immune-checkpoint blockade. Clin
2019;11:1699. Cancer Res 2019;25:1612–23.
216. Diem S, Kasenda B, Martin-Liberal J, Lee A, Chauhan D, Gore M, 235. Brown MC, Holl EK, Boczkowski D, Dobrikova E, Mosaheb M,
et al. Prognostic score for patients with advanced melanoma treated Chandramohan V, et al. Cancer immunotherapy with recombinant
with ipilimumab. Eur J Cancer 2015;51:2785–91. poliovirus induces IFN-dominant activation of dendritic cells and
217. Ding J, Karp JE, Emadi A. Elevated lactate dehydrogenase (LDH) can tumor antigen-specific CTLs. Sci Transl Med 2017;9:eaan4220.
be a marker of immune suppression in cancer: interplay between 236. Aleynick M, Svensson-Arvelund J, Flowers CR, Marabelle A, Brody
hematologic and solid neoplastic clones and their microenviron- JD. Pathogen molecular pattern receptor agonists: treating cancer
ments. Cancer Biomark 2017;19:353–63. by mimicking infection. Clin Cancer Res 2019;25:6283–94.
218. Zhang W, Wang G, Xu ZG, Tu H, Hu F, Dai J, et al. Lactate is 237. Zhou H, Forveille S, Sauvat A, Yamazaki T, Senovilla L, Ma Y, et al.
a natural suppressor of RLR signaling by targeting MAVS. Cell The oncolytic peptide LTX-315 triggers immunogenic cell death.
2019;178:176–89. Cell Death Dis 2016;7:e2134.

April 2021 CANCER DISCOVERY 10TH ANNIVERSARY ISSUE | 897

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
REVIEW Aldea et al.

238. Letourneur D, Danlos FX, Marabelle A. Chemokine biology on phase Ib study in patients with advanced solid tumors. Cancer Res
immune checkpoint-targeted therapies. Eur J Cancer 2020;137: 2018;78:CT176–CT.
260–71. 257. Sullivan RJ, Moschos SJ, Johnson ML, Opyrchal M, Ordentlich P,
239. Sautes-Fridman C, Petitprez F, Calderaro J, Fridman WH. Tertiary Brouwer S, et al. Abstract CT072: efficacy and safety of entinostat
lymphoid structures in the era of cancer immunotherapy. Nat Rev (ENT) and pembrolizumab (PEMBRO) in patients with melanoma pre-
Cancer 2019;19:307–25. viously treated with anti-PD1 therapy. Cancer Res 2019;79:CT072–CT.
240. Watanabe K, Kuramitsu S, Posey AD Jr, June CH. Expanding the 258. Finlay BB, Goldszmid R, Honda K, Trinchieri G, Wargo J, Zitvogel L.
therapeutic window for CAR T cell therapy in solid tumors: the Can we harness the microbiota to enhance the efficacy of cancer
knowns and unknowns of CAR T cell biology. Front Immunol immunotherapy? Nat Rev Immunol 2020;20:522–8.
2018;9:2486. 259. Kim EY, Chung TW, Han CW, Park SY, Park KH, Jang SB,
241. Li D, Li X, Zhou WL, Huang Y, Liang X, Jiang L, et al. Genetically et al. A novel lactate dehydrogenase inhibitor, 1-(Phenylseleno)-
engineered T cells for cancer immunotherapy. Signal Transduct 4-(Trifluoromethyl) benzene, suppresses tumor growth through
Target Ther 2019;4:35. apoptotic cell death. Sci Rep 2019;9:3969.
242. Morris EC, Stauss HJ. Optimizing T-cell receptor gene therapy for 260. Vander Velde R, Yoon N, Marusyk V, Durmaz A, Dhawan A,
hematologic malignancies. Blood 2016;127:3305–11. ­Miroshnychenko D, et al. Resistance to targeted therapies as a
243. Louis CU, Savoldo B, Dotti G, Pule M, Yvon E, Myers GD, et al. Antitu- multifactorial, gradual adaptation to inhibitor specific selective
mor activity and long-term fate of chimeric antigen receptor-­positive pressures. Nat Commun 2020;11:2393.
T cells in patients with neuroblastoma. Blood 2011;118:6050–6. 261. Fattore L, Ruggiero CF, Liguoro D, Mancini R, Ciliberto G. Single
244. Gao J, Ward JF, Pettaway CA, Shi LZ, Subudhi SK, Vence LM, et al. cell analysis to dissect molecular heterogeneity and disease evolu-
VISTA is an inhibitory immune checkpoint that is increased after tion in metastatic melanoma. Cell Death Dis 2019;10:827.
ipilimumab therapy in patients with prostate cancer. Nat Med 262. Pailler E, Faugeroux V, Oulhen M, Mezquita L, Laporte M, Honore A,
2017;23:551–5. et al. Acquired resistance mutations to ALK inhibitors identified by
245. Lee JM, Lee MH, Garon E, Goldman JW, Salehi-Rad R, ­Baratelli FE, single circulating tumor cell sequencing in ALK-rearranged non-
et al. Phase I trial of intratumoral injection of CCL21 gene- small-cell lung cancer. Clin Cancer Res 2019;25:6671–82.
modified dendritic cells in lung cancer elicits tumor-specific 263. Dagogo-Jack I, Rooney M, Lin JJ, Nagy RJ, Yeap BY, Hubbeling H,
immune responses and CD8(+) T-cell infiltration. Clin Cancer Res et al. Treatment with next-generation ALK inhibitors fuels plasma
2017;23:4556–68. ALK mutation diversity. Clin Cancer Res 2019;25:6662–70.
246. Fukumura D, Kloepper J, Amoozgar Z, Duda DG, Jain RK. Enhanc- 264. Parikh AR, Leshchiner I, Elagina L, Goyal L, Levovitz C, Siravegna G,
ing cancer immunotherapy using antiangiogenics: opportunities et al. Liquid versus tissue biopsy for detecting acquired resist-
and challenges. Nat Rev Clin Oncol 2018;15:325–40. ance and tumor heterogeneity in gastrointestinal cancers. Nat Med
247. Kantarjian H, Stein A, Gokbuget N, Fielding AK, Schuh AC, Ribera JM, 2019;25:1415–21.
et al. Blinatumomab versus chemotherapy for advanced acute 265. Aldea M, Hendriks L, Mezquita L, Jovelet C, Planchard D, Auclin E,
lymphoblastic leukemia. N Engl J Med 2017;376:836–47. et al. Circulating tumor DNA analysis for patients with oncogene-
248. Rodriguez-Abreu D, Johnson ML, Hussein MA, Cobo M, Patel AJ, addicted NSCLC with isolated central nervous system progression.
Secen NM, et al. Primary analysis of a randomized, double-blind, J Thorac Oncol 2020;15:383–91.
phase II study of the anti-TIGIT antibody tiragolumab (tira) plus 266. Jovelet C, Ileana E, Le Deley MC, Motte N, Rosellini S, Romero A,
atezolizumab (atezo) versus placebo plus atezo as first-line (1L) et al. Circulating cell-free tumor DNA analysis of 50 genes by next-
treatment in patients with PD-L1-selected NSCLC (CITYSCAPE). generation sequencing in the prospective MOSCATO trial. Clin
J Clin Oncol 2020;38:9503. Cancer Res 2016;22:2960–8.
249. Diab A, Tannir NM, Bentebibel SE, Hwu P, Papadimitrakopoulou V, 267. Mezquita L, Swalduz A, Jovelet C, Ortiz-Cuaran S, Howarth K,
Haymaker C, et al. Bempegaldesleukin (NKTR-214) plus nivolumab Planchard D, et al. Clinical relevance of an amplicon-based liquid
in patients with advanced solid tumors: phase I dose-escalation biopsy for detecting ALK and ROS1 fusion and resistance muta-
study of safety, efficacy, and immune activation (PIVOT-02). Cancer tions in patients with non–small-cell lung cancer. JCO Precis Oncol
Discov 2020;10:1158–73. 2020;4:272–82.
250. Knudson KM, Hicks KC, Alter S, Schlom J, Gameiro SR. Mecha- 268. Zviran A, Schulman RC, Shah M, Hill STK, Deochand S, Khamnei
nisms involved in IL-15 superagonist enhancement of anti-PD-L1 CC, et al. Genome-wide cell-free DNA mutational integration ena-
therapy. J Immunother Cancer 2019;7:82. bles ultra-sensitive cancer monitoring. Nat Med 2020;26:1114–24.
251. Hutmacher C, Gonzalo Nunez N, Liuzzi AR, Becher B, Neri D. Tar- 269. Hugo W, Zaretsky JM, Sun L, Song C, Moreno BH, Hu-Lieskovan S,
geted delivery of IL2 to the tumor stroma potentiates the action of et al. Genomic and transcriptomic features of response to anti-PD-1
immune checkpoint inhibitors by preferential activation of NK and therapy in metastatic melanoma. Cell 2017;168:542.
CD8(+) T cells. Cancer Immunol Res 2019;7:572–83. 270. Sun R, Limkin EJ, Vakalopoulou M, Dercle L, Champiat S, Han SR,
252. Ribas A, Medina T, Kummar S, Amin A, Kalbasi A, Drabick JJ, et al. A radiomics approach to assess tumour-infiltrating CD8 cells
et al. SD-101 in combination with pembrolizumab in advanced and response to anti-PD-1 or anti-PD-L1 immunotherapy: an
melanoma: results of a phase Ib, multicenter study. Cancer Discov imaging biomarker, retrospective multicohort study. Lancet Oncol
2018;8:1250–7. 2018;19:1180–91.
253. Wang S, Campos J, Gallotta M, Gong M, Crain C, Naik E, et al. 271. Gerlinger M, Rowan AJ, Horswell S, Math M, Larkin J, Endesfelder D,
Intratumoral injection of a CpG oligonucleotide reverts resistance et al. Intratumor heterogeneity and branched evolution revealed by
to PD-1 blockade by expanding multifunctional CD8+ T cells. Proc multiregion sequencing. N Engl J Med 2012;366:883–92.
Natl Acad Sci U S A 2016;113:E7240–E9. 272. Davies C, Pan H, Godwin J, Gray R, Arriagada R, Raina V, et al. Long-
254. Yu M, Levine SJ. Toll-like receptor, RIG-I-like receptors and the term effects of continuing adjuvant tamoxifen to 10 years versus
NLRP3 inflammasome: key modulators of innate immune responses stopping at 5 years after diagnosis of oestrogen receptor-positive
to double-stranded RNA viruses. Cytokine Growth Factor Rev 2011; breast cancer: ATLAS, a randomised trial. Lancet 2013;381:805–16.
22:63–72. 273. Ramalingam SS, Vansteenkiste J, Planchard D, Cho BC, Gray JE,
255. Torrejon DY, Abril-Rodriguez G, Champhekar AS, Tsoi J, Camp- Ohe Y, et al. Overall survival with osimertinib in untreated, EGFR-
bell KM, Kalbasi A, et al. Overcoming genetically based resistance mutated advanced NSCLC. N Engl J Med 2020;382:41–50.
mechanisms to PD-1 blockade. Cancer Discov 2020;10:1140–57. 274. Herbst RS, Tsuboi M, John T, Grohé C, Majem M, Goldman JW,
256. Kirkwood JM, Iannotti N, Cho D, O’Day S, Gibney G, Hodi FS, et al. Osimertinib as adjuvant therapy in patients (pts) with stage
et al. Abstract CT176: effect of JAK/STAT or PI3Kδ plus PD-1 inhi- IB–IIIA EGFR mutation positive (EGFRm) NSCLC after complete
bition on the tumor microenvironment: biomarker results from a tumor resection: ADAURA. J Clin Oncol 2020;38:LBA5–LBA.

898 | CANCER DISCOVERY 10TH ANNIVERSARY ISSUE April 2021 AACRJournals.org

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
Overcoming Resistance to Targeted Therapy and Immunotherapy REVIEW

275. Soria JC, Ohe Y, Vansteenkiste J, Reungwetwattana T, Chewasku- rial biomarkers to predict responses to immune checkpoint therapy
lyong B, Lee KH, et al. Osimertinib in untreated EGFR-mutated in mUCC. Sci Transl Med 2020;12:eabc4220.
advanced non-small-cell lung cancer. N Engl J Med 2018;378: 288. Pan D, Kobayashi A, Jiang P, Ferrari de Andrade L, Tay RE, Luoma
113–25. AM, et al. A major chromatin regulator determines resistance of
276. Peters S, Camidge DR, Shaw AT, Gadgeel S, Ahn JS, Kim DW, et al. tumor cells to T cell-mediated killing. Science 2018;359:770–5.
Alectinib versus crizotinib in untreated ALK-positive non-small-cell 289. Mitra A, Andrews MC, Roh W, De Macedo MP, Hudgens CW, Cara-
lung cancer. N Engl J Med 2017;377:829–38. peto F, et al. Spatially resolved analyses link genomic and immune
277. Camidge DR, Kim HR, Ahn MJ, Yang JC, Han JY, Lee JS, et al. Brig- diversity and reveal unfavorable neutrophil activation in melanoma.
atinib versus crizotinib in ALK-positive non-small-cell lung cancer. Nat Commun 2020;11:1839.
N Engl J Med 2018;379:2027–39. 290. Long GV, Dummer R, Hamid O, Gajewski TF, Caglevic C, Dalle S,
278. Sicklick JK, Kato S, Okamura R, Schwaederle M, Hahn ME, et al. Epacadostat plus pembrolizumab versus placebo plus pem-
­Williams CB, et al. Molecular profiling of cancer patients enables brolizumab in patients with unresectable or metastatic melanoma
personalized combination therapy: the I-PREDICT study. Nat Med (ECHO-301/KEYNOTE-252): a phase 3, randomised, double-blind
2019;25:744–50. study. Lancet Oncol 2019;20:1083–97.
279. Misale S, Bozic I, Tong J, Peraza-Penton A, Lallo A, Baldi F, et al. Ver- 291. Powley IR, Patel M, Miles G, Pringle H, Howells L, Thomas A,
tical suppression of the EGFR pathway prevents onset of resistance et al. Patient-derived explants (PDEs) as a powerful preclinical
in colorectal cancers. Nat Commun 2015;6:8305. platform for anti-cancer drug and biomarker discovery. Br J Cancer
280. Yoda S, Lin JJ, Lawrence MS, Burke BJ, Friboulet L, Langenbucher 2020;122:735–44.
A, et al. Sequential ALK inhibitors can select for lorlatinib-resistant 292. Blank CU, Haanen JB, Ribas A, Schumacher TN. Cancer immunol-
compound ALK mutations in ALK-positive lung cancer. Cancer ogy. The “cancer immunogram.” Science 2016;352:658–60.
Discov 2018;8:714–29. 293. Barlesi F, Greillier L, Monville F, Foa C, Treut Jl, Audigier-Valette C,
281. Chau CH, Steeg PS, Figg WD. Antibody-drug conjugates for cancer. et al. LBA53 precision immuno-oncology for advanced non-small
Lancet 2019;394:793–804. cell lung cancer (NSCLC) patients (pts) treated with PD1/L1
282. Tsurutani J, Iwata H, Krop I, Janne PA, Doi T, Takahashi S, et al. immune checkpoint inhibitors (ICIs): a first analysis of the PIO-
Targeting HER2 with trastuzumab deruxtecan: a dose-expansion, NeeR study. Ann Oncol 2020;31:S1183.
phase I study in multiple advanced solid tumors. Cancer Discov 294. Gomez DR, Blumenschein GR Jr, Lee JJ, Hernandez M, Ye R,
2020;10:688–701. Camidge DR, et al. Local consolidative therapy versus maintenance
283. Olson ME, Harris RS, Harki DA. APOBEC enzymes as targets for therapy or observation for patients with oligometastatic non-small-
virus and cancer therapy. Cell Chem Biol 2018;25:36–49. cell lung cancer without progression after first-line systemic ther-
284. Streicher K, Higgs BW, Wu S, Coffman K, Damera G, Durham N, apy: a multicentre, randomised, controlled, phase 2 study. Lancet
et al. Increased CD73 and reduced IFNG signature expression in Oncol 2016;17:1672–82.
relation to response rates to anti-PD-1(L1) therapies in EGFR- 295. Palma DA, Olson R, Harrow S, Gaede S, Louie AV, Haasbeek C, et al.
mutant NSCLC. J Clin Oncol 2017;35:11505. Stereotactic ablative radiotherapy versus standard of care palliative
285. Parsa AT, Waldron JS, Panner A, Crane CA, Parney IF, Barry JJ, et al. treatment in patients with oligometastatic cancers (SABR-COMET):
Loss of tumor suppressor PTEN function increases B7-H1 expres- a randomised, phase 2, open-label trial. Lancet 2019;393:2051–8.
sion and immunoresistance in glioma. Nat Med 2007;13:84–8. 296. Liniker E, Menzies AM, Kong BY, Cooper A, Ramanujam S, Lo S,
286. Takaoka A, Hayakawa S, Yanai H, Stoiber D, Negishi H, Kikuchi et al. Activity and safety of radiotherapy with anti–PD-1 drug
H, et al. Integration of interferon-alpha/beta signalling to p53 therapy in patients with metastatic melanoma. Oncoimmunology
responses in tumour suppression and antiviral defence. Nature 2016;5:e1214788.
2003;424:516–23. 297. Borcoman E, Kanjanapan Y, Champiat S, Kato S, Servois V, ­Kurzrock R,
287. Goswami S, Chen Y, Anandhan S, Szabo PM, Basu S, Blando JM, et al. Novel patterns of response under immunotherapy. Ann Oncol
et al. ARID1A mutation plus CXCL13 expression act as combinato- 2019;30:385–96.

April 2021 CANCER DISCOVERY 10TH ANNIVERSARY ISSUE | 899

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
Research.
Overcoming Resistance to Tumor-Targeted and
Immune-Targeted Therapies
Mihaela Aldea, Fabrice Andre, Aurelien Marabelle, et al.

Cancer Discov 2021;11:874-899.

Updated version Access the most recent version of this article at:
http://cancerdiscovery.aacrjournals.org/content/11/4/874

Cited articles This article cites 295 articles, 103 of which you can access for free at:
http://cancerdiscovery.aacrjournals.org/content/11/4/874.full#ref-list-1

E-mail alerts Sign up to receive free email-alerts related to this article or journal.

Reprints and To order reprints of this article or to subscribe to the journal, contact the AACR Publications
Subscriptions Department at pubs@aacr.org.

Permissions To request permission to re-use all or part of this article, use this link
http://cancerdiscovery.aacrjournals.org/content/11/4/874.
Click on "Request Permissions" which will take you to the Copyright Clearance Center's
(CCC)
Rightslink site.

Downloaded from cancerdiscovery.aacrjournals.org on April 4, 2021. © 2021 American Association for Cancer
View publication stats
Research.

You might also like