You are on page 1of 9

Review doi:10.

1038/nature25183

The biology and management of


non-small cell lung cancer
Roy S. Herbst1, Daniel Morgensztern2 & Chris Boshoff1,3

Important advancements in the treatment of non-small cell lung cancer (NSCLC) have been achieved over the past two
decades, increasing our understanding of the disease biology and mechanisms of tumour progression, and advancing
early detection and multimodal care. The use of small molecule tyrosine kinase inhibitors and immunotherapy has led to
unprecedented survival benefits in selected patients. However, the overall cure and survival rates for NSCLC remain low,
particularly in metastatic disease. Therefore, continued research into new drugs and combination therapies is required
to expand the clinical benefit to a broader patient population and to improve outcomes in NSCLC.

L
ung cancer is the most common cause of cancer death worldwide, and the need for better predictors of responses to immunotherapy,
with an estimated 1.6 million deaths each year1. Approximately 85% new drugs and rationally designed drug combination therapies. In this
of patients have a group of histological subtypes collectively known Review, we provide an overview of the recent progress made in lung
as NSCLC, of which lung adenocarcinoma (LUAD) and lung squamous cancer b
­ iology and treatment, including the most promising strategies
cell carcinoma (LUSC) are the most common subtypes2. The most common that have already made a notable impact in outcomes for patients with
aetiology for lung cancer is tobacco smoking, accounting for more than advanced stage NSCLC.
80% of cases in the United States and other countries where smoking
is common3. Although all major histological subtypes of NSCLC, as Biology of lung cancer
well as small cell lung cancer (SCLC), are associated with smoking, the Lung cancer is a molecularly heterogeneous disease and understanding its
association is stronger with LUSC and SCLC than LUAD, with the latter biology is crucial for the development of effective therapies. The treatment
being the most common histology in never smokers4. Lung cancer in of lung cancer has changed from the empirical use of cytotoxic therapy
never smokers is more common in women and in East Asia, and has based on a physician’s preference to a hallmark of personalized ­medicine,
been associated with environmental exposures including second-hand with subsets of patients treated according to the genetic alterations of
smoking, pollution and occupational carcinogens, and with inherited genetic their tumour and the status of programmed death ligand-1 (PD-L1),
susceptibility3,5,6. which predict for benefit from targeted therapies or immune checkpoint
Eradicating the use of all tobacco-related products is a key goal of ­blockers (ICBs), respectively.
the global fight against cancer and requires a comprehensive approach. Similar to most malignancies, lung cancer is composed of sub-populations
Primary prevention efforts include targeting nicotine addiction by of cells, or clones, with distinct molecular features, resulting in intra-­
­providing effective delivery of nicotine without the co-administration tumoral heterogeneity. A larger sub-clonal mutation fraction may be asso-
of carcinogenic chemicals that are present in cigarettes, such as via ciated with increased likelihood of postsurgical relapse in patients with
e-cigarettes7. Other strategies include the use of varenicline, a partial localized LUAD, implying that there is a greater propensity for metastases
agonist of the nicotinic acetylcholine receptor8, counselling and other early d
­ uring tumour development in those tumours with increased intra-­
socio-economic methods including taxation, advertisement and legis- tumoral ­heterogeneity10. The identification of clonal targetable genetic
lative measures, such as lowering the amount of nicotine in cigarettes to alterations occurring early during cancer evolution has changed the
non-addictive levels, a policy recently announced by the US Food and paradigm of treatment for oncogene-addicted cancers, although few
Drug Administration (FDA) (https://www.fda.gov/tobaccoproducts/­ patients, if any, are cured owing to inherent and acquired resistance
newsevents/ucm568425.htm). Despite best intentions, the use of mecha­nisms, with the latter occurring mostly through the selection of
­e-cigarettes to facilitate smoking cessation remains unproven and resistant sub-clones, pre-existing before targeted drug exposure11,12.
controversial given concerns that they might promote the initiation The most common genetic alterations in LUAD and LUSC are shown
of new ­individuals to use smoking devices with unknown long-term in Box 1. Variant allele frequencies for somatic mutations have shown
consequences9. that mutations in the Kirsten rat sarcoma (KRAS) and epidermal growth
Although a crucial component of the fight against lung cancer, tobacco factor receptor (EGFR) genes, when detected, are usually present in the
prevention strategies are not enough to win the war. Increasingly sophis- founder clones, indicating their roles in tumour initiation and repre-
ticated therapies are required to meaningfully improve clinical outcomes senting attractive targets for therapeutic intervention. KRAS and EGFR
for patients. Progress in this area has been substantial and promising over mutations are usually mutually exclusive, but when they co-exist, KRAS
the past 20 years with the advent of various targeted therapies and the mutations may confer resistance to EGFR inhibitors13. Tumours that
effective application of immunotherapy in some populations of patients harbour oncogenic drivers such as EGFR mutations14 and ROS1 and
with advanced NSCLC. Yet, major challenges still remain, including the anaplastic lymphoma kinase (ALK) rearrangements have a lower-than-­
identification of new driver gene alterations to expand the population that average mutation load, mostly owing to occurrence in never or light
benefit from targeted therapies, better understanding of mechanisms of smokers, although the d ­ ominant driving properties of these oncogenes
resistance to targeted therapy to allow them to be prevented or overcome, may reduce the selection pressure for acquiring additional mutations.

1
Yale Cancer Center, Yale School of Medicine, New Haven, Connecticut, USA. 2Washington University School of Medicine, St Louis, Missouri, USA. 3Pfizer, Inc. New York City, New York, USA.

4 4 6 | N A T U R E | V O L 5 5 3 | 2 5 j anuar y 2 0 1 8
© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
Review insight

Box 1
Alterations in targetable ­oncogenic pathways in LUAD and LUSC
Pathway diagram showing the percentage of NSCLC with alterations preclinical lung cancer models, providing a potential therapeutic
involving key pathway components for receptor tyrosine kinase strategy in dual KEAP1- and KRAS-mutant LUAD139.
signalling, mTOR signalling, oxidative stress response, proliferation Common mutated genes in LUSC include the tumour suppressors
and cell cycle progression. The frequency of alterations is based TP53, which is present in more than 90% of tumours, and CDKN2A.
on the sum of somatic mutations, homozygous deletions, focal The latter, which encodes the p16INK4A and p14ARF proteins, is
amplifications, and by significant up- or downregulation of gene inactivated in over 70% of LUSC through epigenetic silencing by
expression (for example, AKT3, FGFR1, PTEN). methylation (21%), inactivating mutation (18%), exon 1β​ skipping
The most commonly mutated genes in LUAD include KRAS and (4%), or homozygous deletion (29%). Although EGFR amplification
EGFR, and the tumour suppressor genes TP53, KEAP1, STK11 and occurs, unlike LUAD, actionable mutations in receptor tyrosine kinases
NF1. The frequency of EGFR-activating mutations varies greatly are rarely observed in LUSC. (Data compiled from refs 14, 17, 22, 45
by region and ethnicity. KEAP1 inactivation in the presence of and diagram adapted from refs 17, 22.)
KRAS mutations confers sensitivity to inhibition of glutaminase in

.(<
EGFR ERBB2/3 MET FGFR1, 2, 3 ALK RET ROS
ROS
Activated 27% <9% 3% 2–4% 7% 7%, 3%, 2% <8% 1% 2% 2%

Inactivated

LUAD
LUSC PTEN KRAS NRAS RASA
3% 15% 32% 3% <1% <1% 4%
PI3CA
4% 16%
PIK3R1 HRAS RIT1 NF1
<1% <1% 3% 2% 11% 11%

STK11 AKT1 AKT2 AKT3


17% 2% 1% <1% 4% 16%
BRAF
7% 4%

TSC1/2
AMPK
3% 3%
MAP2K1
<1% <1%

MTOR
CDKN2A
43% 70%
Proliferation, cell survival, translation

KEAP1 CUL3 MDM2 ATM CCND1 CDK4 CCNE4


19% 12% <1% 7% 8% 9% 4% 7% 3%

NFE2L2 TP53 RB1


3% 19% 46% 90% 7% 7%

Oxidative stress response Proliferation, cell survival Cell cycle progression

Unlike LUAD, actionable mutations in receptor tyrosine kinases are KRAS and TP53. By contrast, never smokers usually have a predomi-
rarely detected in LUSC15. nant transition of cytosine to thymine (C>​T), and a higher prevalence of
Mutations in tumour protein p53 (TP53) are more commonly observed actionable driving gene alterations including activating EGFR mutations,
with advancing grade, suggesting a role during tumour progression16. and ROS1 and ALK translocations14,17.
By contrast, the frequency of KRAS mutations in LUAD seems constant
across tumour grades, suggesting a role in tumour initiation or early Tumour microenvironment
tumorigenesis, and supporting the presence of KRAS alterations in Genetic events that initiate and drive tumour evolution also shape the
founder clones. tumour microenvironment (TME). Therefore, the genetic architecture of
The genomic landscape of lung cancer is markedly distinct between a tumour determines not only the fitness of the cancer cells, but also the
never smokers and smokers, with the latter containing a significantly composition of the TME. NSCLC has a particularly high somatic tumour
higher mutation frequency, predominantly cytosine to adenine (C>​A) mutation burden (TMB), defined as the number of nonsynonymous
nucleotide transversions and non-actionable mutations such as those in coding mutations per megabase, particularly in smokers, who represent

2 5 j anuar y 2 0 1 8 | V O L 5 5 3 | N A T U R E | 4 4 7
© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
insight Review

KEY Crizotinib (first ALK Crizotinib shown Osimertinib Osimertinib shown to


inhibitor) shown to to be superior to shown to be be superior to
Chemotherapy ALK or ROS1 effective in cytotoxic therapy in
be effective in cytotoxic therapy
EGFR Genomic analysis ALK-positive in 1L ALK-positive EGFR(T790M) previously treated
NSCLC NSCLC mutations EGFR(T790M)
Angiogenesis Immune therapy

Mutational burden
TCGA genomic associated with
Docetaxel Pemetrexed RET and ROS characterization responses to
benefit shown in ALK rearrangements shown activity fusions described of LUAD immune checkpoint
2L NSCLC identified in NSCLC in 1L LUAD in LUAD completed blockers

1997 2000 2004 2006 2007 2008 2009 2010 2011 2012 2013 2014 2015 2016 2017

Discovery of EGFR Gefitinib shown TCGA genomic Crizotinib shown Pembrolizumab


mutations in LUAD better activity characterization to be effective in shown to be
sensitive to gefitinib than cytotoxic of LUSC ROS1-positive superior to
and erlotinib therapy in LUAD completed NSCLC cytotoxic therapy
in 1L PD-L1-high
NSCLC

EGFR Bevacizumab Nivolumab shown Immune checkpoint Alectinib shown


inhibitors shown activity with to be effective in blockers shown to to be superior to
enter clinical cytotoxic therapy NSCLC be superior to crizotinib in 1L
development in 1L NSCLC docetaxel in 2L ALK-positive
NSCLC NSCLC

Figure 1 | Timeline illustrating the development of targeted therapies have transformed the management of NSCLC over the past two decades.
and immunotherapies for the treatment of NSCLC over two decades. 1L, first line; 2L, second line; TCGA, The Cancer Genome Atlas.
Timeline highlights some of the pivotal discoveries and clinical studies that

the majority of patients. Overall, the number of mutations is significantly of ­bevacizumab, a monoclonal antibody against vascular endothelial
higher in metastases than in primary lung lesions18. growth factor (VEGF) or necitumumab, an antibody that targets EGFR,
Some mutations create neoantigens, which may be recognized by ­produced modest improvements in survival for patients with non-LUSC
tumour-infiltrating cytotoxic T cells. A high clonal neoantigen burden and LUSC histological subtypes, respectively26,27. In the second-line
in LUAD is associated with an inflamed TME, enriched with activated ­setting, the standard of care has been docetaxel, with or without the anti-
effector T cells and the expression of proteins associated with antigen VEGF receptor-2 antibody ramucirumab28.
presentation, T cell migration (CXCL-10 and CXCL-9), and effector T cell Cytotoxic regimens have demonstrated their greatest effect on earlier
function, as well as negative regulators of T-cell activity including PD-L1, stage disease. Surgical resection is the most effective therapy for stages I
programmed death-1 (PD-1) and lymphocyte activation gene-3 (LAG-3)19. to II and selected cases of stage IIIA NSCLC29. However, despite its cura-
This phenotype may confer sensitivity to treatment with an ICB. Loss of tive intent, a high percentage of tumours will recur, with 5-year overall
mismatch-repair function, which confers the microsatellite instability survival ranging from 83% for stage IA to 36% for stage IIIA disease30.
phenotype, is an extreme example of cancer with a high TMB, with these Adjuvant cytotoxic therapy with a cisplatin-based doublet is associated
tumours demonstrating T cell infiltration and marked responses to ICBs, with improved survival in patients with completely resected stage II and
independent of the tissue of origin20. IIIA NSCLC, with a likely benefit for stage IB disease measuring 4 cm
Genetic alterations may affect the TME in several ways. An example or more31. Morbidity and outcomes have also improved for early stage
is the inactivation of the tumour suppressor serine/threonine kinase 11 lung cancer owing to advances in surgical and radiation technologies,
(STK11; also known as LKB1), occurring in one-third of KRAS-mutated including robotic and video-assisted surgery32, as well as stereotactic and
LUAD, which skews the TME towards the accumulation of immunosup- hyper-fractionated approaches33. The standard therapy for patients with
pressive neutrophils and loss of PD-L1 expression, and is associated with unresectable locally advanced NSCLC is the combination of cytotoxic
fewer tumour-infiltrating lymphocytes21. Large-scale studies to corre- therapy and thoracic radiation, with a survival advantage for concurrent
late the genomes of NSCLC to the cellular constituents of the TME are therapy over sequential approaches34,35.
required to understand how different genotypes determine the cellular The high lung cancer mortality is due to the presence of meta-
make-up of the TME. static disease at the time of diagnosis in most patients, indicating that
improvements in long-term survival will require more effective systemic
Therapeutic advances during the past two decades therapies36,37. Molecularly targeted therapies in patients with NSCLC
Over the past 20 years, treatment has evolved from the empiric use of cyto- were initially used in the late 1990s with the introduction of gefitinib, an
toxic therapies to effective and better tolerated regimens that are targeted oral EGFR tyrosine kinase inhibitor (TKI). In unselected populations,
to specific molecular subtypes in LUAD, and therapies in d ­ evelopment to the response rates were approximately 10%, with increased frequency
target LUSC17,22 (Fig. 1). Platinum-based doublet therapy (for e­ xample, of responses noted in females, never smokers, and patients of Asian
cisplatin in combination with another cytotoxic therapy) has been the descent38,39. Erlotinib, another TKI against EGFR, was associated with
standard therapy for patients with advanced stage NSCLC and good improved survival compared to the best supportive care in patients with
performance status, with the option of maintenance therapy in patients previously treated advanced-stage NSCLC40. Retrospective studies sub-
with non-LUSC histology who achieve tumour control after the initial sequently demonstrated that activating EGFR mutations were observed
four to six cycles23. Overall, there have been no clinically meaningful in the vast majority of patients who benefited from EGFR TKIs41,42. Since
differences in outcome among the multiple cytotoxic regimens used then, additional gene alterations, including ALK rearrangements, ROS1
in patients with advanced stage NSCLC24, with the exception of fusions and BRAF mutations led to the development of effective targeted
­pemetrexed, which is less effective in patients with LUSC25. The ­addition therapies43.

4 4 8 | N A T U R E | V O L 5 5 3 | 2 5 j anuar y 2 0 1 8
© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
Review insight

An important advance in the management of advanced stage NSCLC updates on overall survival, may provide a guide to first-line therapy
occurred in 2015, when the US FDA approved the ICB nivolumab for in previously untreated patients with tumours harbouring an EGFR
the treatment of patients whose disease progressed during or after mutation.
platinum-based therapy, heralding a new era in the management of lung One of the mechanisms for acquired resistance to third-generation
cancer44. EGFR TKIs is the C797S mutation65 which in combination with the sensi-
tizing mutation and without T790M causes resistance to third-­generation
Targeted therapy EGFR TKIs, but not to gefitinib or afatinib. The presence of triple mutants
The identification of targetable gene alterations has transformed the (sensitizing mutation, T790M and C797S) however, leads to resistance
­management of lung cancer, with the incorporation of tumour genotyping to all three generations of EGFR TKIs66. Promising approaches to ­triple
to allow individualized therapy and leading to remarkable responses in mutants include the allosteric inhibitor EAI045 in tumours with ­original
selected patients treated with matched TKIs (Table 1). In the multicentre L858R-sensitizing mutation, and brigatinib, an ALK inhibitor with activity
Lung Cancer Mutation Consortium, targetable oncogenic drivers were against EGFR mutations, in tumours harbouring exon 19 d ­ eletion, both
observed in 64% of patients with LUAD, for whom the use of genotype-­ in combination with cetuximab, an anti-EGFR monoclonal antibody67,68.
directed therapy was associated with improved survival compared to
those treated without targeted therapies45. ALK
ALK encodes a transmembrane receptor tyrosine kinase with unclear
EGFR function in humans69. In ALK rearrangements, the most common partner
EGFR belongs to a receptor tyrosine kinase family that also includes is the echinoderm microtubule-associated protein-like 4 (EML4) gene
human epidermal growth factor receptor 2 (HER2, also known as ERBB2), (EML4-ALK)70. Crizotinib is an oral competitive ATP inhibitor of ALK,
HER3 (ERBB3) and HER4 (ERBB4). The receptor contains four extra­ MET and ROS1 tyrosine kinases with activity against ALK fusion-positive
cellular domains, a transmembrane domain, a tyrosine kinase domain, NSCLC71. Crizotinib is associated with improved ORRs and median PFS
and a carboxy tail46. Binding of activating ligands leads to EGFR dimeri- compared to cytotoxic therapy in both previously treated and untreated
zation and trans-phosphorylation of the tyrosine residues in the carboxy patients72,73. Most patients previously treated with crizotinib benefit
tail, with activation of downstream pathways involved in cell proliferation, from second-generation ALK inhibitors including ceritinib, alectinib
survival, invasion and angiogenesis47. Heterozygous mutations ­clustering and brigatinib74–76. Ceritinib also increased median PFS compared to
around the ATP-binding pocket of the tyrosine kinase domain may lead first-line cytotoxic therapy in patients with ALK-positive NSCLC77. Two
to constitutive EGFR activation and ligand independence41,42. The most randomized studies showed increased ORRs and median PFS for alectinib
common EGFR mutations associated with sensitivity to EGFR TKIs compared to crizotinib in patients with previously untreated ALK-positive
include exon 19 deletions and a missense mutation on exon 21 (L858R)48. NSCLC, establishing alectinib as a first-line treatment option78,79.
First-generation EGFR TKIs, including gefitinib and erlotinib, have Resistance to ALK inhibitors may occur owing to ALK alterations such
shown higher objective response rates (ORRs) and progression-free as mutations and amplification, or upregulation of bypass signalling path-
­survival (PFS) compared to cytotoxic therapy in previously untreated ways including EGFR and mitogen-activated protein kinase (MAPK).
patients with EGFR mutations49–54. In contrast to first-generation EGFR Secondary ALK mutations are the predominant mechanism of resistance
TKIs, which are reversible competitive ATP inhibitors that target only to second-generation TKIs80. The most common ALK resistance mutation
EGFR, second-generation inhibitors including afatinib and dacom- among patients treated with second-generation TKIs is G1202R, which is
itinib are irreversible inhibitors that also target HER2 and HER4. Both
­afatinib and dacomitinib showed improved PFS compared to gefitinib55,56. Table 1 | Selected randomized trials with first-line targeted
Afatinib also demonstrated a significant improvement in the median therapies
overall survival compared to platinum-based cytotoxic therapy in patients PFS
with an exon 19 deletion, but not in those with an L858R mutation57. Study Target Design ORR (%) (months)
Differences in outcomes between these two most common EGFR muta- IPASS49,50 EGFR Gefitinib vs 72.1 vs 47.3 9.5 vs 6.3
tions may occur owing to distinct conformational changes within the carboplatin plus
ATP-binding pocket and patterns of auto-phosphorylation induced by paclitaxel
NEJ00251 EGFR Gefitinib vs 73.7 vs 30.7 10.8 vs 5.4
each mutation58. carboplatin plus
The most common cause of acquired resistance to first-generation paclitaxel
TKIs is a second EGFR mutation in exon 20, with a threonine-to- WJTOG-340552 EGFR Gefitinib vs 62.1 vs 32.2 9.2 vs 6.3
methionine substitution on codon 790 (T790M)59,60. This mutation affects cisplatin plus
the initial EGFR TKI efficacy either from steric hindrance or by increased docetaxel
EURTAC54 EGFR Erlotinib vs 58 vs 15 9.7 vs 5.2
affinity of the tyrosine kinase domain for ATP. Other mechanisms of platinum doublet
resistance include amplifications in HER2 or mutations in MET, BRAF OPTIMAL53 EGFR Erlotinib vs 83 vs 36 13.1 vs 4.6
or phosphatidylinositol-4,5-bisphosphate 3-kinase catalytic subunit alpha carboplatin plus
(PIK3CA), and SCLC transformation61, indicating that repeated mole­ gemcitabine
cular profiling at progression is needed to determine the next appropriate LUX-Lung-755 EGFR Afatinib vs 72.5 vs 56 11 vs 10.9
gefitinib
treatment. Third-generation EGFR TKIs are selective inhibitors of both ARCHER-105056 EGFR Dacomitinib vs 75 vs 72 14.7 vs 9.2
the original sensitizing and T790M mutations, while sparing wild-type gefitinib
EGFR. These drugs bind covalently to cysteine on codon 797, overcoming FLAURA 64
EGFR Osimertinib 80 vs 76 18.9 vs 10.2
the enhanced ATP affinity from the T790M mutation. Osimertinib, a vs gefitinib or
erlotinib
third-generation EGFR TKI, is effective in patients with NSCLC har-
PROFILE -101473 ALK Crizotinib vs 74 vs 45 10.9 vs 7
bouring EGFR(T790M) mutations following progression after first- cisplatin plus
generation EGFR TKI62, and showed increased ORRs and PFS ­compared pemetrexed
to platinum-based cytotoxic therapy63. In a randomized trial that ASCEND-477 ALK Ceritinib vs 72.5 vs 26.7 16.6 vs 8.1
­compared osimertinib to either erlotinib or gefitinib in previously platinum plus
pemetrexed
untreated patients with advanced stage NSCLC harbouring either EGFR
ALEX79 ALK Alectinib vs 82.9 vs 75.5 NR vs 11.1
exon 19 deletion or L858R mutation, osimertinib was associated with crizotinib
a significant improvement in PFS, establishing osimertinib as a first- The IPASS trial encompasses a subgroup of patients with EGFR mutation. NR, not reached; ORR,
line EGFR TKI option64. Additional follow-up from this trial, including objective response rate; PFS, progression-free survival.

2 5 j anuar y 2 0 1 8 | V O L 5 5 3 | N A T U R E | 4 4 9
© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
insight Review

Table 2 | Randomized phase 3 clinical trials comparing ICBs to Immunotherapy


cytotoxic therapy Harnessing the host immune response to treat cancer is not a new
ORR PFS OS concept102, and the introduction of ICBs such as monoclonal antibodies
Study Drug (%) (months) (months) that target cytotoxic T-lymphocyte antigen-4 (CTLA-4) and antibodies
Second-line against PD-1 or PD-L1 have signalled a new direction for lung cancer
CheckMate-017 Nivolumab 20 3.5 9.2 care. The first ICB approved by the US FDA was ipilimumab, a human
(squamous)110 3 mg kg−1 Q2W immunoglobulin G1 monoclonal antibody that blocks CTLA-4, for the
Checkmate-057 Nivolumab 19 2.3 12.2 treatment of metastatic melanoma103,104. During tumorigenesis, PD-1
(non-squamous)111 3 mg kg−1 Q2W
Keynote-010 (NSCLC Pembrolizumab 18 3.9 10.4
signalling, driven primarily by adaptive expression of PD-L1 within
PD-L1 positive)112 2 mg kg−1 Q3W the tumour, inactivates T cells that recognize tumour-specific antigens,
Oak (NSCLC)113 Atezolizumab 14 2.8 13.8 allowing tumour progression and metastasis105,106. Blocking the PD-1
1,200 mg Q3W and PD-L1 axis with antibodies offers an approach to restoring T cell-­
Combined docetaxel Docetaxel 9–13 2.8–4.2 6–9.6 mediated antitumour immunity107–109. ICBs have shown significant
results* (refs 110–113) 75 mg m−2 Q3W
First-line
­benefit in a broad population of patients with NSCLC (Table 2). Current
Keynote-024 Pembrolizumab 44.8 10.3 NR ICBs approved or in development for NSCLC include the anti-PD-1 anti-
(PD-L1 ≥​ 50%)114 200 mg Q3W bodies nivolumab (human IgG4) and pembrolizumab (humanized IgG4),
Checkmate-026 Nivolumab 26 4.2 14.4 as well as the anti-PD-L1 antibodies atezolizumab (human IgG1, with the
(PD-L1 ≥​ 5%)115 3 mg kg−1 Q2W Fc domain engineered to prevent antibody-directed cell cytotoxicity),
Combined Platinum-based 27–33 5.9–6 13.2
chemotherapy results† chemotherapy
durvalumab (human IgG1 engineered), and avelumab (human IgG1
(refs 114, 115) demonstrating preclinical antibody-directed cell cytotoxicity activity)29.
OS, overall survival; Q2W, every two weeks; Q3W, every three weeks. ICBs have been approved as a standard of care for patients with
*Combined control arm results from CheckMate trials 017 and 057, Keynote-010 and Oak. advanced NSCLC whose tumours progress on first-line cytotoxic t­ herapy.
†Combined control arm results from CheckMate-026 and Keynote-024.
Treatment with nivolumab was associated with significantly longer
median overall survival compared to treatment with docetaxel among
associated with in vitro resistance to all currently available ALK inhibitors patients with metastatic NSCLC who had disease progression during or
except for lorlatinib, a potent third-generation ALK inhibitor with activity after platinum-based cytotoxic therapy110,111. Subsequently, other ICBs
against most known ALK resistance mutations and efficacy in patients showed improvement in overall survival compared to treatment with
previously treated with up to three previous lines of ALK inhibitors81. docetaxel, including pembrolizumab112 and atezolizumab113.
In the first-line NSCLC setting, pembrolizumab was established as a
ROS1 new standard of care for patients with advanced or metastatic NSCLC
ROS1 encodes a receptor tyrosine kinase that becomes constitutively with tumour PD-L1 expression levels of 50% or more (by a companion
activated when a rearrangement leads to the fusion of its tyrosine kinase immunohistochemistry test), which is present in up to 30% of NSCLC. In
domain with a partner gene such as CD7482. Owing to the high h ­ omology these patients, pembrolizumab is associated with a significant improve-
between the kinase domains of ROS1 and ALK, drugs used to treat ALK- ment in ORRs, PFS and overall survival compared to platinum-based
positive tumours including crizotinib83, ceritinib84 and lorlatinib81 have cytotoxic therapy114. By contrast, among patients with PD-L1 expression
also shown marked activity in ROS1-positive tumours. Mechanisms levels of 5% or more, nivolumab was not associated with improvements in
of acquired resistance of ROS1 rearrangements to crizotinib include PFS or overall survival compared to cytotoxic therapy115. Among patients
secondary mutations, most commonly G2032R, wild-type EGFR in the nivolumab group with both PD-L1 expression levels above 50% and
­signalling activation, KRAS and KIT mutations85,86. high TMB, the ORR was 75%, confirming previous observations that both
TMB and PD-L1 expression may predict for benefit from ICB. The better
Other alterations toxicity profile with ICBs and equivalent survival outcome could support
Among patients with NSCLC and BRAF mutation, approximately half their selection as first-line therapy for patients with PD-L1-expressing
have a single transversion at exon 15, in which valine is replaced by advanced NSCLC, especially in those not eligible for cytotoxic therapy116.
­glutamate at residue 600 (V600E)87,88, predicting for sensitivity to the Cytotoxic therapy could synergize with ICBs by killing tumour cells,
BRAF inhibitors vemurafenib and dabrafenib as single agents89,90, or improving the T-cell-to-tumour ratio and restoring the ­metabolic restric-
dabrafenib in combination with the MEK inhibitor trametinib91. tions that result in T cell hyporesponsiveness in cancer117,118. Cytotoxic
Somatic mutations that affect MET exon 14, which contains the Y1003 ­therapy could also reduce immunosuppressive factors released by
residue required for the recruitment of CBL ubiquitin ligase that targets tumours or promote the release of antigens for presentation, b­ roadening
MET for ubiquitin-mediated degradation, lead to increased MET stability the antitumour T cell response. The combination of pembrolizumab, car-
and prolonged signalling from hepatocyte growth factor stimulation92. boplatin and pemetrexed resulted in improved ORRs and PFS ­compared
Patients with NSCLC harbouring MET exon 14 skipping may respond to to cytotoxic therapy alone, and this combination could be an effective
MET inhibitors including crizotinib or cabozantinib93,94. and tolerable first-line treatment option for patients with advanced non-
Other potential targetable gene alterations include mutations in HER2, LUSC119. Several randomized studies are evaluating the role of ICBs in
rearrangements in the proto-oncogene RET, which encodes a receptor combination with platinum-based cytotoxic regimens, with or without
tyrosine kinase, and fusions of the neurotrophic tyrosine receptor kinase bevacizumab, in first-line advanced or recurrent NSCLC. The magnitude
(NTRK) genes 1, 2 and 3, which code for tropomyosin receptor kinases of the overall survival benefit of such combinations will determine their
(TRK) A, B and C, respectively. Initial results from targeted treatment future use.
against HER2 and RET alterations have shown modest activity compared Combining anti-PD-(L)1 and anti-CTLA-4 monoclonal antibodies
to other oncogenic targets, which may reflect their roles as dominant may result in higher and more durable responses in NSCLC, as observed
clonal drivers95–99. By contrast, selective TRK TKIs have demonstrated in experimental models120, and suggested in a single-arm clinical study121.
histology-agnostic efficacy in patients with NTRK fusion-positive Randomized studies that combine nivolumab with ipilimumab, and
cancers, which occur in less than 1% of NSCLC. Although responses durvalumab with the anti-CTLA-4 human IgG4 antibody tremelimumab,
to TRK inhibition are notable and durable100, the duration of response are continuing and overall survival data from these studies are eagerly
may eventually be limited by resistance, and a next-generation TKI has awaited. It is expected that the safety profile for such combinations may
been identified to overcome acquired resistance to previous TRK kinase not be as favourable as those with ICB plus cytotoxic therapy combina-
inhibition101. tions, especially in terms of immune-related adverse events. Nevertheless,

4 5 0 | N A T U R E | V O L 5 5 3 | 2 5 j anuar y 2 0 1 8
© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
Review insight

0 6 12 18 24 30 36

EGFR 3rd gen TKI TKI/chemo/ICB


Low

15–17%
EGFR 1st gen T790M 3rd gen TKI TKI/chemo/ICB
Oncogene addiction
(incidence: 25%)

ALK 2nd gen TKI TKI/chemo/ICB


3–4%
ALK 1st gen TKI ALK 2nd gen TKI TKI/chemo/ICB
2–3% MET exon 14 TKI TKI/chemo/ICB

1% ROS1 1st gen TKI TKI/chemo/ICB


1% BRAF/MEK TKI Chemo/ICB

<1% NTRK TKI TKI/chemo/ICB


PD-L1 expression

Chemo PD-(L)1
(incidence: 50%)

Chemo PD-(L)1 Chemo/IT

Chemo + PD-(L)1 Chemo/IT


Other

PD-(L)1 + CTLA-4 Chemo/IT

PD-(L)1 + IDO1 Chemo/IT


KEY
CHEMO + PD-(L)1 + CTLA-4 Chemo/IT Indicates PFS range estimated
from ongoing study
(incidence: 25–30%)

Experimental or
MSI high PD-1 Chemo/IT
PD-L1 high+++

clinical trial option


PD-1 Chemo/IT Biopsy indicated or recommended
PD-(L)1 + chemo Approved treatment option
Investigational treatment
High

PD-(L)1 + CTLA-4 (ongoing phase II or III clinical study)

0 6 12 18 24 30 36
PFS (months)
Figure 2 | Current and investigative treatment options for advanced is not the best indicator to capture the overall true benefit of ICBs, as a
or metastatic NSCLC. Illustration of the current and future personalized proportion of patients remain alive or disease-free even after long-term
treatment options for NSCLC. Targetable oncogenic drivers account for follow-up. In patients with tumours with high (>​50%) or low (>​1%)
approximately 25% of NSCLC, of which EGFR mutations are the most expression levels of PD-L1, current studies are assessing the benefit of
frequent45. Biopsies are indicated at the time of disease progression anti-PD-(L)1 combinations with cytotoxic therapy, anti-CTLA-4, or other
to determine the best treatment option. For patients with tumours immunotherapy (IT) approaches. PFS estimates illustrated for targeted
expressing high levels of PD-L1 (>​50%) or high levels of microsatellite therapies from refs 52, 54, 64, 73, 79, 83, 100 and for ICBs from refs 114,
instability (MSI), single agent ICB is indicated114. In general, median PFS 119, 121.

despite encouraging results with prolonged benefit in selected patients, for NSCLC, include the use of autologous vaccines that use genomic
most lung tumours are either inherently resistant or will adapt to or information from a specific tumour to predict neo-epitopes that could be
become resistant to current immunotherapies. The challenge is to develop presented to T cells, for the design and manufacture of a vaccine unique
rational combinations that will increase responses, or delay the onset of for each patient127.
resistance122. Most patients who achieve an initial benefit from an ICB eventually
Whereas ICB monotherapy may be appropriate for tumours with high develop resistance. Some of the mechanisms for acquired resistance to
expression levels of PD-L1 or a high nonsynonymous TMB, a d ­ ifferent ICBs include defects in interferon-γ​signalling or major histocompatibility
approach may be required for tumours with fewer T cells and a lower complex presentation128, and increased levels of the enzyme indoleamine
TMB (Supplementary Table 1). The induction of immunogenic ­cancer 2,3-dioxygenase (IDO1), which catabolizes tryptophan, an amino acid
death with the use of cytotoxic therapy, epigenetic modifiers123 or required for optimal T cell function129,130.
­oncolytic viruses124 appears promising in preclinical models or in early ICBs are poised to move to earlier stages of lung cancer therapy, in
phase 1 c­ linical studies. Another strategy includes combination with anti-­ an attempt to improve survival after surgery or radiotherapy, in which
angiogenic drugs, as VEGF contributes to an immunosuppressive TME by the goal is curative. A randomized trial of durvalumab as a sequential
recruiting suppressive immune cells, such as myeloid-derived suppressor treatment in patients with locally advanced unresectable stage NSCLC
cells and regulatory T cells125. Furthermore, angiogenic ­inhibitors may who had not progressed after standard concurrent platinum-based cyto-
increase immune cell infiltration126. toxic and radiation therapy showed a notable improvement in median
In tumours with a low TMB, few T cells and low PD-L1 expression PFS compared to placebo131. The role of ICBs in patients with curable
(‘cold tumours’), the challenge for immunotherapeutic approaches is not non-metastatic disease will be further clarified with the results of many
only to attract effector T cells into the TME, but also to present tumour continuing trials accruing in both the perioperative setting and in patients
antigens to T cells. Possible approaches for such cold NSCLCs could with locally advanced disease treated with concurrent cytotoxic and
include the use of adoptive transfer of autologous tumour-infiltrating ­radiation therapy.
lymphocytes or chimaeric antigen receptor T cell therapy. For the latter, The notable clinical success of cancer immunotherapy over a short
NSCLC-specific or unique cell-surface antigens will need to be identified. period of time suggests that it may form the foundation of future ­curative-
Other approaches being developed for solid tumours, and in the future intent regimens for many malignancies, including NSCLC (Fig. 2).

2 5 j anuar y 2 0 1 8 | V O L 5 5 3 | N A T U R E | 4 5 1
© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
insight Review

Advancing personalized medicine and trial design information with the cellular components of the TME could allow the
The rapid development of targeted therapies with newer and more use of rationally designed combination therapies138.
potent generations of drugs, and the characterization of the mecha- Over the past 20 years, there has been enormous progress in the
nisms of acquired resistance, established the role for repeated genomic understanding of the biology and management of NSCLC, with targeted
­profiling at the time of tumour progression, particularly in patients with and immunotherapies providing a new foundation towards rationally
EGFR mutations, in which the most common cause of resistance, the designed therapeutic regimens with manageable toxicity profiles and
T790M mutation, may be successfully treated with osimertinib. The improvement in survival. Although the treatment of patients with met-
repeated genomic profile has traditionally been performed through astatic NSCLC has long been considered palliative, continuing drug
repeated biopsy, which may not always be feasible and carries risks for development provides the hope for prolonged survival in an increasing
complications. An ­emerging option is the use of plasma genotyping with number of patients and, for the first time, raises the possibility of a cure
sequencing circulating tumour DNA (ctDNA). In a retrospective analysis in those with metastatic disease. Furthermore, the use of new therapeutic
including 216 patients with both central tissue and plasma genotyping modalities including immunotherapy may have an even higher impact in
available before treatment with osimertinib in a clinical trial, the sensi- patients with earlier non-detectable metastatic disease, in whom treatment
tivity of plasma genotyping for EGFR(T790M) was 70%132. Furthermore, is given with curative intent.
plasma EGFR(T790M) mutations were found in 31% of patients with
tissue-negative results. Because the outcomes with osimertinib t­ reatment, received 18 August; accepted 29 November 2017.
including ORRs and PFS, were similar for T790M mutations detected
by tissue or plasma, a repeated biopsy could be avoided in those with 1. Torre, L. A. et al. Global cancer statistics, 2012. CA Cancer J. Clin. 65, 87–108
positive plasma results. By contrast, patients with negative plasma results (2015).
2. Molina, J. R., Yang, P., Cassivi, S. D., Schild, S. E. & Adjei, A. A. Non-small cell
should undergo a repeated biopsy. Multianalytical ctDNA tests for lung cancer: epidemiology, risk factors, treatment, and survivorship. Mayo Clin.
multiple molecular markers, including the detection of high TMB, are Proc. 83, 584–594 (2008).
being developed to support individualized lung cancer treatment. 3. Alberg, A. J., Brock, M. V., Ford, J. G., Samet, J. M. & Spivack, S. D. Epidemiology
of lung cancer: diagnosis and management of lung cancer, 3rd ed.: American
Because many gene alterations are uncommon or rare, and h ­ undreds College of Chest Physicians evidence-based clinical practice guidelines. Chest
of combination studies are being conducted in NSCLC, recruiting 143, e1S–e29S (2013).
patients to clinical trials is becoming more difficult. To adjust to this 4. Sun, S., Schiller, J. H. & Gazdar, A. F. Lung cancer in never smokers--a different
disease. Nat. Rev. Cancer 7, 778–790 (2007).
current ­situation, innovative master protocols are being used in which 5. Vineis, P. et al. Environmental tobacco smoke and risk of respiratory cancer
several questions can be answered in one study. Such master protocols are and chronic obstructive pulmonary disease in former smokers and never
designed to encompass a collection of trials that share key design features smokers in the EPIC prospective study. Br. Med. J. 330, 277 (2005).
and used to assess either multiple targeted therapies or immunotherapy 6. Hackshaw, A. K., Law, M. R. & Wald, N. J. The accumulated evidence on lung
cancer and environmental tobacco smoke. Br. Med. J. 315, 980–988 (1997).
combinations for a single disease (umbrella trials) or a single targeted 7. Shahab, L. et al. Nicotine, carcinogen, and toxin exposure in long-term
therapy for multiple diseases (basket trials). The Biomarker-integrated e-cigarette and nicotine replacement therapy users: a cross-sectional study.
Approaches of Targeted Therapy for Lung cancer Elimination (BATTLE) Ann. Intern. Med. 166, 390–400 (2017).
8. Hays, J. T. & Ebbert, J. O. Varenicline for tobacco dependence. N. Engl. J. Med.
and Lung Cancer Master Protocol (Lung-MAP) trials are among the first 359, 2018–2024 (2008).
umbrella trials for patients with NSCLC, with the former showing the 9. Brandon, T. H. et al. Electronic nicotine delivery systems: a policy statement
feasibility of performing fresh biopsies to guide the next line of therapy, a from the American Association for Cancer Research and the American Society
of Clinical Oncology. J. Clin. Oncol. 33, 952–963 (2015).
principle that has been often used in the most recent studies133,134. 10. Zhang, J. et al. Intratumor heterogeneity in localized lung adenocarcinomas
delineated by multiregion sequencing. Science 346, 256–259 (2014).
Future perspectives 11. Engelman, J. A. et al. MET amplification leads to gefitinib resistance in lung
cancer by activating ERBB3 signaling. Science 316, 1039–1043 (2007).
The treatment of NSCLC has undergone remarkable changes. A better 12. Turke, A. B. et al. Preexistence and clonal selection of MET amplification in
understanding of the tumour biology enabled the development of EGFR mutant NSCLC. Cancer Cell 17, 77–88 (2010).
targeted therapies that heralded the era of personalized medicine. 13. Pao, W. et al. KRAS mutations and primary resistance of lung
adenocarcinomas to gefitinib or erlotinib. PLoS Med. 2, e17 (2005).
Furthermore, the introduction of ICBs has led to prolonged survival in 14. Ding, L. et al. Somatic mutations affect key pathways in lung adenocarcinoma.
selected patients. Nevertheless, this unprecedented benefit from current Nature 455, 1069–1075 (2008).
standard therapies is still observed in only a minority of patients, with 15. Reck, M. & Rabe, K. F. Precision diagnosis and treatment for advanced
targeted therapies restricted to non-LUSC histological subtypes that contain non-small-cell lung cancer. N. Engl. J. Med. 377, 849–861 (2017).
16. Ahrendt, S. A. et al. p53 mutations and survival in stage I non-small-cell lung
actionable driver mutations, and durable responses from immunotherapy cancer: results of a prospective study. J. Natl. Cancer Inst. 95, 961–970 (2003).
occurring uncommonly. One of the main concerns with targeted therapy 17. Cancer Genome Atlas Research Network. Comprehensive molecular profiling
is the emergence of secondary clones that may not be effectively of lung adenocarcinoma. Nature 511, 543–550 (2014).
18. Robinson, D. R. et al. Integrative clinical genomics of metastatic cancer. Nature
targeted by the initial treatment directed at the founder clone. To improve 548, 297–303 (2017).
outcomes further, there is a need to understand better the mechanisms of 19. McGranahan, N. et al. Clonal neoantigens elicit T cell immunoreactivity and
acquired resistance to allow their prevention or effective treatment at the sensitivity to immune checkpoint blockade. Science 351, 1463–1469 (2016).
20. Le, D. T. et al. PD-1 blockade in tumors with mismatch-repair deficiency.
time of emergence. Therefore, focusing on both dominant and sub-clones N. Engl. J. Med. 372, 2509–2520 (2015).
may be required for a more effective and durable benefit from targeted 21. Koyama, S. et al. STK11/LKB1 deficiency promotes neutrophil recruitment
therapies135. and proinflammatory cytokine production to suppress T-cell activity in the
lung tumor microenvironment. Cancer Res. 76, 999–1008 (2016).
The application of ctDNA to track the evolutionary dynamics of 22. Cancer Genome Atlas Research Network. Comprehensive genomic
early stage lung cancer should be expanded to detect both oncogenic characterization of squamous cell lung cancers. Nature 489, 519–525 (2012).
drivers and track resistant mutations, providing a future approach for References 17 and 22 are landmark genomics analyses that describe the
molecular landscapes of lung adenocarcinoma and squamous cell
ctDNA-driven targeted therapeutics136. A systematic approach to collect carcinoma, respectively.
tissue samples not only at the time of diagnosis but also serially at the 23. Hanna, N. et al. Systemic therapy for stage IV non-small-cell lung cancer:
times of relapse to evaluate the dynamic clonal evolution that occurs over american society of clinical oncology clinical practice guideline update. J. Clin.
time and possibly at different metastatic sites will be crucial for further Oncol. 35, 3484–3515 (2017).
24. Schiller, J. H. et al. Comparison of four chemotherapy regimens for advanced
therapeutic advances. non-small-cell lung cancer. N. Engl. J. Med. 346, 92–98 (2002).
Better predictors for response to immunotherapy are critical for its 25. Scagliotti, G. V. et al. Phase III study comparing cisplatin plus gemcitabine with
­optimal use. Although both PD-L1 and TMB may be used to select cisplatin plus pemetrexed in chemotherapy-naive patients with advanced-
stage non-small-cell lung cancer. J. Clin. Oncol. 26, 3543–3551 (2008).
patients for therapy137, most patients will not fit the ideal profile based 26. Sandler, A. et al. Paclitaxel-carboplatin alone or with bevacizumab for
on these two biomarkers. Furthermore, correlating cancer genomic non-small-cell lung cancer. N. Engl. J. Med. 355, 2542–2550 (2006).

4 5 2 | N A T U R E | V O L 5 5 3 | 2 5 j anuar y 2 0 1 8
© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
Review insight

27. Thatcher, N. et al. Necitumumab plus gemcitabine and cisplatin versus 54. Rosell, R. et al. Erlotinib versus standard chemotherapy as first-line treatment
gemcitabine and cisplatin alone as first-line therapy in patients with stage IV for European patients with advanced EGFR mutation-positive non-small-cell
squamous non-small-cell lung cancer (SQUIRE): an open-label, randomised, lung cancer (EURTAC): a multicentre, open-label, randomised phase 3 trial.
controlled phase 3 trial. Lancet Oncol. 16, 763–774 (2015). Lancet Oncol. 13, 239–246 (2012).
28. Garon, E. B. et al. Ramucirumab plus docetaxel versus placebo plus docetaxel 55. Paz-Ares, L. et al. Afatinib versus gefitinib in patients with EGFR mutation-
for second-line treatment of stage IV non-small-cell lung cancer after disease positive advanced non-small-cell lung cancer: overall survival data from the
progression on platinum-based therapy (REVEL): a multicentre, double-blind, phase IIb LUX-Lung 7 trial. Ann. Oncol. 28, 270–277 (2017).
randomised phase 3 trial. Lancet 384, 665–673 (2014). 56. Wu, Y. L. et al. Dacomitinib versus gefitinib as first-line treatment for patients
29. Hirsch, F. R. et al. Lung cancer: current therapies and new targeted treatments. with EGFR-mutation-positive non-small-cell lung cancer (ARCHER 1050): a
Lancet 389, 299–311 (2017). randomised, open-label, phase 3 trial. Lancet Oncol. 18, 1454–1466 (2017).
30. Goldstraw, P. et al. The IASLC Lung Cancer Staging Project: Proposals for 57. Yang, J. C. et al. Afatinib versus cisplatin-based chemotherapy for EGFR
Revision of the TNM Stage Groupings in the Forthcoming (Eighth) Edition of mutation-positive lung adenocarcinoma (LUX-Lung 3 and LUX-Lung 6):
the TNM Classification for Lung Cancer. J. Thorac. Oncol. 11, 39–51 (2016). analysis of overall survival data from two randomised, phase 3 trials.
31. Kris, M. G. et al. Adjuvant systemic therapy and adjuvant radiation therapy for Lancet Oncol. 16, 141–151 (2015).
stage I to IIIA completely resected non-small-cell lung cancers: American 58. Okabe, T. et al. Differential constitutive activation of the epidermal growth
Society of Clinical Oncology/Cancer Care Ontario clinical practice guideline factor receptor in non-small cell lung cancer cells bearing EGFR gene
update. J. Clin. Oncol. 35, 2960–2974 (2017). mutation and amplification. Cancer Res. 67, 2046–2053 (2007).
32. Liang, H. et al. Robotic versus video-assisted lobectomy/segmentectomy for 59. Kobayashi, S. et al. EGFR mutation and resistance of non-small-cell lung
lung cancer: a meta-analysis. Ann. Surg. https://doi.org/10.1097/ cancer to gefitinib. N. Engl. J. Med. 352, 786–792 (2005).
SLA.0000000000002346 (2017). 60. Sequist, L. V. et al. Genotypic and histological evolution of lung cancers
33. Chi, A., Chen, H., Wen, S., Yan, H. & Liao, Z. Comparison of particle beam acquiring resistance to EGFR inhibitors. Sci. Transl. Med. 3, 75ra26 (2011).
therapy and stereotactic body radiotherapy for early stage non-small cell lung 61. Camidge, D. R., Pao, W. & Sequist, L. V. Acquired resistance to TKIs in solid
cancer: A systematic review and hypothesis-generating meta-analysis. tumours: learning from lung cancer. Nat. Rev. Clin. Oncol. 11, 473–481
Radiother. Oncol. 123, 346–354 (2017). (2014).
34. Dillman, R. O. et al. A randomized trial of induction chemotherapy plus 62. Jänne, P. A. et al. AZD9291 in EGFR inhibitor-resistant non-small-cell lung
high-dose radiation versus radiation alone in stage III non-small-cell lung cancer. N. Engl. J. Med. 372, 1689–1699 (2015).
cancer. N. Engl. J. Med. 323, 940–945 (1990). 63. Mok, T. S. et al. Osimertinib or platinum-pemetrexed in EGFR T790M-positive
35. Curran, W. J., Jr et al. Sequential vs. concurrent chemoradiation for stage III lung cancer. N. Engl. J. Med. 376, 629–640 (2017).
non-small cell lung cancer: randomized phase III trial RTOG 9410. J. Natl. 64. Soria, J. C. et al. Osimertinib in untreated EGFR-mutated advanced non-small
Cancer Inst. 103, 1452–1460 (2011). cell lung cancer. N. Engl. J. Med. http://doi.org/10.1056/NEJMoa1713137
36. Herbst, R. S., Heymach, J. V. & Lippman, S. M. Lung cancer. N. Engl. J. Med. (2017).
359, 1367–1380 (2008). 65. Thress, K. S. et al. Acquired EGFR C797S mutation mediates resistance to
37. Morgensztern, D., Ng, S. H., Gao, F. & Govindan, R. Trends in stage distribution AZD9291 in non-small cell lung cancer harboring EGFR T790M. Nat. Med. 21,
for patients with non-small cell lung cancer: a National Cancer Database 560–562 (2015).
survey. J. Thorac. Oncol. 5, 29–33 (2010). 66. Niederst, M. J. et al. The allelic context of the C797S mutation acquired upon
38. Fukuoka, M. et al. Multi-institutional randomized phase II trial of gefitinib for treatment with third-generation EGFR inhibitors impacts sensitivity to
previously treated patients with advanced non-small-cell lung cancer subsequent treatment strategies. Clin. Cancer Res. 21, 3924–3933 (2015).
(The IDEAL 1 Trial) [corrected]. J. Clin. Oncol. 21, 2237–2246 (2003). 67. Jia, Y. et al. Overcoming EGFR(T790M) and EGFR(C797S) resistance with
39. Kris, M. G. et al. Efficacy of gefitinib, an inhibitor of the epidermal growth factor mutant-selective allosteric inhibitors. Nature 534, 129–132 (2016).
receptor tyrosine kinase, in symptomatic patients with non-small cell lung 68. Uchibori, K. et al. Brigatinib combined with anti-EGFR antibody overcomes
cancer: a randomized trial. J. Am. Med. Assoc. 290, 2149–2158 (2003). osimertinib resistance in EGFR-mutated non-small-cell lung cancer.
40. Shepherd, F. A. et al. Erlotinib in previously treated non-small-cell lung cancer. Nat. Commun. 8, 14768 (2017).
N. Engl. J. Med. 353, 123–132 (2005). 69. Lin, J. J., Riely, G. J. & Shaw, A. T. Targeting ALK: precision medicine takes on
41. Lynch, T. J. et al. Activating mutations in the epidermal growth factor receptor drug resistance. Cancer Discov. 7, 137–155 (2017).
underlying responsiveness of non-small-cell lung cancer to gefitinib. N. Engl. 70. Soda, M. et al. Identification of the transforming EML4-ALK fusion gene in
J. Med. 350, 2129–2139 (2004). non-small-cell lung cancer. Nature 448, 561–566 (2007).
42. Paez, J. G. et al. EGFR mutations in lung cancer: correlation with clinical This study describes the discovery of ALK rearrangements in NSCLC.
response to gefitinib therapy. Science 304, 1497–1500 (2004). 71. Kwak, E. L. et al. Anaplastic lymphoma kinase inhibition in non-small-cell lung
References 41 and 42 were among the first studies to demonstrate that cancer. N. Engl. J. Med. 363, 1693–1703 (2010).
EGFR mutations in NSCLC confer sensitivity to anti-EGFR tyrosine kinase This study is the first to report the activity of crizotinib in patients with ALK
inhibitors. rearrangements.
43. Rikova, K. et al. Global survey of phosphotyrosine signaling identifies 72. Shaw, A. T. et al. Crizotinib versus chemotherapy in advanced ALK-positive
oncogenic kinases in lung cancer. Cell 131, 1190–1203 (2007). lung cancer. N. Engl. J. Med. 368, 2385–2394 (2013).
44. Kazandjian, D. et al. FDA approval summary: nivolumab for the treatment of 73. Solomon, B. J. et al. First-line crizotinib versus chemotherapy in ALK-positive
metastatic non-small cell lung cancer with progression on or after platinum- lung cancer. N. Engl. J. Med. 371, 2167–2177 (2014).
based chemotherapy. Oncologist 21, 634–642 (2016). 74. Shaw, A. T. et al. Ceritinib in ALK-rearranged non-small-cell lung cancer.
45. Kris, M. G. et al. Using multiplexed assays of oncogenic drivers in lung cancers N. Engl. J. Med. 370, 1189–1197 (2014).
to select targeted drugs. J. Am. Med. Assoc. 311, 1998–2006 (2014). 75. Shaw, A. T. et al. Alectinib in ALK-positive, crizotinib-resistant, non-small-cell
46. Lemmon, M. A., Schlessinger, J. & Ferguson, K. M. The EGFR family: not so lung cancer: a single-group, multicentre, phase 2 trial. Lancet Oncol. 17,
prototypical receptor tyrosine kinases. Cold Spring Harb. Perspect. Biol. 6, 234–242 (2016).
a020768 (2014). 76. Kim, D. W. et al. Brigatinib in patients with crizotinib-refractory anaplastic
47. Wheeler, D. L., Dunn, E. F. & Harari, P. M. Understanding resistance to EGFR lymphoma kinase-positive non-small-cell lung cancer: a randomized,
inhibitors-impact on future treatment strategies. Nat. Rev. Clin. Oncol. 7, multicenter phase II trial. J. Clin. Oncol. 35, 2490–2498 (2017).
493–507 (2010). 77. Soria, J. C. et al. First-line ceritinib versus platinum-based chemotherapy in
48. Sharma, S. V., Bell, D. W., Settleman, J. & Haber, D. A. Epidermal growth factor advanced ALK-rearranged non-small-cell lung cancer (ASCEND-4): a
receptor mutations in lung cancer. Nat. Rev. Cancer 7, 169–181 (2007). randomised, open-label, phase 3 study. Lancet 389, 917–929 (2017).
49. Mok, T. S. et al. Gefitinib or carboplatin-paclitaxel in pulmonary 78. Hida, T. et al. Alectinib versus crizotinib in patients with ALK-positive
adenocarcinoma. N. Engl. J. Med. 361, 947–957 (2009). non-small-cell lung cancer (J-ALEX): an open-label, randomised phase 3 trial.
This study relates to a change in the era of personalized therapy, and Lancet 390, 29–39 (2017).
demonstrates that an anti-EGFR tyrosine kinase inhibitor is superior to 79. Peters, S. et al. Alectinib versus crizotinib in untreated ALK-positive
cytotoxic therapy in patients with tumours that contain an activating EGFR non-small-cell lung cancer. N. Engl. J. Med. 377, 829–838 (2017).
mutation. 80. Gainor, J. F. et al. Molecular mechanisms of resistance to first- and second-
50. Fukuoka, M. et al. Biomarker analyses and final overall survival results from a generation ALK inhibitors in ALK-rearranged lung cancer. Cancer Discov. 6,
phase III, randomized, open-label, first-line study of gefitinib versus 1118–1133 (2016).
carboplatin/paclitaxel in clinically selected patients with advanced non-small- 81. Shaw, A. T. et al. Lorlatinib in non-small-cell lung cancer with ALK or ROS1
cell lung cancer in Asia (IPASS). J. Clin. Oncol. 29, 2866–2874 (2011). rearrangement: an international, multicentre, open-label, single-arm
51. Maemondo, M. et al. Gefitinib or chemotherapy for non-small-cell lung cancer first-in-man phase 1 trial. Lancet Oncol. 18, 1590–1599 (2017).
with mutated EGFR. N. Engl. J. Med. 362, 2380–2388 (2010). 82. Facchinetti, F. et al. Oncogene addiction in non-small cell lung cancer: focus on
52. Mitsudomi, T. et al. Gefitinib versus cisplatin plus docetaxel in patients with ROS1 inhibition. Cancer Treat. Rev. 55, 83–95 (2017).
non-small-cell lung cancer harbouring mutations of the epidermal growth 83. Shaw, A. T. et al. Crizotinib in ROS1-rearranged non-small-cell lung cancer.
factor receptor (WJTOG3405): an open label, randomised phase 3 trial. N. Engl. J. Med. 371, 1963–1971 (2014).
Lancet Oncol. 11, 121–128 (2010). 84. Lim, S. M. et al. Open-label, multicenter, phase II study of ceritinib in patients
53. Zhou, C. et al. Erlotinib versus chemotherapy as first-line treatment for with non-small-cell lung cancer harboring Ros1 rearrangement. J. Clin. Oncol.
patients with advanced EGFR mutation-positive non-small-cell lung cancer 35, 2613–2618 (2017).
(OPTIMAL, CTONG-0802): a multicentre, open-label, randomised, phase 3 85. Awad, M. M. et al. Acquired resistance to crizotinib from a mutation in
study. Lancet Oncol. 12, 735–742 (2011). CD74-ROS1. N. Engl. J. Med. 368, 2395–2401 (2013).

2 5 j anuar y 2 0 1 8 | V O L 5 5 3 | N A T U R E | 4 5 3
© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
insight Review

86. Davies, K. D. et al. Resistance to ROS1 inhibition mediated by EGFR pathway 116. Herbst, R. S. & Sznol, M. Diminished but not dead: chemotherapy for the
activation in non-small cell lung cancer. PLoS One 8, e82236 (2013). treatment of NSCLC. Lancet Oncol. 17, 1464–1465 (2016).
87. Marchetti, A. et al. Clinical features and outcome of patients with non-small-cell 117. Chang, C. H. et al. Metabolic competition in the tumor microenvironment is a
lung cancer harboring BRAF mutations. J. Clin. Oncol. 29, 3574–3579 (2011). driver of cancer progression. Cell 162, 1229–1241 (2015).
88. Cardarella, S. et al. Clinical, pathologic, and biologic features associated with 118. Zitvogel, L., Galluzzi, L., Smyth, M. J. & Kroemer, G. Mechanism of action of
BRAF mutations in non-small cell lung cancer. Clin. Can. Res. 19, 4532–4540 conventional and targeted anticancer therapies: reinstating
(2013). immunosurveillance. Immunity 39, 74–88 (2013).
89. Hyman, D. M. et al. Vemurafenib in multiple nonmelanoma cancers with BRAF 119. Langer, C. J. et al. Carboplatin and pemetrexed with or without pembrolizumab
V600 mutations. N. Engl. J. Med. 373, 726–736 (2015). for advanced, non-squamous non-small-cell lung cancer: a randomised,
90. Planchard, D. et al. Dabrafenib in patients with BRAF(V600E)-positive phase 2 cohort of the open-label KEYNOTE-021 study. Lancet Oncol. 17,
advanced non-small-cell lung cancer: a single-arm, multicentre, open-label, 1497–1508 (2016).
phase 2 trial. Lancet Oncol. 17, 642–650 (2016). 120. Curran, M. A., Montalvo, W., Yagita, H. & Allison, J. P. PD-1 and CTLA-4
91. Planchard, D. et al. Dabrafenib plus trametinib in patients with previously combination blockade expands infiltrating T cells and reduces regulatory
treated BRAF(V600E)-mutant metastatic non-small cell lung cancer: an T and myeloid cells within B16 melanoma tumors. Proc. Natl Acad. Sci. USA
open-label, multicentre phase 2 trial. Lancet Oncol. 17, 984–993 (2016). 107, 4275–4280 (2010).
92. Frampton, G. M. et al. Activation of MET via diverse exon 14 splicing alterations 121. Hellmann, M. D. et al. Nivolumab plus ipilimumab as first-line treatment for
occurs in multiple tumor types and confers clinical sensitivity to MET advanced non-small-cell lung cancer (CheckMate 012): results of an
inhibitors. Cancer Discov. 5, 850–859 (2015). open-label, phase 1, multicohort study. Lancet Oncol. 18, 31–41 (2017).
93. Paik, P. K. et al. Response to MET inhibitors in patients with stage IV lung 122. Sharma, P., Hu-Lieskovan, S., Wargo, J. A. & Ribas, A. Primary, adaptive, and
adenocarcinomas harboring MET mutations causing exon 14 skipping. acquired resistance to cancer immunotherapy. Nat. Med. 23, 1362–1368
Cancer Discov. 5, 842–849 (2015). (2017).
94. Awad, M. M. et al. MET exon 14 mutations in non-small-cell lung cancer are 123. Bezu, L. et al. Combinatorial strategies for the induction of immunogenic cell
associated with advanced age and stage-dependent MET genomic death. Front. Immunol. 6, 187 (2015).
amplification and c-Met overexpression. J. Clin. Oncol. 34, 721–730 (2016). 124. Lawler, S. E., Speranza, M. C., Cho, C. F. & Chiocca, E. A. Oncolytic viruses in
95. Mazières, J. et al. Lung cancer that harbors an HER2 mutation: epidemiologic cancer treatment: a review. JAMA Oncol. 3, 841–849 (2017).
characteristics and therapeutic perspectives. J. Clin. Oncol. 31, 1997–2003 125. Voron, T. et al. Control of the immune response by pro-angiogenic factors.
(2013). Front. Oncol. 4, 70 (2014).
96. Mazières, J. et al. Lung cancer patients with HER2 mutations treated with 126. Tian, L. et al. Mutual regulation of tumour vessel normalization and
chemotherapy and HER2-targeted drugs: results from the European EUHER2 immunostimulatory reprogramming. Nature 544, 250–254 (2017).
cohort. Ann. Oncol. 27, 281–286 (2016). 127. Sahin, U. et al. Personalized RNA mutanome vaccines mobilize poly-specific
97. Kohno, T. et al. KIF5B-RET fusions in lung adenocarcinoma. Nat. Med. 18, therapeutic immunity against cancer. Nature 547, 222–226 (2017).
375–377 (2012). 128. Gettinger, S. et al. Impaired HLA class I antigen processing and presentation as
98. Drilon, A. et al. Cabozantinib in patients with advanced RET-rearranged a mechanism of acquired resistance to immune checkpoint inhibitors in lung
non-small-cell lung cancer: an open-label, single-centre, phase 2, single-arm cancer. Cancer Discov. 7, 1420–1435 (2017).
trial. Lancet Oncol. 17, 1653–1660 (2016). 129. Zaretsky, J. M. et al. Mutations associated with acquired resistance to PD-1
99. Gautschi, O. et al. Targeting RET in patients with RET-rearranged lung cancers: blockade in melanoma. N. Engl. J. Med. 375, 819–829 (2016).
results from the global, multicenter RET registry. J. Clin. Oncol. 35, 1403–1410 130. Holmgaard, R. B., Zamarin, D., Munn, D. H., Wolchok, J. D. & Allison, J. P.
(2017). Indoleamine 2,3-dioxygenase is a critical resistance mechanism in antitumor
100. Hyman, D. M. et al. The efficacy of larotrectinib (LOXO-101), a selective T cell immunotherapy targeting CTLA-4. J. Exp. Med. 210, 1389–1402
tropomyosin receptor kinase (TRK) inhibitor, in adult and pediatric TRK fusion (2013).
cancers. J. Clin. Oncol. 35, LBA2501–LBA2501 (2017). 131. Antonia, S. J. et al. Durvalumab after chemoradiotherapy in stage III non-small-
101. Drilon, A. et al. A next-generation TRK kinase inhibitor overcomes acquired cell lung cancer. N. Engl. J. Med. 377, 1919–1929 (2017).
resistance to prior TRK kinase inhibition in patients with TRK fusion-positive 132. Oxnard, G. R. et al. Association between plasma genotyping and outcomes of
solid tumors. Cancer Discov. 7, 963–972 (2017). treatment with osimertinib (AZD9291) in advanced non-small-cell lung
102. Coley, W. B. The treatment of malignant tumors by repeated inoculations of cancer. J. Clin. Oncol. 34, 3375–3382 (2016).
erysipelas. With a report of ten original cases. 1893. Clin. Orthop. Relat. Res. 133. Kim, E. S. et al. The BATTLE trial: personalizing therapy for lung cancer.
(262):3–11 (1991). Cancer Discov. 1, 44–53 (2011).
103. Leach, D. R., Krummel, M. F. & Allison, J. P. Enhancement of antitumor 134. Herbst, R. S. et al. Lung Master Protocol (Lung-MAP)-A biomarker-driven
immunity by CTLA-4 blockade. Science 271, 1734–1736 (1996). protocol for accelerating development of therapies for squamous cell lung
104. Hodi, F. S. et al. Improved survival with ipilimumab in patients with metastatic cancer: SWOG S1400. Clin. Cancer Res. 21, 1514–1524 (2015).
melanoma. N. Engl. J. Med. 363, 711–723 (2010). 135. Blakely, C. M. et al. Evolution and clinical impact of co-occurring genetic
105. Dong, H. et al. Tumor-associated B7-H1 promotes T-cell apoptosis: a potential alterations in advanced-stage EGFR-mutant lung cancers. Nat. Genet. 49,
mechanism of immune evasion. Nat. Med. 8, 793–800 (2002). 1693–1704 (2017).
106. Iwai, Y. et al. Involvement of PD-L1 on tumor cells in the escape from host 136. Abbosh, C. et al. Phylogenetic ctDNA analysis depicts early-stage lung cancer
immune system and tumor immunotherapy by PD-L1 blockade. Proc. Natl evolution. Nature 545, 446–451 (2017).
Acad. Sci. USA 99, 12293–12297 (2002). This study introduces ctDNA profiling to track the subclonal nature of lung
107. Topalian, S. L. et al. Safety, activity, and immune correlates of anti-PD-1 cancer progression, providing an approach for ctDNA-driven therapeutic
antibody in cancer. N. Engl. J. Med. 366, 2443–2454 (2012). studies.
108. Garon, E. B. et al. Pembrolizumab for the treatment of non-small-cell lung 137. Rizvi, N. A. et al. Cancer immunology. Mutational landscape determines
cancer. N. Engl. J. Med. 372, 2018–2028 (2015). sensitivity to PD-1 blockade in non-small cell lung cancer. Science 348,
109. Herbst, R. S. et al. Predictive correlates of response to the anti-PD-L1 antibody 124–128 (2015).
MPDL3280A in cancer patients. Nature 515, 563–567 (2014). This is a landmark study indicating that lung cancers with high non-
110. Brahmer, J. et al. Nivolumab versus docetaxel in advanced squamous-cell synonymous mutation burden are more responsive to ICB.
non-small-cell lung cancer. N. Engl. J. Med. 373, 123–135 (2015). 138. Chen, D. S. & Mellman, I. Elements of cancer immunity and the cancer-
111. Borghaei, H. et al. Nivolumab versus docetaxel in advanced nonsquamous immune set point. Nature 541, 321–330 (2017).
non-small-cell lung cancer. N. Engl. J. Med. 373, 1627–1639 (2015). 139. Romero, R. et al. Keap1 loss promotes Kras-driven lung cancer and results in
References 110 and 111 were the first phase 3 studies to show increased dependence on glutaminolysis. Nat. Med. 23, 1362–1368 (2017).
survival for ICBs compared to cytotoxic therapy in patients with previously
treated advanced-stage NSCLC, heralding the era of immunotherapy for Supplementary Information is available in the online version of the paper.
NSCLC.
112. Herbst, R. S. et al. Pembrolizumab versus docetaxel for previously treated, Acknowledgements We would like to thank L. Chen and A. M. Incassati for
PD-L1-positive, advanced non-small-cell lung cancer (KEYNOTE-010): a editorial assistance. R. Herbst is supported by the Yale SPORE in Lung Cancer
randomised controlled trial. Lancet 387, 1540–1550 (2016). (P50CA196530).
113. Rittmeyer, A. et al. Atezolizumab versus docetaxel in patients with previously
treated non-small-cell lung cancer (OAK): a phase 3, open-label, multicentre Author Contributions All authors contributed to the writing of this Review.
randomised controlled trial. Lancet 389, 255–265 (2017).
114. Reck, M. et al. Pembrolizumab versus chemotherapy for PD-L1-positive Author Information Reprints and permissions information is available at
non-small-cell lung cancer. N. Engl. J. Med. 375, 1823–1833 (2016). www.nature.com/reprints. The authors declare competing financial interests:
This study provides evidence that in selected patients with high tumour details are available in the online version of the paper. Readers are welcome to
expression of PD-L1, ICBs are more effective than cytotoxic therapy in the comment on the online version of the paper. Publisher’s note: Springer Nature
first-line setting. remains neutral with regard to jurisdictional claims in published maps and
115. Carbone, D. P. et al. First-line nivolumab in stage IV or recurrent non-small-cell institutional affiliations. Correspondence and requests for materials should be
lung cancer. N. Engl. J. Med. 376, 2415–2426 (2017). addressed to R.S.H. (roy.herbst@yale.edu) or C.B. (chris.boshoff@pfizer.com).

4 5 4 | N A T U R E | V O L 5 5 3 | 2 5 j anuar y 2 0 1 8
© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.

You might also like