You are on page 1of 19

Articles

https://doi.org/10.1038/s41588-018-0209-6

Pharmacogenomic landscape of patient-derived


tumor cells informs precision oncology therapy
Jin-Ku Lee   1,2,3,23, Zhaoqi Liu4,5,23, Jason K. Sa1,3,23, Sang Shin1,6,23, Jiguang Wang   7,8,9,23,
Mykola Bordyuh4,5, Hee Jin Cho1,3, Oliver Elliott4,5, Timothy Chu4,5, Seung Won Choi   1,6,
Daniel I. S. Rosenbloom   4,5, In-Hee Lee   1,3, Yong Jae Shin1,2,3, Hyun Ju Kang1,3, Donggeon Kim1,3,
Sun Young Kim10, Moon-Hee Sim10, Jusun Kim10, Taehyang Lee10, Yun Jee Seo1,3, Hyemi Shin1,6,
Mijeong Lee1,6, Sung Heon Kim1,2, Yong-Jun Kwon1, Jeong-Woo Oh1,6, Minsuk Song1, Misuk Kim1,3,
Doo-Sik Kong2, Jung Won Choi2, Ho Jun Seol2, Jung-Il Lee2, Seung Tae Kim10, Joon Oh Park6,10,
Kyoung-Mee Kim11, Sang-Yong Song11, Jeong-Won Lee12, Hee-Cheol Kim13, Jeong Eon Lee13,
Min Gew Choi13, Sung Wook Seo14, Young Mog Shim15, Jae Ill Zo15, Byong Chang Jeong16, Yeup Yoon3,6,
Gyu Ha Ryu3, Nayoung K. D. Kim   3,17, Joon Seol Bae3,17, Woong-Yang Park3,6,17, Jeongwu Lee18,
Roel G. W. Verhaak   19, Antonio Iavarone20,21,22, Jeeyun Lee   6,10*, Raul Rabadan4,5*
and Do-Hyun Nam1,2,6*

Outcomes of anticancer therapy vary dramatically among patients due to diverse genetic and molecular backgrounds, high-
lighting extensive intertumoral heterogeneity. The fundamental tenet of precision oncology defines molecular characterization
of tumors to guide optimal patient-tailored therapy. Towards this goal, we have established a compilation of pharmacological
landscapes of 462 patient-derived tumor cells (PDCs) across 14 cancer types, together with genomic and transcriptomic pro-
filing in 385 of these tumors. Compared with the traditional long-term cultured cancer cell line models, PDCs recapitulate the
molecular properties and biology of the diseases more precisely. Here, we provide insights into dynamic pharmacogenomic
associations, including molecular determinants that elicit therapeutic resistance to EGFR inhibitors, and the potential repur-
posing of ibrutinib (currently used in hematological malignancies) for EGFR-specific therapy in gliomas. Lastly, we present
a potential implementation of PDC-derived drug sensitivities for the prediction of clinical response to targeted therapeutics
using retrospective clinical studies.

G
enomic and molecular tumor profiling enables the identifica- dard in vitro cancer cell line models16–23. In particular, large-scale
tion of effective drugs tailored to cancer patients1–9. However, drug-screening systems, using conventional cancer cell lines have
predicting successful anticancer therapy remains extremely provided reference points for gene–drug associations, enabling
challenging10–12, largely due to extensive inter- and intratumoral the discovery of molecular markers that may predict therapeutic
heterogeneity13–15. Recent efforts have established a framework response16,17. However, there are several challenges that hamper
for genetic predictions of anticancer drug responses using stan- broad clinical utility of the current gene–drug association map in

1
Institute for Refractory Cancer Research, Samsung Medical Center, Seoul, Republic of Korea. 2Department of Neurosurgery, Samsung Medical Center,
Sungkyunkwan University School of Medicine, Seoul, Republic of Korea. 3Research Institute for Future Medicine, Samsung Medical Center, Seoul, Republic
of Korea. 4Department of Systems Biology, Columbia University, New York, NY, USA. 5Department of Biomedical Informatics, Columbia University,
New York, NY, USA. 6Department of Health Sciences and Technology, Samsung Advanced Institute for Health Science and Technology, Sungkyunkwan
University, Seoul, Republic of Korea. 7Division of Life Science, Hong Kong University of Science and Technology, Hong Kong, China. 8Department of
Chemical and Biological Engineering, Hong Kong University of Science and Technology, Hong Kong, China. 9Center of Systems Biology and Human Health,
Hong Kong University of Science and Technology, Hong Kong, China. 10Division of Hematology-Oncology, Department of Medicine, Samsung Medical
Center, Sungkyunkwan University School of Medicine, Seoul, Republic of Korea. 11Department of Pathology and Translational Genomics, Samsung Medical
Center, Sungkyunkwan University School of Medicine, Seoul, Republic of Korea. 12Department of Obstetrics and Gynecology, Samsung Medical Center,
Sungkyunkwan University School of Medicine, Seoul, Republic of Korea. 13Deparment of Surgery, Samsung Medical Center, Sungkyunkwan University
School of Medicine, Seoul, Republic of Korea. 14Department of Orthopedic Surgery, Samsung Medical Center, Sungkyunkwan University School of
Medicine, Seoul, Republic of Korea. 15Department of Thoracic and Cardiovascular Surgery, Samsung Medical Center, Sungkyunkwan University School of
Medicine, Seoul, Republic of Korea. 16Deparment of Urology, Samsung Medical Center, Sungkyunkwan University School of Medicine, Seoul, Republic of
Korea. 17Samsung Genome Institute, Samsung Medical Center, Seoul, Republic of Korea. 18Department of Stem Cell Biology and Regenerative Medicine,
Lerner Research Institute, Cleveland Clinic, Cleveland, OH, USA. 19The Jackson Laboratory for Genomic Medicine, Farmington, CT, USA. 20Institute for
Cancer Genetics, Columbia University, New York, NY, USA. 21Department of Neurology, Columbia University, New York, NY, USA. 22Department of
Pathology, Columbia University, New York, NY, USA. 23These authors contributed equally: Jin-Ku Lee, Zhaoqi Liu, Jason K. Sa, Sang Shin, Jiguang Wang.
*e-mail: jyunlee@skku.edu; rr2579@cumc.columbia.edu; nsnam@skku.edu

Nature Genetics | www.nature.com/naturegenetics


Articles NaTuRE GEnETICs

the oncology clinic. First, as most solid cancers harbor multiple the samples with available WES data (Supplementary Fig. 1 and
molecular aberrations, predicting therapeutic efficacy of a targeted Supplementary Table 4). RNA sequencing (RNA-Seq) analysis was
agent based on genomic profiling alone can be a complicated pro- performed on 107 tumors to identify structure variations and gene
cess. Second, the prediction of treatment outcome extrapolated expression profiles. Tumor cell isolates were cultured under serum-
from conventional cancer cell lines may not recapitulate each free conditions27,28. Patient-tumor-derived short-term cultures were
cancer patient’s tumor. To address these challenges, we present a further subjected to targeted exome sequencing (GliomaSCAN or
comprehensive integrated approach using genomic analysis of the CancerSCAN; n =​ 122) and/or RNA-Seq to determine whether the
patient tumor, ex vivo assessment of drug effects on patient tumor PDCs retained the spectrum of genomic alterations found in the
derivatives, and in vivo validation of the selected compounds’ patient specimen (Fig. 1a and Supplementary Table 1). We observed
therapeutic efficacies. that somatic variations in major cancer-driven genes including TP53,
While patient-derived xenograft (PDX) systems respect both PTEN, PIK3CA, EGFR, NF1, APC, KRAS, and ATRX were well pre-
interpatient genomic diversity and the intratumor microenviron- served from the parental tumors to PDCs across major cancer types
ment24, the generation of PDX models exhibits a relatively lower including gliomas, metastatic brain tumors, gastric cancers, CRCs
tumor-formation rate, and requires a longer establishment period and lung adenocarcinomas (n =​  122; Fig. 1b and Supplementary
compared with patient-derived tumor cells (PDCs). In glioblas- Fig. 2a,b). Next, we measured transcriptomic similarity between
toma, it takes approximately two to three weeks to establish neu- the parental tumors and PDCs and discovered a strong posi-
rosphere-like PDCs, whereas well-defined tumor masses in vivo tive correlation (Fig. 1c and Supplementary Fig. 3). Gene fusions
are generally observed after six to seven weeks of patient tumor- were identified by chimerascan and characterized by Pegasus38.
sphere transplantation20,25,26. More importantly, chemical screen- We found that the expression levels of gene fusions were strongly
ing using PDCs presents a significant advantage over PDXs in correlated between tumors and PDCs (Supplementary Fig. 2c).
that it substantially enlarges the number of chemicals that could In addition, we observed that the mutation and gene expression
be employed in multiple doses, which is necessary to generate reli- profiles of our PDC panel were notably similar to tumor tissues
able drug response parameters. PDCs represent the unique biology (The Cancer Genome Atlas (TCGA) dataset) compared with the
of each corresponding tumor and provide a more accurate model previously established conventional cell lines (Genomics of Drug
system for assessing drug response. We and others have previously Sensitivity in Cancer (GDSC) and Cancer Cell Line Encyclopedia
shown that PDCs and PDX tumors retain the genomic and biologi- (CCLE) datasets), underscoring significant use of PDCs over
cal characteristics of tumors in situ26–29. However, drug sensitivity traditional cancer cell line models (Supplementary Fig. 4)1,16,17,23.
screening of PDCs remains at the level of proof-of-concept22,23,30, Taken together, our results suggest that PDCs provide clinically
as large-scale studies have yet to identify statistically robust con- pertinent genomic/transcriptomic proxies for primary tumors,
nections between pharmacology and genomics across multiple and are less divergent from the patient tumors than conventional
cancer types. Moreover, the number of clinical studies guided by cancer cell lines.
PDC-based drug screening is insufficient to draw concrete conclu- A 60-drug panel was used for chemical screening of PDCs
sions. In the present study, we report a unique resource, present- across multiple tumor lineages. Our panel included well-known
ing the pharmacological landscape of 462 PDCs in 14 cancer types, receptor tyrosine kinase (RTK) inhibitors, including the epider-
based on treatment with 60 anticancer agents. We show that PDCs mal growth factor receptor (EGFR), platelet-derived growth fac-
faithfully recapitulate the molecular profiles (RNA expression and tor receptor (PDGFR)/vascular endothelial growth factor receptor
mutations) of the original primary tumors. Using a combination (VEGFR), and phosphoinositide 3-kinase (PI3K)/protein kinase
of machine learning and statistical methods, we have discovered B (AKT)/mammalian target of rapamycin (mTOR) (PAM path-
lineage-specific drug sensitivities and molecular correlates of drug way), as well as multi-target drugs and inhibitors for proteasome,
sensitivity and resistance. Finally, we demonstrate a high clinical poly (ADP-ribose) polymerase, and histone deacetylase (HDAC)
concordance rate between clinical response and drug susceptibility (Supplementary Tables 5 and 6)39–52.
prediction in retrospective studies. Cell viability was determined after six or seven days of drug
treatment using multi-parameter drug sensitivity analyses, includ-
Results ing the half-maximal inhibitory concentration (IC50) and area
Genetic and transcriptomic similarity between primary tumors under the curve (AUC) of the dose–response curve (DRC; Fig. 1a
and PDCs. To establish a large-scale PDC library, we derived 462 and Supplementary Tables 7 and 8)53–55. The drug panel sensitiv-
short-term tumor cell cultures from surgically resected tumor spec- ity profiles were highly concordant among technical replicates
imens or ascites-derived tumor cells that were isolated from 14 dif- (Supplementary Fig. 5a). To evaluate whether tumor cell expansion
ferent cancer types, including malignant gliomas, metastatic brain affected biological properties, including drug sensitivity, we mea-
tumors, gastric cancers, colorectal cancers (CRCs), lung, breast, sured post-drug treatment cell viability using four different sets of
and ovary cancers, renal cell carcinoma and sarcomas (Fig. 1a the matched PDCs across various in vitro passages or expansion
and Supplementary Table 1). To identify somatic mutations and stages (Supplementary Fig. 5b). Drug sensitivities of biological rep-
copy-number alterations (CNAs), exome sequencing was per- licates from the same PDCs at different passages showed significant
formed on 368 tumor specimens and matched normal specimens. correlations with minimal variations.
Among them, 94 samples were subjected to whole-exome sequenc-
ing (WES), while 274 tissue specimens were sequenced for full cod- Pan-cancer pharmacological profiling identifies lineage-specific
ing exons of 80 commonly mutated cancer genes (CancerSCAN; drug sensitivity. Next, we established a pharmacological landscape
Supplementary Tables 1 and 2)31–34. Forty-one specimens from of 462 PDCs from 14 different tumor lineages, based on antican-
malignant gliomas were analyzed using a massive parallel tar- cer drug responses to 60 compounds. A total of 27,720 drug–PDC
geted sequencing platform, covering exons of cancer-driven and/ combinations were evaluated and analyzed (Supplementary Fig
or glioma-associated genes (GliomSCAN; Supplementary Tables 1 6). The median IC50 values varied from 0.003–53.22 μ​M, where
and 3)1,9,35. We used the statistical algorithm for variant frequency 1 μ​M is a commonly applied threshold for pharmacological rel-
identification (SAVI2) to identify somatic single-nucleotide vari- evance (Supplementary Fig. 6a). A subset of compounds, includ-
ants, as well as short insertions and deletions36. Somatic variants ing AUY922 (heat shock protein), BEZ235, PKI-587 (PI3K/
with a mutant allele frequency of >​5% were considered. To explore mTOR), bortezomib, carfilzomib (proteasome), neratinib (EGFR),
copy-number variations (CNVs), EXCAVATOR37 was applied on panobinostat (HDAC), and trametinib (mitogen-activated

Nature Genetics | www.nature.com/naturegenetics


NaTuRE GEnETICs Articles
a b Conserved
TP53

Cancer patients Genomic profiling


(462 samples from 14 lineages)
Cancer types (14)
Patients (355)
GBM Brain metastases GBM (75)
Tissues (385)
WES (94)
WES blood (68)
GS (38)
Breast GS blood (28)
Lung CS (274)
Others (14)
RNA-Seq (107) PIK3CA PTEN
Stomach PDC (128) Sarcoma (7)
Colon PDC_GS (35) EGFR
Renal cell
Ovary carcinoma (13) NF1
PDC_RNA (34) APC ARID1A, CDH1, ATRX
… PDC_CS (89) KRAS
Ovarian cancer (13)
KDR, NCOR1, JAK3 MLH1
ERBB4, CDKN2A, FGFR2
DNA sequencing NSCLC (SqCC) (18) CREBBP, BRAF, RB1
Tissue Mutation PTPN11
CNA… NSCLC (adenocarcinoma) (20)

Sequencing Tissue PDC


Metastatic brain (others) (12)

Metastatic brain (lung) (22)


c Spearman’s correlation
PDC Colorectal cancer
RNA-Seq (48)

0
.5

5
Breast cancer (15)

0.
Expression

–0
Fusion…
g
ru Gastric cancer (98)
e –d tion

mRNA expression on tissue


en ia
G soc
as Clinical response prediction
Chemical response (60 drugs) Before After
treatment treatment
120 12
% survival

80 10
e
8 sitiv
40 6 Sen
4 Resista
0 nt
–3 –2 –1 0 1 2 2
ATP-based cell log drug concentration 0
survival analysis (µM)
0
0
0
0
0
0
15
20
25
30
35
40

Compute DRC mRNA expression on PDC


z-score calculation
Calculate AUC

Fig. 1 | Patient tumor and derived cell resources for pharmacogenomics analysis. a, Overview of the procedure for pharmacogenomic analysis in PDCs.
A total of 462 PDCs from 14 cancer types were isolated. Genomic contexts were analyzed to identify somatic variants and/or gene expression profiles.
Comprehensive genomic profiles from 14 cancer lineages are summarized in a Circos plot, demonstrating detailed data structure and size for each type
of available molecular data. Short-term cultured PDCs underwent drug sensitivity screening to 60 molecular targeted agents. The clinical feasibility of
PDC screening-guided precision therapy was evaluated. NSCLC, non-small-cell lung carcinoma; SqCC, squamous cell carcinoma; GS, GliomaSCAN;
CS, CancerSCAN. b, Three-dimensional bubble plot showing the frequency of somatic non-synonymous mutations exclusively in tissue (red, left axis),
exclusively in PDC (black, right axis), and common to the two (yellow, upper axis). In total, 122 samples with DNA sequencing data from tumor tissue
and matched PDC samples were considered in this analysis. c, Comparison of mRNA expression profiles between primary tissue and PDC on 24 paired
samples with matched RNA-Seq data. Spearman’s correlations of mRNA expression between tissue and PDC are shown as a heat map. Paired samples are
located along the diagonal.

protein kinase/extracellular signal-regulated kinase), showed Supplementary Table 9). Interestingly, PDCs from malignant
notably high drug activities. In contrast, several agents, such as gliomas demonstrated considerable resistance to erlotinib, gefi-
dabrafenib (B-raf proto-oncogene), olaparib (poly (ADP-ribose) tinib (EGFR) and BKM120 (PI3K), while the CDK4/6 inhibitors
polymerase), sotrastaurin (protein kinase C), vismodegib (hedge- LY2835219 and palbociclib were significantly more sensitive in
hog), and XL147 (PI3K) exhibited low antitumor activities, each glioma PDCs compared with the rest (Fig. 2a and Supplementary
with median IC50 values of >​10  μ​M. RTK blockers targeting EGFR, Fig. 6d)1,56–58. To explore dynamic pharmacological associations
c-MET, fibroblast growth factor receptors (FGFR), PDGFR and among different drug profiles, we performed topological data
PAM showed a wide spectrum of drug sensitivities, suggesting a analysis (TDA) using Mapper (see Supplementary Methods)—a
strong association between the pharmacological drug response and computational technique to reduce the dimensionality of large
genomic aberration (Supplementary Fig. 6b). datasets while retaining the local high-dimensional structure13,59–61
As tumor lineages portray diverse drug responses16,17, we of lineage-specific drug associations (Fig. 2b and Supplementary
first analyzed lineage-specific drug sensitivity within our cohort Table 10). PDCs isolated from CRCs were relatively more resistant
(Fig. 2a). Overall, we discovered robust lineage-specific drug sen- to multiple PAM inhibitors, such as BEZ235, PKI-587, AZD2014,
sitivities across multiple cancer types, among which significant everolimus, and BYL719 (Fig. 2b). Additionally, gastric cancer
diversity was observed in glioma, breast, CRC, and gastric cancer. PDCs were highly sensitive to PI3K inhibitors, including BYL719
Hierarchical clustering of drug sensitivities (normalized AUCs) and BKM12062, which were comparatively less potent in malig-
identified three distinct clusters, enriched with gastric cancer, nant gliomas (Fig. 2b,c and Supplementary Fig. 7). Consistently,
glioma, and CRC, suggesting lineage-specific drug response pat- PAM pathways were found to be more activated in gastric cancer
terns (Supplementary Fig. 6c). Of note, EGFR inhibitors (erlo- PDCs compared with glioblastoma (GBM) PDCs (Fig. 2d)63–65.
tinib, dacomitinib and afatinib) showed significantly high activity These observations suggest a higher likelihood of antitumor activ-
in both gastric and breast cancer models, while demonstrating ity of PAM inhibitors in gastric cancer patients compared with
minimal responses in malignant gliomas and CRC (Fig. 2a and GBM patients.

Nature Genetics | www.nature.com/naturegenetics


Articles NaTuRE GEnETICs

a b
Erlotinib BYL719 (gastric)
Gefitinib Relatively sensitive
CO.1686
AEE788 Relatively resistant
AC480
Neratinib
Afatinib
Canertinib
Lapatinib
Dacomitinib
BKM120 20
BYL719 BKM120 (gastric)
XL147

–log10[TDA q value]
Sensitive AZD2014
Everolimus PKI.587 (colorectal)
Resistant PKI.587
BEZ235
XL765 BYL719 (glioma)
AZD5363
Axitinib Erlotinib (gastric)
Cediranib
Imatinib CO.1686 (gastric)
Pazopanib
Regorafenib 10 LDK378 (gastric)
Sunitinib BKM120 (glioma)
Tandutinib Dacomitinib (breast)
Tivozanib Dasatinib (glioma)
Vandetanib
Cabozantinib AUY922 (other brain)
Crizotinib Afatinib (breast)
Foretinib
INCB28060 Bortezomib (gastric)
LEE011 BGJ398 (other solid)
LY2835219
Palbociclib
AZD4547
BGJ398 0
Dovitinib
Bosutinib
Dasatinib
Nilotinib –1.0 –0.5 0 0.5 1.0
Selumetinib
Trametinib log2[fold-change AUC]
Bortezomib
Carfilzomib c d
ABT.199
LDK378
Ruxolitinib 0.4
AUY922 P = 2.53 ×10–10
Vemurafenib
Ibrutinib
Olaparib q = 7.4 × 10–28
PF.04449913
Veliparib 0.2

Enrichment score
EGFR inhibitor Dabrafenib
PI3K inhibitor Erismodegib
Sotrastaurin Gastric
VEGFR inhibitor Vismodegib Fraction %
LGK.974 0 100
Met inhibitor Panobinostat 0
CDK4/6 inhibitor
N eta st Lu ore ast
C Br ma

Saino n

a
LC tati ic b g (o as al
(a br rain thetric

lid m )
no n ( lun )

er ain q )

ca va er
ca ors

rc ma
O S car the g)
O er b LC noms)
so tu CC
de ai ( rs

th r (S a

rc ria

om
SC s at n G ct

ll O nc
th C ci r
ol e

FGFR inhibitor
lio
G

Bcr-Abl inhibitor BYL719


q = 8.0 × 10–15 –0.2
ce

MEK inhibitor
N
c

al
en

AUC
M eta

Proteasome inhibitor
R
M

Low High
Gliomas Gastric GBM

Fig. 2 | Therapeutic landscape of PDCs and lineage-specific responses a, Identification of tumor lineage-specific drug associations (n =​ 462 biologically
independent samples). A two-sided Wilcoxon rank-sum test was applied to determine the relative differences in drug sensitivity between specific tumor
types and the rest. Only significant associations are marked (q value <​ 0.05). Drugs are ordered based on their known targets. NSCLC, non-small-cell lung
cancer; SqCC, squamous cell carcinoma. b, Volcano plot representation of TDA analysis showing the magnitude (fold-change) and significance (TDA
q value) of all tumor–drug associations (n =​ 462 biologically independent samples). Each circle represents a single tumor–drug interaction, and the size is
proportional to the cohort size of that tumor. c, Distribution of gastric, glioma PDCs and BYL719 drug AUC profiles over the topological representation of
PDCs (n =​ 462 biologically independent samples). Each node represents a set of PDCs with similar AUC profiles. A PDC can appear in several nodes, and
two nodes are connected by an edge if they have at least one PDC in common. The P values were calculated using a Pearson’s correlation test between
the fraction distribution of gastric or glioma cell lines and mean AUC values of BYL719 drug over the nodes, and they were adjusted using the Benjamini–
Hochberg method. d, Violin plots measure the activity level of the PAM pathway on gastric and GBMs using TCGA RNA-Seq datasets (n =​ 100 biologically
independent samples for each group). We adopt the enrichment score derived from ssGSEA as an assessment. The P value was calculated from a two-
sided Wilcoxon rank-sum test. Horizontal lines within the violin plot represent 0.25, 0.50, and 0.75 quantiles.

Genomic predictors of drug sensitivity/resistance based on PDC proto-oncogene), BYL719 (PI3K) and trametinib (MEK) (Fig. 3b)73–
genome–drug mapping. Genomic variations are increasingly 76
. Although KRAS mutations drive tumor progression through con-
being utilized as reliable biomarkers for predicting clinical response stitutive activation of mitogen-activated protein kinase signaling
to cancer therapy4,5,10,11,16,17. To identify significant genomic corre- pathway, conflicting evidence has been reported for the efficacy of
lates of pharmacological sensitivity in PDC models, we assessed MEK target therapy in KRAS-mutant cancers77–80. We showed that
individual drug profiles across 360 tumor cells based on single treatment with MEK inhibitors increased the efficacy of EGFR fam-
genomic alterations (Fig. 3a and Supplementary Table 11). EGFR ily blockers (dacomitinib and gefitinib) in a KRAS-mutant (G12V)
somatic alterations, including single-nucleotide variations, copy- CRC cell line, SW48080, suggesting that simultaneous targeting of
number amplifications, and structure variations, were mostly found EGFR and MEK signaling pathways could potentially treat KRAS-
in GBM tissues and conferred increased sensitivity to relevant mutant CRC patients (Fig. 3c)19,81,82. Furthermore, mutation in the
EGFR inhibitors, such as dacomitinib, and interestingly to vande- adenomatous polyposis coli (APC) gene was significantly corre-
tanib (VEGFR) and ibrutinib (Bruton’s tyrosine kinase (BTK))66–68. lated with minimal responses of multiple drugs, including PKI-587
Previous studies have demonstrated a strong association between (PI3K/mTOR), AZD5363 (AKT), and AZD2014 (mTOR) (Fig. 3a
KRAS mutation and therapeutic resistance to PAM pathway inhibi- and Supplementary Fig. 8), consistent with previous observations
tors, including PKI-587, AZD5363, AZD2014, and multiple EGFR that activation of the β​-catenin signaling pathway may elicit multi-
inhibitors, such as neratinib, erlotinib, afatinib, dacomitinib, and drug resistance across various cancer types83–85.
canertinib69–72. In contrast, we discovered that KRAS mutation con- Activation of various RTKs is a key molecular feature of GBM.
ferred increased sensitivity to a specific group of drugs, including As RTK-targeting approaches against GBM have been widely pur-
dasatinib (SRC proto-oncogene and breakpoint cluster region-ABL sued, we examined pharmacological sensitivities of various RTK

Nature Genetics | www.nature.com/naturegenetics


NaTuRE GEnETICs Articles
a
PKI.587
APC mutation Sensitive
Resistant

6 PKI.587
KRAS mutation
AZD2014 AZD5363
SMAD4 KRAS mutation
mutation Bosutinib
AZD5363
–log10[rank-sum P value]
NF1 mutation Vandetanib
APC mutation
BYL719 Axitinib EGFRvIII
4 ABT.199
NF1 mutation APC mutation
Neratinib CIC mutation
AZD2014 ABT.199
APC mutation KRAS mutation Dacomitinib
DYSF mutation
BYL719 EGFR amplification
Ibrutinib
KRAS mutation
EGFRvIII Trametinib
Dasatinib
KRAS mutation KRAS
2 mutation

–0.4 –0.2 0 0.2 0.4


log2[fold-change AUC]

b c
2
–log10[P ]

0 120

% survival
80
Sensitive to the KRAS mutation
4
Resistance to the KRAS mutation Control
40 Trametinib (0.1 μM)
Trametinib (1 μM)
ni s

D ib
lo b
er ib

PK 363
87
BY tinib
Ve me 19
ur ib

Y b
C xit 2
D ner ib
om nib
Af inib
Pa ZD inib

no 4
Ev ilo t
ol b
N sta

Su imu
AU eni
A 92

bi 01

er tini

Er tini

AZ atin
m tin

a in

N tin

I.5
Tr L7

ac ti
it
at
no 2

5
a

af

0
as

a
D

–2 –1 0 1 2
log dacomitinib (μM)
d
120
% survival

EGFR PAM pathway 80


MET Others
Control
40 Trametinib (0.1 μM)
Ibrutinib Trametinib (1 μM)

0
–2 –1 0 1 2
log gefitinib (μM)
0
–2
–1

1
2

z-score
e f 1.0
Ibrutinib response
Fraction survival

EGFR CN (P = 0.00017) 0.8

EGFRvIII (P = 0.0026) 0.6


EGFR expression 0.4 Vehicle
0.2 Ibrutinib (50 mg/kg)
EGFR copy number log2[CN] > 1 log2[CN] > log2[CN] < log2[CN] < P = 0.002
0.5 –0.5 –1 0
EGFR expression RPKM >100 RPKM > 50 0 30 60 90 120 150 180
Time (days)

Fig. 3 | Pharmacogenomic interactions in PDCs. a, Volcano plot representation of a correlation analysis showing the magnitude and significance of
gene–drug associations (n =​ 360 biologically independent samples). b, Waterfall plot enumerating significant associations between KRAS mutation and
drug sensitivity (n =​ 360 biologically independent samples). The two horizontal dashed lines indicate statistical significance. c, SW480 (KRAS G12V) was
treated with DMSO (control) or trametinib (0.1 or 1 mM), followed by incubation with two EGFR inhibitors (dacomitinib (top) and gefitinib (bottom)). Cell
viability for each dose was normalized to DMSO or trametinib (0.1 or 1 mM) treatment-only cells. d, Probability distribution of drug–target families over
the topological network. Each node represents a set of drugs with similar AUC profiles. A drug can appear in several nodes, and two nodes are connected
by an edge if they have at least one drug in common. The colors of the nodes correspond to mean RGB values of drug families. Ibrutinib belongs to three
nodes on the network encompassed by an oval. e, Drug sensitivities to ibrutinib in 67 PDCs. The red color in the heat map represents sensitivity, while the
blue color indicates resistance. EGFR alterations, including genomic amplification, vIII, and expression are shown. f, Kaplan–Meier survival plots for the
sample P2.T (EGFR amplification/vIII) orthotopic mice model. Once the intracranial model was established, vehicle (0.5% methylcellulose) or ibrutinib
(50 mg) was administrated orally for 5 consecutive days and 2 days of resting period per cycle (n =​ 8 per group). P values in a, b and e: two-sided Wilcoxon
rank-sum test. P values in f: two-sided log-rank test.

Nature Genetics | www.nature.com/naturegenetics


Articles NaTuRE GEnETICs
a Drug data b c
300 300
120 P = 1.93 × 10–5 P = 6.62 × 10–5
% survival

AUC panobinostat

AUC panobinostat
80 PCC = 0.48 PCC = 0.48
mRNA SIK2
40 Others ARID1B 200 200
HDAC4
0 –log10[P] PCGF5 mutation
–3 –2 –1 0 1 2
1
log drug concentration (μM) 75 2 100 100
3 ZHX2
REST

Frequency of bootstrap
Known 50 50
4
drug target
IDH1 mutation
CDKN2A|CNV

1.0 1.5 2.0 2.5 3.0 3.5 2.0 2.5 3.0 3.5 4.0
50
EGFRvIII

WDR77 HDAC4 expression SIK2 expression


PIK3CA

INCB28060

Association
d
XL147
Panobinostat

EGFR|CNV

finding WDR73
Ruxolitinib

FBXW7

1.6 1.6
PKI.587

Dacomitinib

PRONEURAL
PTEN|CNV

Bortezomib

POLI
Afatinib

KRAS

ATRX

GNAQ

1.4 Non-target 1.4 Non-target


Ibrutinib

KDM5D

Relative cell viability

Relative cell viability


CCDC6
LEE011 LDK378
BEZ235

CDK4 mutation
Lapatinib

LGK.974
Nilotinib

shHDAC4-1 shSIK2-1
ARID1A
Dabrafenib
EGFR

Biological 1.2 1.2


Axitinib
Olaparib

GATAD1
TP53

Neratinib

PF.04449913

25
TP73|CNV CO.1686

IDH1
Pazopanib

Cediranib

Canertinib Palbociclib

Gefitinib

ATM
Dasatinib
Tandutinib

shHDAC4-2
ABT.199

shSIK2-2
LY2835219

network
PTEN

AUY922 SMAD4
BYL719
BGJ398

AEE788
APC

Imatinib

Erlotinib
Everolimus Vandetanib

Bosutinib

1.0 1.0
Vemurafenib
Dovitinib

Sunitinib

AC480 Cabozantinib
NF1
RELN

shHDAC4-3
PIK3R1
KDR
BKM120 Trametinib Selumetinib
PDGFRA|CNV

0.8 * 0.8
0.6 0.6 *
**
0.4 ** 0.4
0 * **
Genomic profiling 0.2
** 0.2 * *
–0.2 –0.1 0 0.1 0.2 0 0
1.2 dose 4.8 dose 1.2 dose 4.8 dose
Sensitive Averaged weights Resistant
(nM) (nM) (nM) (nM)
Panobinostat Panobinostat

Fig. 4 | Genomic and transcriptomic correlates of panobinostat sensitivity a, Schematic overview of the dNetFS. In brief, the dNetFS integrates genomic/
pharmaceutical data, protein–protein interaction networks, and previous knowledge of drug–targets interactions to prioritize genetic and gene expression
features of PDCs that predict drug responses. b, Predictive features of panobinostat response identified by the dNetFS are plotted for their frequency
and effect size. Associations are colored in red for expression features and blue for others. Node size is proportional to the single drug–feature linear
correlation. c, Scatter plot showing linear correlation between panobinostat AUC and HDAC4 expression (left) and SIK2 expression (right) (n =​ 69
biologically independent samples). The correlation coefficients (PCCs) and the P values were obtained using Pearson’s correlation tests. d, Drug-response
assessment of panobinostat (1.2 or 4.8 nM) with shRNA-mediated knockdown of HDAC4 or non-target (left) and SIK2 or non-target (right). Cell viability
for each dose was normalized to sole transduced cells only. Data are means ±​ s.d. of n =​ 3 technical replicates. Experiments were repeated three times with
similar results. *P ≤​ 0.05; **P ≤​ 0.01; two-tailed t-test.

inhibitors in GBM PDCs via drug-centric TDA (see Supplementary were further confirmed using different ibrutinib compounds
Methods) (Fig. 3d). Interestingly, we discovered several compounds obtained from five different vendors (Supplementary Fig. 11c).
that are clustered with different classes of agents, including erlotinib Furthermore, in vivo assessment of ibrutinib (50 mg/kg orally every
with PI3K, foretinib and BGJ398 with VEGFR, cediranib with FGFR, day for 5 days) demonstrated prolonged survival and reduced tumor
and ibrutinib with EGFR compounds (Fig. 3d and Supplementary volume in PDX models (Supplementary Fig. 11d). Histological and
Fig. 9). The ibrutinib–EGFR TDA result was consistent with pre- immunocytochemical analyses showed attenuated expression of
vious genomic and pharmacological correlate analysis (Fig. 3a). In EGFR and its downstream molecules (Supplementary Fig. 11d,e).
preclinical and clinical studies on hematologic malignancies, ibru- Collectively, our results indicate that EGFR amplification/vIII could
tinib has been reported to promote cellular arrest via inhibition of be a potential biomarker for sensitivity to ibrutinib in GBM.
BTK phosphorylation68,86–88. Although a potent antitumor effect of
ibrutinib was previously demonstrated in non-small-cell lung can- Integrative multi-layer data analysis identifies molecular deter-
cer cells67, it was exclusive to tumors harboring the EGFR T790M minants that dictate drug sensitivity to panobinostat. Prediction
mutation, which is an erlotinib resistance-associated event. Notably, of drug response is a complex process that often depends on mul-
we showed that EGFRvIII and EGFR amplification were signifi- tiple variables, including genetic and transcriptomic profiling of
cantly linked to increased sensitivity to ibrutinib (Fig. 3e). To deter- tumors16,23,48. Given such extensive diversity, we next applied the dif-
mine whether the sensitivity to ibrutinib was exclusively associated fusion kernel based network method for feature selection of drug
with EGFR aberration, we surveyed the transcriptome levels of BLK, sensitivity (dNetFS)—a regression model-based analysis to uncover
BMX, and BTK—the proposed targets of ibrutinib. Expression levels cooperative interactions among multiple layers of variables90. The
of these genes were generally low in GBMs and, more importantly, dNetFS integrates protein–protein interaction networks and previ-
there was no significant correlation between their transcriptional ous knowledge of drug target interactions to prioritize genetic and
levels and ibrutinib activity (Supplementary Fig. 10a). In contrast, gene expression features that are potentially associated with drug
the ibrutinib response significantly correlated with EGFR messen- responses (Fig. 4a and Supplementary Fig. 12a). Using the dNetFS,
ger RNA (mRNA) expression levels (Supplementary Fig. 10b). This we detected previously known associations between EGFR, FGFR,
observation was significantly enhanced when only the EGFRvIII- CDK4/6, KIT, and MET alterations and drugs that directly target
positive tumors were selected. these molecules (Supplementary Fig. 12b). In addition, we also
To further validate and explore the ibrutinib–EGFR association found that PDCs harboring an EPHB4 mutation showed substan-
and its biological effects in GBM, we performed a series of func- tial sensitivity to EGFR inhibitors, including afatinib, dacomitinib
tional in vitro and in vivo experiments. Consistent with previous and AEE788. EPHB4 encodes for the RTK EphB4, which governs
findings, ibrutinib significantly decreased clonogenic growth of tumor progression and portrays dismal prognosis across various
EGFR amplification/vIII-harboring GBM PDCs compared with cancer types91–95. Our result suggests that EPHB4 mutation could
those with the wild-type EGFR gene (Supplementary Fig. 11a)89. also confer increased sensitivity to EGFR target therapies, for which
Immunoblot analyses of EGFR and its downstream effectors, pSTAT3, its underlying mechanism needs to be further evaluated.
pERK, pAKT and pS6K, showed robust inhibition in the presence We further discovered a group of molecular predictors for
of ibrutinib treatment (Supplementary Fig. 11b). These activities panobinostat (HDAC)—a drug that has rarely been analyzed

Nature Genetics | www.nature.com/naturegenetics


NaTuRE GEnETICs Articles
a c d P = 0.001011
49 GBM samples with EGFR alteration NRG1 400
Altered Wild type
High

–6
EGFR CNV
EGFR mutation

10

4
EGFRvIII 300

×
×
EGFR fusion AEE788

10
4

AUC/expression
–5
EGFR-SEPT14 Afatinib

=
Erlotinib Dacomitinib

P
P59.TR1p
P59.TR2p
P58.TI1p
P58.TI2p

P10.TRn
P10.TRp

P59.TRn

P56.TR2
P56.TR3

P41.TR1
P41.TR2
P41.TR3
P5.TIR
P42.TI3

P72.T1

P42.TI2

P42.TI1
P55.TI1
P55.TI2

P77.T1

P21.TI2
P21.TR
P76.T1

P83.T1

P81.T1
P58.TIn

P36.TIp

P74.T1
P36.TIn
P21.TI1
P14.TR
P24.T2

P57.TI3
P57.TI2
P24.T1

P64.T1
P69.T1
P75.T1
P33.T
P34.T
P23.T

P19.T
P47.T

P48.T
P39.T
P45.T
P35.T
P17.T
P37.T
P13.T

P43.T
P46.T
P44.T
P30.T
P50.T
P11.T
P5.TIL

P54.T
P10.TI

P38.T
P53.T
P18.T
P15.T
P22.T
P12.T
P31.T

P40.T
P20.T
P28.T

P5.TR

P27.T
P51.T
P52.T
P16.T
P14.TI
P26.T

P49.T
P2.T

P9.T

P3.T

P7.T

P8.T

P6.T
BMS-599626

AUC
CO-1686
200
Canertinib
Dacomitinib
Erlotinib

b
Gefitinib
EGFR 100 Lapatinib

0.

3
02
00
Neratinib

0.
7
80 Low
TP73 CNV 0
NRG1

Non-target shNRG1
TP73 CNV
60 EGFR
e f –5
Frequency of bootstrap

EGFR fusion
P = 3.926 × 10
EGFR mutation NRG1
0.4 300
NRG4 CDK4 mutation
TP73 CNV

Median ES in resistant cases


40 NRG4
CDKN2A CNV
NRG1 0.2 KRAS signaling
AEE788
AEE788 Myc targets
CDKN2A BMS-599626
EGFR mutation
Afatinib PAM pathway signaling 200
EGFR CNV CI-1033

AUC
BMS-599626
CO-1686
EGFR fusion Canertinib 0
NRG1 CO-1686 Erlotinib
20 Dacomitinib Higher in Gefitinib
CDK4 CNV NRG1
Erlotinib Lapatinib
–0.2 resistance
PIK3R1 mutation
TP73 CNV Gefitinib 100 Neratinib
PIK3CA mutation
Lapatinib
NRG1
EGFR mutation
NRG4
Neratinib Higher in
0 –0.4 sensitive
–0.2 –0.1 0 0.1 0.2 –0.4 –0.2 0 0.2 0.4 DMSO PAM
inhibitors
Sensitive Averaged weights Resistance Median ES in sensitive cases

Fig. 5 | Predictive biomarkers for response to EGFR inhibitors in EGFR-altered GBM PDCs. a, Mutational landscape of EGFR alterations in the GBM
cohort. Each column represents individual tumor sample and detailed patient information are included in the Supplementary Table 1. b, For the
10 EGFR inhibitors, top drug–feature associations identified by the dNetFS are plotted for their frequency and effect size (n =​ 49 biologically
independent samples from a). Node size is proportional to the single drug–feature linear correlation. c, Gene expression profiles of EGFR and
NRG1, and AUC drug-response profiles of erlotinib and dacomitinib, over the topological representation (n =​ 44 biologically independent samples).
d, Drug-response assessment of EGFR inhibitors with shRNA-mediated knockdown of NRG1 or non-target. Cell viability for each dose was normalized
to shNRG1 or non-target-transduced cells only (n =​ 10 independent experiments with 3 technical replicates). e, Comparisons of cancer pathway
activities between two EGFR-altered GBM subgroups that were most sensitive and most resistant. We adopt the enrichment score (ES) derived from
ssGSEA for assessment. f, Drug-response assessment of EGFR inhibitors with PAM inhibitors or DMSO. Cell viability for each dose was normalized to
PAM- or DMSO-treated cells only. Mean AUC values for 4 PAM inhibitors (BYL719, BKM120, BEZ235, and AZD2014) are plotted (n =​ 8 independent
experiments with 3 technical replicates). P values in c: Pearson’s correlation test. P values in d and f: two-sided Wilcoxon rank-sum test. Horizontal
lines within the violin plots represent 0.25, 0.50, and 0.75 quantiles.

in pharmacogenomics studies (Fig. 4b and Supplementary Genomic correlates of EGFR inhibitor sensitivity in an EGFR-
Fig. 12c)96–98. We identified that high expression of the RE1- altered subgroup. Clinical response to molecular-targeted drugs is
silencing transcription factor gene (REST) confers cellular vulner- likely to depend on the multiplicity of biological processes in tumor
ability to panobinostat, consistent with the previous notion that cells, even in subpopulations harboring alterations in the corre-
HDAC inhibitors could potentially be employed as therapeutic sponding target molecules107–109. Genomic aberrations of EGFR,
agents against REST-positive tumors99. Furthermore, mutations including copy-number amplification, mutation, and structure
in IDH1 and ARID1B, as well as high expression of zinc finger variations, largely contribute to glioma malignancy in over 50%
homeobox 2 (ZHX2—a transcriptional repressor), polycomb of patients1,110. However, targeting EGFR has shown disappointing
group ring finger 5 (PCGF5—a chromatin modifier), HDAC4 clinical outcomes due to various resistance-associated biological
(a transcriptional repressor and class IIa HDAC) and salt inducible mechanisms48,111–114. Therefore, we analyzed the pharmacological
kinase 2 (SIK2—an upstream regulator of class IIa HDACs), were response of 10 EGFR inhibitors on 49 glioma PDCs harboring EGFR
significantly associated with resistance to panobinostat (Fig. 4b, alterations, including copy-number amplification (n =​  48), muta-
Supplementary Fig. 12c and Supplementary Table 12)100–103. tions (n =​  10), vIII (n =​ 13), and fusions including EGFR–SEPT14
Interestingly, although panobinostat is a non-selective inhibi- (n =​  4) (Fig. 5a). All ten EGFR inhibitors showed similar activities
tor for HDAC, transcriptome expressions of HDAC4 and SIK2 across GBM PDCs: a subset of tumor cells that are sensitive to a
were highly correlated with its resistance (Fig. 4c). Consistently, particular EGFR compound was more likely to respond to other
small hairpin RNA (shRNA)-mediated knockdown of HDAC4 or EGFR inhibitors, and vice versa (Supplementary Fig. 10b). EGFR-
SIK2 conferred increased sensitivity to panobinostat (Fig. 4d and altered tumors clustered into two major groups based on their cel-
Supplementary Fig. 12d). Panobinostat was reported to confine lular responses to EGFR inhibitors. Next, we applied the dNetFS
HDAC4 in the cytoplasm104. Cytoplasmic HDAC4, which could and TDA to investigate the molecular correlates of drug sensitiv-
also be induced by SIK2 activation, may attenuate the sensitiv- ity to EGFR inhibitors (Fig. 5b,c and Supplementary Fig. 13). As
ity to panobinostat by previously proposed mechanisms, such expected, EGFR transcriptional expression, mutations, and fusions
as activation of the HDAC–mitogen-activated protein kinase– conferred sensitivity to multiple EGFR inhibitors (Supplementary
activator protein-1 signaling axis or hypoxia inducible factor Table 13). Strikingly, neuregulin 1 (NRG1) transcriptional expres-
1-alpha functions100,103–106. These associations represent candidate sion emerged as a robust hit against cellular resistance to a variety
genomic markers of drug sensitivity that may be useful for future of EGFR inhibitors. We further observed similar unresponsive-
biomarker-adaptive clinical trial design. ness behavior to multiple EGFR inhibitors via TDA (Fig. 5c and

Nature Genetics | www.nature.com/naturegenetics


Articles NaTuRE GEnETICs

a Clinical responses Cancer types d


2 Stable/progressive disease P10 (52/F): GBM, P30 (41/F): GBM,
Gastric cancer
Complete/partial response
Glioblastoma
PTEN p.Gly165Arg PTEN deletion
1 Lung adenocarcinoma
Atypical meningioma 400
z -score

0 300 P30.T

AUC
P10.TRp
–1 200

100
–2
Day –6 Day 0 Day 17 0 Day –36 Day 0 Day 60
Pan-cancer
Lapatinib (ERBB2)
Gefitinib (EGFR)
Lapatinib (ERBB2)
Everolimus (mTOR)
Gefitinib (EGFR)
Afatinib (EGFR)
Pazopanib (VEGFR)
Classified (MET)
Crizotinib (MET)
Everolimus (mTOR)
Afatinib (EGFR)
Vandetanib (VEGFR)
Crizotinib (MET)
Afatinib (EGFR)
Lapatinib (ERBB2)
Crizotinib (MET)
Everolimus (mTOR)
Crizotinib (MET)
Everolimus (mTOR)
Gefitinib (EGFR)
Everolimus (mTOR)
Sunitinib (PDGFRA)
Classified (MET)
Gefitinib (EGFR)
Pazopanib (VEGFR)
Lapatinib (ERBB2)
Crizotinib (MET)
Afatinib (EGFR)
Sunitinib (PDGFRA)
Afatinib (EGFR)
Pazopanib (VEGFR)
Everolimus (20 mg d–1) (n = 432) Everolimus (20 mg d–1)

b c e f P436 (62/M): gastric cancer,


P = 0.002337 P215 (49/F): atypical MNG HER2 amplification
2 Before lapatinib
100 400
1 80 400
Sensitivity (%)

300
z-score

0 60
AUC

AUC
40 300
–1 AUC: 0.8125 200 P436.T Lapatinib (1,250 mg, 84 d)
20 P215.T1
95% CI: 0.6582–0.9668
–2 P value: 0.003056
0 200
0 20 40 60 80 100 100
CR/PR SD/PD Day –124 Day –14 Day 0 Day 45
100 – specificity (%) Pan-cancer Pan-cancer
(n = 16) (n = 15) (n = 456) (n = 462)
Sunitinib (25 mg d–1)

Fig. 6 | Clinical feasibility of PDC drug-screening-guided precision oncology. a, Bar graph representing normalized AUCs (z-scores from the pan-cancer
dataset) of the indicated drugs (n =​ 31 biologically independent samples). Clinical response was determined according to RECIST. Multiple-target drugs
are classified based on their corresponding representative genomic targets. b, Representative box plot of a. CR, complete response; PD, progressive
disease; PR, partial response; SD, stable disease. Abbreviations are as per RECIST criteria. Box plots span from the first to third quartiles, and the whiskers
represent the 1.5 interquartile range. c, ROC curve plotted for the sensitivity versus 100 −​ specificity values for predicting clinical response rates using the
z-scores in a. CI, confidence interval. d–f, T1-weighted contrast enhanced magnetic resonance images (d), T2-flare magnetic resonance images (e) and
computed tomography images (f) for the indicated patients before and after drug treatment. Patient labels show the patient number, followed by their age
and sex (in brackets), cancer type and mutation. Gray circles on the bar represent days when the presented images were obtained. Day 0 refers to the day
when drug treatment started. Red arrows indicate measurable or progressed tumors. Orange arrows represent partial response. The vertical scatter plots
show the AUCs of the indicated drugs for the pan-cancer AUC reference dataset. AUCs of the PDCs isolated from the illustrated patients are highlighted.
P value in b: two-sided Wilcoxon rank-sum test. P value in c: two-sided binomial exact test.

Supplementary Fig. 13). To determine whether NRG1 promotes (Supplementary Fig. 10b). Interestingly, activation of Myc sig-
cellular resistance to EGFR inhibitors, we evaluated the prolifera- naling pathways conferred superior sensitivity to all ten EGFR
tive kinetics of EGFR-amplified PDCs in response to 50 ng/ml exog- inhibitors, while upregulation of KRAS and/or PI3K/AKT/mTOR
enous NRG1 protein. As suspected, NRG1 treatment induced drug signaling pathways was associated with resistance to a subset of
resistance to EGFR inhibitors compared with vehicle-treated cells EGFR-targeting drugs (Fig. 5e). Our results highlight a potential
(Supplementary Fig. 14a). Furthermore, silencing of NRG1 through drug evasion process where sustainable activation of RAS–MEK or
either shRNA-mediated knockdown or neutralizing antibody dem- PI3K–AKT signaling molecules could prevent cell death from EGFR
onstrated sensitizing effects to EGFR inhibitors, further support- inhibition in EGFR-driven tumors108,120. Therefore, we conducted a
ing our observation that NRG1 expression potentially dictates the two-drug combinational strategy of applying PAM inhibitors with
cellular response to EGFR inhibitors (Fig. 5d and Supplementary EGFR compounds in EGFR-altered tumors. Notably, combining
Fig. 14b,c). NRG1 has been reported to enhance the survival and PAM inhibitors significantly increased the sensitizing effects of
proliferation of cancer cells including GBM115,116, and elevated EGFR inhibitors (Fig. 5f and Supplementary Fig. 14d), highlighting
expression of NRG1 has been associated with increased sensitivity the potential combinatorial implementation of PAM inhibitors to
to EGFR inhibitors through activation of human epidermal growth overcome EGFR-mediated therapeutic resistance121–124.
factor receptor 3 (HER3) kinases in HER2 non-amplified cancer
cells117. However, our results support that NRG1 promotes cellu- Clinical feasibility of patient-centric drug screening-guided ther-
lar resistance to EGFR target therapy in a subpopulation of EGFR- apy. Pharmacogenomic mapping using PDCs has identified several
altered tumors, potentially due to triggering heterodimerization of genetic aberrations that may potentially guide biomarker-driven
EGFR and HER3, as previously reported118,119. clinical trials. As demonstrated in recent studies, most refractory can-
To identify cellular signaling pathways that influence pharma- cers have multiple genomic aberrations125, rending optimization of
cological sensitivity/resistance to EGFR target therapy, we per- genome-based treatment difficult for each patient. To translate our ex
formed single sample gene set enrichment analysis (ssGSEA) vivo analysis into a clinical framework, we compared PDC-based drug
between two previously identified major EGFR-altered subgroups sensitivities with clinical response in retrospective clinical studies.

Nature Genetics | www.nature.com/naturegenetics


NaTuRE GEnETICs Articles
EGFR inhibitor
s
Lapatinib
Alterations
Canertinib AC480 CO.1686
Afatinib Erlotinib
AEE788 Dacomitinib Mutation mRNA
Ibrutinib Gefitinib
Vismodegib Neratinib CNV Fusion
Olaparib BKM120
EGFR
PF.04449913 BYL719

PI
3K
Bortezomib BEZ235
Lineage-specific

/m
TO
BGJ398 PARP2 AZD2014
R

EGFRvIII PIK3R1 Colorectal

R in
FGF

AZD4547 XL147

hibit
MTOR
Gastric
Dovitinib PKI.587

ors
FGFR3 EGFR
NRG1
APC
ERBB2 Everolimus
Lung
LDK378

ABT.199 AZD5363 GBM


TP53

Ruxolitinib KRAS
FGFR2 Axitinib
PDGFRA
REST
Dabrafenib Regorafenib
HDAC4
Panobinostat Imatinib
FGFR3 KDR
SIK2
LGK.974 Tandutinib

IT
ARID1A

,K
Selumetinib PDGFRA Cediranib

FR
DG
Trametinib Vandetanib
ME

,P
K

Vemurafenib MET Sunitinib

FR
AUY922 CDK4

G
Pazopanib

VE
Dasatinib Tivozanib
Bosutinib Foretinib
Nilotinib INCB28060 Interactions
Bc LY2835219
Palbociclib
Crizotinib
Cabozantinib
r-A LEE011 Sensitive
b I K
, AL Resistant
CDK4/6 MET

Fig. 7 | Schematic illustration of the major lineage-specific and genomic associated drug interactions. Molecular targeting agents are clustered based on
drug family classification and connected to various genomic alterations, including mutation, mRNA expression, CNV, and fusion. Edges represent sensitive
or resistant gene–drug interactions. Lineage-specific drug associations are highlighted in each drug node based on specific lineage type.

To this end, we explored whether normalized AUCs (z-scores clinical outcome and drug screening results13,130,131. EGFR amplifica-
in a pan-cancer AUC dataset; n =​ 462) of drugs targeting EGFR, tion with concurrent EGFR mutations, including vIII, was only found
VEGFR, PI3K/mTOR, MET, or platelet-derived growth factor recep- in the right frontal tumor, and showed partial response to treatment
tor alpha could be used to predict the clinical response of 31 patients with afatinib (EGFR), while the left frontal tumor did not respond to
across 4 major cancer types, including gastric cancer (n =​  17), GBM this therapy (Supplementary Fig. 15c)13,113,132. These clinical results
(n =​ 8), lung adenocarcinoma (n =​ 5), and atypical meningioma are well correlated with the drug sensitivity profiles of the corre-
(n =​ 1). We discovered that the majority of the PDCs with a z-score sponding PDCs in response to afatinib. Another instance for drug-
of less than −​0.5 (indicating inclusion in the top approximately screening correlation with clinical sensitivities was demonstrated in
30% most sensitive to a particular drug) predicted a partial or com- a GBM patient (P10) harboring a PTEN mutation, who showed a
plete response with an average duration of 6 months (±​3  months) short-term partial response to everolimus (mTOR) treatment, while
(Fig. 6a)126. In concordance with the clinical outcome, the z-scores another GBM patient (P30) with a PTEN deletion did not respond to
for drugs derived from drug sensitivity analysis using patients’ prog- such therapeutic approach (Fig. 6d and Supplementary Fig. 15d,e)133.
eny PDCs demonstrated considerably lower values compared with The AUC of the P10-derived tumor cells treated with everolimus was
the values in clinically resistant tumors (Fig. 6b). Next, we evaluated notably lower compared with that of the P30-derived tumor cells,
the accuracy of the z-score-based drug sensitivity profile as a bio- highlighting the accuracy of our drug screening results in relevance
marker for predicting clinical response by applying receiver-oper- to clinical response. P215—a patient who had been diagnosed with
ating characteristic (ROC) curve analysis127. When we integrated atypical meningioma—showed remarkable in vitro sensitivity to
z-score and clinical response, our in vitro drug sensitivity of PDCs treatment with sunitinib (VEGFR, PDGFR, etc.), and accordingly
(n =​ 31) demonstrated high concordance with the clinical response presented a partial response on both meningeal and abdominal
(AUC =​  0.8125, P =​  0.003045; AUC  =​ 0.5 represents the chance dis- metastatic lesions (Fig. 6e and Supplementary Fig. 15f)134. HER2
crimination as a diagnostic accuracy)128. amplification is found in ~13–23% of gastric cancers, and phase
Here, we demonstrate examples for clinical concordance of drug III trials with anti-HER2-targeted agents have shown controversial
response prediction using the progeny PDCs’ drug sensitivity. Two results135–137. PDC progeny from a gastric cancer patient (P436) with
GBM patients had been treated with gefitinib (EGFR inhibitor) based HER2 amplification showed profound in vitro sensitivity to lapatinib
on the presence of EGFR amplification with concomitant point muta- (HER2 and EGFR), which was consistent with the clinical partial
tions, as well as PTEN deletion (Supplementary Fig. 15a,b)129. Despite response (Fig. 6f). Although larger-scale clinical studies are essen-
such promising genomic indications, neither patient responded to tial to determine the clinical utility of PDCs in predicting clinical
the drug treatment and their clinical outcomes were consistent with response, our findings suggest that integrative analysis of genome
our patient-centric drug sensitivity analysis. Furthermore, a multi- and PDC-based drug screening could be a potential tool for patient
centric GBM patient showed notable concordance between their enrichment trials.

Nature Genetics | www.nature.com/naturegenetics


Articles NaTuRE GEnETICs

This accumulating evidence collectively proposes clinical feasi- alternative avenue for predicting clinical response via integrative
bility of combined patient-centric drug sensitivity screening with pharmacological drug responses with genomic characterizations.
genomic profiling to facilitate the prediction of clinical response to As prospective validation, we are currently enrolling gastric cancer
targeted agents. patients and using PDC models to screen for drug sensitivity and to
facilitate optimized clinical trials based on integrative genome–drug
Discussion mapping analysis (clinicaltrials.gov identifier National Clinical Trial
Precision oncology aims to provide personalized treatment options number: 03170180).
by identifying and targeting genomic and molecular aberrations of Collectively, our systematic method, using a unique drug sen-
the individual patient tumor. This approach holds the potential to sitivity dataset of molecularly annotated patient derivatives, repre-
greatly improve clinical outcomes3. Rapidly accumulating genomic sents a significant conceptual advance toward precision oncology
data, spearheaded by TCGA and multiple global collaborative therapy. In addition, we have developed an interactive, publicly
efforts, including the International Cancer Genome Consortium, available web resource (the Cancer-Drug eXplorer (cDx); see URLs)
have painted a comprehensive portrait of the tumor genome for exploration of our pharmacogenomic dataset.
atlas1,8,138. Large-scale drug-screening efforts using human cancer URLs. Cancer-Drug eXplorer (cDx; temporary username and
cell line models have begun to establish a collection of gene–drug password: cdx), http://cancerdrugexplorer.org/; Picard, http://picard.
associations and have uncovered potential molecular markers pre- sourceforge.net; cBioPortal, http://www.cbioportal.org/study?id=​
dictive of therapeutic responses16,17,20,21,23. However, conventional gbm_tcga#summary; CCLE, https://portals.broadinstitute.org/ccle;
cell line models have failed to demonstrate accurate genomic rep- UCSC Xena, https://xenabrowser.net/datapages/; Sanger, http://
resentations of the parental tumors, which rarely lead to success- www.cancerrxgene.org/gdsc1000/GDSC1000_WebResources/
ful clinical implementation. Therefore, systematic evaluation and Home.html; Ayasdi, https://www.ayasdi.com/, https://github.com/
clinical application of patient-derived resources to assess the genetic MLWave/kepler-mapper; MSigDB, http://software.broadinstitute.
variations that underlie pharmacological drug responses could pro- org/gsea/msigdb.
vide breakthroughs for precision treatment in cancer patients.
In this study, we have demonstrated the feasibility and clini- Methods
cal relevance of a PDC-based drug screening system for pharma- Methods, including statements of data availability and any asso-
cogenomic analyses. We report the therapeutic landscape of 60 ciated accession codes and references, are available at https://doi.
molecular-targeted compounds on 462 PDCs across 14 cancer org/10.1038/s41588-018-0209-6.
types, identifying lineage-specific drug associations, such as gastric
cancers and PI3K inhibitors. We have also integrated cancer-driven Received: 14 June 2017; Accepted: 27 July 2018;
genomic variations, including somatic mutations, CNAs, and/or Published: xx xx xxxx
transcriptome expression. Through large-scale pharmacogenomic
analyses, we have suggested therapeutic options for KRAS-mutant References
tumors and identified molecular determinants, such as HDAC4 1. Brennan, C. W. et al. The somatic genomic landscape of glioblastoma. Cell
155, 462–477 (2013).
and SIK2, that dictate panobinostat sensitivity. HDAC4 (a class IIa 2. Weinstein, J. N. et al. The Cancer Genome Atlas Pan-Cancer analysis
HDAC) is preferentially expressed in the heart, brain, and muscles, project. Nat. Genet. 45, 1113–1120 (2013).
inferring a potential role in brain pathogenesis, including brain 3. Hamburg, M. A. & Collins, F. S. The path to personalized medicine.
tumors and acquired/intrinsic resistance to chemotherapies100. We N. Engl. J. Med. 363, 301–304 (2010).
4. Slamon, D. J. et al. Use of chemotherapy plus a monoclonal antibody
propose the potential functional role of HDAC4 and its regulator
against HER2 for metastatic breast cancer that overexpresses HER2.
SIK2 in inducing resistance to the HDAC inhibitor, the underlying N. Engl. J. Med. 344, 783–792 (2001).
mechanisms of which should be further evaluated. Furthermore, we 5. Chapman, P. B. et al. Improved survival with vemurafenib in melanoma
have identified molecular predictors of intrinsic resistance to EGFR with BRAF V600E mutation. N. Engl. J. Med. 364, 2507–2516 (2011).
inhibitors, such as NRG1, in EGFR-altered tumors. These results 6. O’Brien, S. G. et al. Imatinib compared with interferon and low-dose
cytarabine for newly diagnosed chronic-phase chronic myeloid leukemia.
have led us to propose the potential strategy of combining anti- N. Engl. J. Med. 348, 994–1004 (2003).
NRG1 therapy to overcome unresponsiveness to EGFR-targeted 7. Loeb, L. A. Human cancers express mutator phenotypes: origin,
treatment139. Moreover, we have suggested EGFR alterations, consequences and targeting. Nat. Rev. Cancer 11, 450–457 (2011).
including genomic amplification and structure variation (vIII) as a 8. Vogelstein, B. et al. Cancer genome landscapes. Science 339,
novel genomic biomarker for the BTK inhibitor, ibrutinib, in GBM 1546–1558 (2013).
9. Ceccarelli, M. et al. Molecular profiling reveals biologically discrete subsets
patients. Ibrutinib previously showed notable sensitivity to EGFR- and pathways of progression in diffuse glioma. Cell 164, 550–563 (2016).
mutant tumors, such as L858R, Del19 and T790M in non-small- 10. Rubio-Perez, C. et al. In silico prescription of anticancer drugs to cohorts of
cell lung cancer cells67,140, which are intracellular domain events and 28 tumor types reveals targeting opportunities. Cancer Cell 27, 382–396
critically different from the EGFR alterations that are often found (2015).
in GBM. As ibrutinib has been suggested to be able to penetrate the 11. Altman, R. B. Predicting cancer drug response: advancing the DREAM.
Cancer Discov. 5, 237–238 (2015).
blood–brain barrier141, our results provide important groundwork 12. Geeleher, P., Cox, N. J. & Huang, R. S. Clinical drug response can be
for the clinical feasibility of repurposing ibrutinib for EGFR-altered predicted using baseline gene expression levels and in vitro drug sensitivity
GBM patients. Collectively, our findings demonstrate significant lin- in cell lines. Genome Biol. 15, R47 (2014).
eage-specific or molecular correlates of diverse drug agents (Fig. 7). 13. Lee, J. K. et al. Spatiotemporal genomic architecture informs precision
Lastly, we have demonstrated clinical concordance between oncology in glioblastoma. Nat. Genet. 49, 594–599 (2017).
14. Burrell, R. A., McGranahan, N., Bartek, J. & Swanton, C. The causes and
PDC-based drug sensitivity and clinical response, highlighting consequences of genetic heterogeneity in cancer evolution. Nature 501,
promising opportunities for drug screening-guided personalized 338–345 (2013).
treatment. Although the evaluation of clinical sensitivity was based 15. Yates, L. R. et al. Subclonal diversification of primary breast cancer revealed
on early radiological tumor responses, initial response rates have by multiregion sequencing. Nat. Med. 21, 751–759 (2015).
largely been accepted to be significantly correlated with patient sur- 16. Garnett, M. J. et al. Systematic identification of genomic markers of drug
sensitivity in cancer cells. Nature 483, 570–575 (2012).
vival142,143. Furthermore, as previous clinical studies have demon- 17. Barretina, J. et al. The Cancer Cell Line Encyclopedia enables predictive
strated, predicting the therapeutic efficacy of a targeted compound modelling of anticancer drug sensitivity. Nature 483, 603–607 (2012).
based on genomic profiling alone is a difficult task as tumors often 18. Shoemaker, R. H. The NCI60 human tumour cell line anticancer drug
harbor multiple genetic aberrations144. Therefore, we provide an screen. Nat. Rev. Cancer 6, 813–823 (2006).

Nature Genetics | www.nature.com/naturegenetics


NaTuRE GEnETICs Articles
19. Basu, A. et al. An interactive resource to identify cancer genetic and lineage 51. Cloughesy, T. F., Cavenee, W. K. & Mischel, P. S. Glioblastoma: from
dependencies targeted by small molecules. Cell 154, 1151–1161 (2013). molecular pathology to targeted treatment. Annu. Rev. Pathol. 9, 1–25 (2014).
20. Holbeck, S. L., Collins, J. M. & Doroshow, J. H. Analysis of Food and Drug 52. Mellinghoff, I. K. et al. Molecular determinants of the response of
Administration-approved anticancer agents in the NCI60 panel of human glioblastomas to EGFR kinase inhibitors. N. Engl. J. Med. 353, 2012–2024
tumor cell lines. Mol. Cancer Ther. 9, 1451–1460 (2010). (2005).
21. Garnett, M. J. & McDermott, U. The evolving role of cancer cell line-based 53. Fallahi-Sichani, M., Honarnejad, S., Heiser, L. M., Gray, J. W. & Sorger, P. K.
screens to define the impact of cancer genomes on drug response. Curr. Metrics other than potency reveal systematic variation in responses to
Opin. Genet. Dev. 24, 114–119 (2014). cancer drugs. Nat. Chem. Biol. 9, 708–714 (2013).
22. Van de Wetering, M. et al. Prospective derivation of a living organoid 54. Jang, I. S., Neto, E. C., Guinney, J., Friend, S. H. & Margolin, A. A.
biobank of colorectal cancer patients. Cell 161, 933–945 (2015). Systematic assessment of analytical methods for drug sensitivity prediction
23. Iorio, F. et al. A landscape of pharmacogenomic interactions in cancer. Cell from cancer cell line data. Pac. Symp. Biocomput. 2014, 63–74 (2014).
166, 740–754 (2016). 55. Huang, S. & Pang, L. Comparing statistical methods for quantifying drug
24. Gao, H. et al. High-throughput screening using patient-derived tumor sensitivity based on in vitro dose-response assays. Assay Drug Dev. Technol.
xenografts to predict clinical trial drug response. Nat. Med. 21, 10, 88–96 (2012).
1318–1325 (2015). 56. Raub, T. J. et al. Brain exposure of two selective dual CDK4 and CDK6
25. Galli, R. et al. Isolation and characterization of tumorigenic, stem-like inhibitors and the antitumor activity of CDK4 and CDK6 inhibition in
neural precursors from human glioblastoma. Cancer Res. 64, 7011–7021 combination with temozolomide in an intracranial glioblastoma xenograft.
(2004). Drug Metab. Dispos. 43, 1360–1371 (2015).
26. Joo, K. M. et al. Patient-specific orthotopic glioblastoma xenograft models 57. Cen, L. et al. p16–Cdk4–Rb axis controls sensitivity to a cyclin-dependent
recapitulate the histopathology and biology of human glioblastomas in situ. kinase inhibitor PD0332991 in glioblastoma xenograft cells. Neuro. Oncol.
Cell Rep. 3, 260–273 (2013). 14, 870–881 (2012).
27. Lee, J. et al. Tumor stem cells derived from glioblastomas cultured in bFGF 58. Schroder, L. B. & McDonald, K. L. CDK4/6 inhibitor PD0332991 in
and EGF more closely mirror the phenotype and genotype of primary glioblastoma treatment: does it have a future? Front. Oncol. 5, 259 (2015).
tumors than do serum-cultured cell lines. Cancer Cell 9, 391–403 (2006). 59. Nicolau, M., Levine, A. J. & Carlsson, G. Topology based data analysis
28. Lee, J. Y. et al. Patient-derived cell models as preclinical tools for identifies a subgroup of breast cancers with a unique mutational profile and
genome-directed targeted therapy. Oncotarget 6, 25619–25630 (2015). excellent survival. Proc. Natl Acad. Sci. USA 108, 7265–7270 (2011).
29. Xie, Y. et al. The human glioblastoma cell culture resource: validated 60. Camara, P. G., Rosenbloom, D. I., Emmett, K. J., Levine, A. J. & Rabadan, R.
cell models representing all molecular subtypes. EBioMedicine 2, Topological data analysis generates high-resolution, genome-wide maps of
1351–1363 (2015). human recombination. Cell Syst. 3, 83–94 (2016).
30. Kanabur, P. et al. Patient-derived glioblastoma stem cells respond 61. Rizvi, A. H. et al. Applied topology delineates developmental progression
differentially to targeted therapies. Oncotarget 7, 86406–86419 (2016). with single-cell resolution. Nat. Biotech. (in the press).
31. Park, Y. H. et al. Role of HER2 mutations in refractory metastatic breast 62. Bhattacharya, B. et al. Pharmacologic synergy between dual
cancers: targeted sequencing results in patients with refractory breast phosphoinositide-3-kinase and mammalian target of rapamycin inhibition
cancer. Oncotarget 6, 32027–32038 (2015). and 5-fluorouracil in PIK3CA mutant gastric cancer cells. Cancer Biol. Ther.
32. Lim, S. H. et al. The implication of FLT3 amplification for FLT targeted 13, 34–42 (2012).
therapeutics in solid tumors. Oncotarget 8, 3237–3245 (2017). 63. Tapia, O. et al. The PI3K/AKT/mTOR pathway is activated in gastric cancer
33. Yoo, K. H. et al. Genomic alterations in biliary tract cancer using targeted with potential prognostic and predictive significance. Virchows Arch. 465,
sequencing. Transl. Oncol. 9, 173–178 (2016). 25–33 (2014).
34. Song, H. N. et al. Molecular characterization of colorectal cancer patients 64. Ying, J. et al. The expression of the PI3K/AKT/mTOR pathway in gastric
and concomitant patient-derived tumor cell establishment. Oncotarget 7, cancer and its role in gastric cancer prognosis. OncoTargets Ther. 8,
19610–19619 (2016). 2427–2433 (2015).
35. Suzuki, H. et al. Mutational landscape and clonal architecture in grade II 65. The Cancer Genome Atlas Research Network. Comprehensive molecular
and III gliomas. Nat. Genet. 47, 458–468 (2015). characterization of gastric adenocarcinoma. Nature 513, 202–209 (2014).
36. Trifonov, V., Pasqualucci, L., Tiacci, E., Falini, B. & Rabadan, R. SAVI: 66. Yiin, J. J. et al. ZD6474, a multitargeted inhibitor for receptor tyrosine
a statistical algorithm for variant frequency identification. BMC Syst. Biol. 7, kinases, suppresses growth of gliomas expressing an epidermal growth
S2 (2013). factor receptor mutant, EGFRvIII, in the brain. Mol. Cancer Ther. 9,
37. Magi, A. et al. EXCAVATOR: detecting copy number variants from 929–941 (2010).
whole-exome sequencing data. Genome Biol. 14, R120 (2013). 67. Gao, W. et al. Selective antitumor activity of ibrutinib in EGFR-mutant
38. Abate, F. et al. Pegasus: a comprehensive annotation and prediction tool for non-small cell lung cancer cells. J. Natl Cancer. Inst. 106, dju204 (2014).
detection of driver gene fusions in cancer. BMC Syst. Biol. 8, 97 (2014). 68. Byrd, J. C. et al. Targeting BTK with ibrutinib in relapsed chronic
39. Lemmon, M. A. & Schlessinger, J. Cell signaling by receptor tyrosine lymphocytic leukemia. N. Engl. J. Med. 369, 32–42 (2013).
kinases. Cell 141, 1117–1134 (2010). 69. Misale, S. et al. Emergence of KRAS mutations and acquired resistance to
40. Gschwind, A., Fischer, O. M. & Ullrich, A. The discovery of receptor tyrosine anti-EGFR therapy in colorectal cancer. Nature 486, 532–536 (2012).
kinases: targets for cancer therapy. Nat. Rev. Cancer 4, 361–370 (2004). 70. Garassino, M. C. et al. Different types of K-Ras mutations could affect drug
41. Nakada, M. et al. Aberrant signaling pathways in glioma. Cancers (Basel) 3, sensitivity and tumour behaviour in non-small-cell lung cancer. Ann. Oncol.
3242–3278 (2011). 22, 235–237 (2011).
42. Joo, K. M. et al. MET signaling regulates glioblastoma stem cells. Cancer 71. Lievre, A. et al. KRAS mutation status is predictive of response to
Res. 72, 3828–3838 (2012). cetuximab therapy in colorectal cancer. Cancer Res. 66, 3992–3995 (2006).
43. Wen, P. Y., Lee, E. Q., Reardon, D. A., Ligon, K. L. & Alfred Yung, W. K. 72. Belmont, P. J. et al. Resistance to dual blockade of the kinases PI3K and
Current clinical development of PI3K pathway inhibitors in glioblastoma. mTOR in KRAS-mutant colorectal cancer models results in combined
Neuro. Oncol. 14, 819–829 (2012). sensitivity to inhibition of the receptor tyrosine kinase EGFR. Sci. Signal. 7,
44. Filbin, M. G. et al. Coordinate activation of Shh and PI3K signaling in ra107 (2014).
PTEN-deficient glioblastoma: new therapeutic opportunities. Nat. Med. 19, 73. Hutchinson, L. Targeted therapies: dasatinib sensitizes KRAS-mutant
1518–1523 (2013). colorectal cancer tumors to cetuximab. Nat. Rev. Clin. Oncol. 8, 193 (2011).
45. Wen, P. Y. & Kesari, S. Malignant gliomas in adults. N. Engl. J. Med. 359, 74. Ku, B. M. et al. BYL719, a selective inhibitor of phosphoinositide 3-kinase
492–507 (2008). alpha, enhances the effect of selumetinib (AZD6244, ARRY-142886) in
46. Ohka, F., Natsume, A. & Wakabayashi, T. Current trends in targeted KRAS-mutant non-small cell lung cancer. Invest. New Drugs 33, 12–21
therapies for glioblastoma multiforme. Neurol. Res. Int. 2012, 878425 (2012). (2015).
47. Puputti, M. et al. Amplification of KIT, PDGFRA, VEGFR2, and EGFR in 75. Jing, J. et al. Comprehensive predictive biomarker analysis for MEK
gliomas. Mol. Cancer Res. 4, 927–934 (2006). inhibitor GSK1120212. Mol. Cancer Ther. 11, 720–729 (2012).
48. Taylor, T. E., Furnari, F. B. & Cavenee, W. K. Targeting EGFR for treatment 76. Infante, J. R. et al. Safety, pharmacokinetic, pharmacodynamic, and efficacy
of glioblastoma: molecular basis to overcome resistance. Curr. Cancer Drug data for the oral MEK inhibitor trametinib: a phase 1 dose-escalation trial.
Targets 12, 197–209 (2012). Lancet Oncol. 13, 773–781 (2012).
49. Snuderl, M. et al. Mosaic amplification of multiple receptor tyrosine kinase 77. Hatzivassiliou, G. et al. Mechanism of MEK inhibition determines efficacy
genes in glioblastoma. Cancer Cell. 20, 810–817 (2011). in mutant KRAS- versus BRAF-driven cancers. Nature 501, 232–236 (2013).
50. Szerlip, N. J. et al. Intratumoral heterogeneity of receptor tyrosine kinases 78. Blumenschein, G. R. Jr. et al. A randomized phase II study of the MEK1/
EGFR and PDGFRA amplification in glioblastoma defines subpopulations MEK2 inhibitor trametinib (GSK1120212) compared with docetaxel in
with distinct growth factor response. Proc. Natl Acad. Sci. USA 109, KRAS-mutant advanced non-small-cell lung cancer (NSCLC). Ann. Oncol.
3041–3046 (2012). 26, 894–901 (2015).

Nature Genetics | www.nature.com/naturegenetics


Articles NaTuRE GEnETICs
79. Manchado, E. et al. A combinatorial strategy for treating KRAS-mutant lung 108. Hopper-Borge, E. A. et al. Mechanisms of tumor resistance to EGFR-
cancer. Nature 534, 647–651 (2016). targeted therapies. Expert Opin. Ther. Targets 13, 339–362 (2009).
80. Yeh, J. J. et al. KRAS/BRAF mutation status and ERK1/2 activation as 109. Spaans, J. N. & Goss, G. D. Drug resistance to molecular targeted therapy
biomarkers for MEK1/2 inhibitor therapy in colorectal cancer. Mol. Cancer and its consequences for treatment decisions in non-small-cell lung cancer.
Ther. 8, 834–843 (2009). Front. Oncol. 4, 190 (2014).
81. Sun, C. et al. Intrinsic resistance to MEK inhibition in KRAS mutant lung 110. Fan, Q. W. et al. EGFR phosphorylates tumor-derived EGFRvIII driving
and colon cancer through transcriptional induction of ERBB3. Cell Rep. 7, STAT3/5 and progression in glioblastoma. Cancer Cell. 24, 438–449 (2013).
86–93 (2014). 111. Nathanson, D. A. et al. Targeted therapy resistance mediated by dynamic
82. Heinemann, V., Stintzing, S., Kirchner, T., Boeck, S. & Jung, A. Clinical regulation of extrachromosomal mutant EGFR DNA. Science 343,
relevance of EGFR- and KRAS-status in colorectal cancer patients treated 72–76 (2014).
with monoclonal antibodies directed against the EGFR. Cancer Treat. Rev. 112. Thiessen, B. et al. A phase I/II trial of GW572016 (lapatinib) in recurrent
35, 262–271 (2009). glioblastoma multiforme: clinical outcomes, pharmacokinetics and
83. Cui, J., Jiang, W., Wang, S., Wang, L. & Xie, K. Role of Wnt/beta-catenin molecular correlation. Cancer Chemother. Pharmacol. 65, 353–361 (2010).
signaling in drug resistance of pancreatic cancer. Curr. Pharm. Des. 18, 113. Reardon, D. A. et al. Phase I/randomized phase II study of afatinib, an
2464–2471 (2012). irreversible ErbB family blocker, with or without protracted temozolomide
84. Yeung, J. et al. Beta-catenin mediates the establishment and drug resistance in adults with recurrent glioblastoma. Neuro. Oncol. 17, 430–439 (2015).
of MLL leukemic stem cells. Cancer Cell 18, 606–618 (2010). 114. Uhm, J. H. et al. Phase II evaluation of gefitinib in patients with newly
85. Nagaraj, A. B. et al. Critical role of Wnt/beta-catenin signaling in diagnosed grade 4 astrocytoma: Mayo/North Central Cancer Treatment
driving epithelial ovarian cancer platinum resistance. Oncotarget 6, Group Study N0074. Int. J. Radiat. Oncol. Biol. Phys. 80, 347–353 (2011).
23720–23734 (2015). 115. Ritch, P. S., Carroll, S. L. & Sontheimer, H. Neuregulin-1 enhances survival
86. Ivanescu, A. M., Oprea, M., Turbatu, A., Colita, A. & Lupu, A. R. Ibrutinib, of human astrocytic glioma cells. Glia 51, 217–228 (2005).
a novel agent in relapsed or refractory chronic lymphocytic leukemia. 116. Sheng, Q. et al. An activated ErbB3/NRG1 autocrine loop supports in vivo
Maedica (Buchar) 9, 217–218 (2014). proliferation in ovarian cancer cells. Cancer Cell 17, 298–310 (2010).
87. Rushworth, S. A., MacEwan, D. J. & Bowles, K. M. Ibrutinib in relapsed 117. Wilson, T. R., Lee, D. Y., Berry, L., Shames, D. S. & Settleman, J.
chronic lymphocytic leukemia. N. Engl. J. Med. 369, 1277–1278 (2013). Neuregulin-1-mediated autocrine signaling underlies sensitivity to
88. Wang, M. L. et al. Targeting BTK with ibrutinib in relapsed or refractory HER2 kinase inhibitors in a subset of human cancers. Cancer Cell 20,
mantle-cell lymphoma. N. Engl. J. Med. 369, 507–516 (2013). 158–172 (2011).
89. Hu, Y. & Smyth, G. K. ELDA: extreme limiting dilution analysis for 118. Dong, X., Fernandez-Salas, E., Li, E. & Wang, S. Elucidation of resistance
comparing depleted and enriched populations in stem cell and other assays. mechanisms to second-generation ALK inhibitors alectinib and ceritinib in
J. Immunol. Methods 347, 70–78 (2009). non-small cell lung cancer cells. Neoplasia 18, 162–171 (2016).
90. Wang, J., Kribelbauer, J. & Rabadan, R. Network propagation reveals novel 119. Dhomen, N. S., Mariadason, J., Tebbutt, N. & Scott, A. M. Therapeutic
genetic features predicting drug response of cancer cell lines. Curr. targeting of the epidermal growth factor receptor in human cancer. Crit.
Bioinform. 11, 8 (2016). Rev. Oncog. 17, 31–50 (2012).
91. Munarini, N. et al. Altered mammary epithelial development, pattern 120. Dempke, W. C. & Heinemann, V. Ras mutational status is a biomarker for
formation and involution in transgenic mice expressing the EphB4 receptor resistance to EGFR inhibitors in colorectal carcinoma. Anticancer Res. 30,
tyrosine kinase. J. Cell Sci. 115, 25–37 (2002). 4673–4677 (2010).
92. Kumar, S. R. et al. Receptor tyrosine kinase EphB4 is a survival factor in 121. Fan, Q. W. et al. A dual phosphoinositide-3-kinase alpha/mTOR inhibitor
breast cancer. Am. J. Pathol. 169, 279–293 (2006). cooperates with blockade of epidermal growth factor receptor in
93. Yang, N. Y., Pasquale, E. B., Owen, L. B. & Ethell, I. M. The EphB4 PTEN-mutant glioma. Cancer Res. 67, 7960–7965 (2007).
receptor-tyrosine kinase promotes the migration of melanoma cells through 122. Yi, Y. W. et al. Inhibition of the PI3K/AKT pathway potentiates cytotoxicity
Rho-mediated actin cytoskeleton reorganization. J. Biol. Chem. 281, of EGFR kinase inhibitors in triple-negative breast cancer cells. J. Cell. Mol.
32574–32586 (2006). Med. 17, 648–656 (2013).
94. Ferguson, B. D. et al. The EphB4 receptor tyrosine kinase promotes lung 123. Tricker, E. M. et al. Combined EGFR/MEK inhibition prevents the
cancer growth: a potential novel therapeutic target. PLoS ONE 8, emergence of resistance in EGFR-mutant lung cancer. Cancer Discov. 5,
e67668 (2013). 960–971 (2015).
95. Pasquale, E. B. Eph receptors and ephrins in cancer: bidirectional signalling 124. Huang, M. H. et al. MEK inhibitors reverse resistance in epidermal growth
and beyond. Nat. Rev. Cancer 10, 165–180 (2010). factor receptor mutation lung cancer cells with acquired resistance to
96. Cai, Y., Yan, X., Zhang, G., Zhao, W. & Jiao, S. The predictive value of gefitinib. Mol. Oncol. 7, 112–120 (2013).
ERCC1 and p53 for the effect of panobinostat and cisplatin combination 125. Zehir, A. et al. Mutational landscape of metastatic cancer revealed from
treatment in NSCLC. Oncotarget 6, 18997–19005 (2015). prospective clinical sequencing of 10,000 patients. Nat. Med. 23,
97. Lee, E. Q. et al. Phase II study of panobinostat in combination with 703–713 (2017).
bevacizumab for recurrent glioblastoma and anaplastic glioma. Neuro. 126. Eisenhauer, E. A. et al. New response evaluation criteria in solid tumours:
Oncol. 17, 862–867 (2015). revised RECIST guideline (version 1.1). Eur. J. Cancer 45, 228–247 (2009).
98. Grasso, C. S. et al. Functionally defined therapeutic targets in diffuse 127. Soreide, K. Receiver-operating characteristic curve analysis in diagnostic,
intrinsic pontine glioma. Nat. Med. 21, 555–559 (2015). prognostic and predictive biomarker research. J. Clin. Pathol. 62,
99. Taylor, P. et al. REST is a novel prognostic factor and therapeutic target for 1–5 (2009).
medulloblastoma. Mol. Cancer Ther. 11, 1713–1723 (2012). 128. Hajian-Tilaki, K. Receiver operating characteristic (ROC) curve analysis
100. Wang, Z., Qin, G. & Zhao, T. C. HDAC4: mechanism of regulation and for medical diagnostic test evaluation. Caspian J. Intern. Med. 4,
biological functions. Epigenomics 6, 139–150 (2014). 627–635 (2013).
101. Plass, C. et al. Mutations in regulators of the epigenome and their 129. Rich, J. N. et al. Phase II trial of gefitinib in recurrent glioblastoma. J. Clin.
connections to global chromatin patterns in cancer. Nat. Rev. Genet. 14, Oncol. 22, 133–142 (2004).
765–780 (2013). 130. Lasocki, A., Gaillard, F., Tacey, M., Drummond, K. & Stuckey, S.
102. Kawata, H. et al. Zinc-fingers and homeoboxes (ZHX) 2, a novel member Multifocal and multicentric glioblastoma: improved characterisation
of the ZHX family, functions as a transcriptional repressor. Biochem. J. 373, with FLAIR imaging and prognostic implications. J. Clin. Neurosci. 31,
747–757 (2003). 92–98 (2016).
103. Walkinshaw, D. R. et al. The tumor suppressor kinase LKB1 activates the 131. Liu, Q. et al. Genetic, epigenetic, and molecular landscapes of multifocal
downstream kinases SIK2 and SIK3 to stimulate nuclear export of class IIa and multicentric glioblastoma. Acta Neuropathol. 130, 587–597 (2015).
histone deacetylases. J. Biol. Chem. 288, 9345–9362 (2013). 132. Alshami, J. et al. Afatinib, an irreversible ErbB family blocker, with
104. Geng, L. et al. Histone deacetylase (HDAC) inhibitor LBH589 increases protracted temozolomide in recurrent glioblastoma: a case report.
duration of gamma-H2AX foci and confines HDAC4 to the cytoplasm in Oncotarget 6, 34030–34037 (2015).
irradiated non-small cell lung cancer. Cancer Res. 66, 11298–11304 (2006). 133. Ma, D. J. et al. A phase II trial of everolimus, temozolomide, and
105. Geng, H. et al. HDAC4 protein regulates HIF1alpha protein lysine radiotherapy in patients with newly diagnosed glioblastoma: NCCTG
acetylation and cancer cell response to hypoxia. J. Biol. Chem. 286, N057K. Neuro. Oncol. 17, 1261–1269 (2015).
38095–38102 (2011). 134. Kaley, T. J. et al. Phase II trial of sunitinib for recurrent and progressive
106. Choi, M. C. et al. A direct HDAC4–MAP kinase crosstalk activates muscle atypical and anaplastic meningioma. Neuro. Oncol. 17, 116–121 (2015).
atrophy program. Mol. Cell 47, 122–132 (2012). 135. Bang, Y. J. et al. Trastuzumab in combination with chemotherapy versus
107. Ellis, L. M. & Hicklin, D. J. Resistance to targeted therapies: refining chemotherapy alone for treatment of HER2-positive advanced gastric or
anticancer therapy in the era of molecular oncology. Clin. Cancer Res. 15, gastro-oesophageal junction cancer (ToGA): a phase 3, open-label,
7471–7478 (2009). randomised controlled trial. Lancet 376, 687–697 (2010).

Nature Genetics | www.nature.com/naturegenetics


NaTuRE GEnETICs Articles
136. Hecht, J. R. et al. Lapatinib in combination with capecitabine plus and C6002-17G). The biospecimens for this study were provided by the Samsung
oxaliplatin in human epidermal growth factor receptor 2-positive advanced Medical Center BioBank.
or metastatic gastric, esophageal, or gastroesophageal adenocarcinoma:
TRIO-013/LOGiC—a randomized phase III trial. J. Clin. Oncol. 34,
443–451 (2016). Author contributions
J.-K.L., Z.L., J.K.S., S.S. and J.W. are co-first authors. J.-K.L., Z.L., J.K.S., S.S. and J.W.
137. Kim, S. T. et al. Impact of genomic alterations on lapatinib treatment
performed the majority of the experiments and analyses. Z.L. and M.B. analyzed the
outcome and cell-free genomic landscape during HER2 therapy in
therapeutic landscape of PDCs and pharmacogenomic interactions. D.I.S.R., O.E. and
HER2-positive gastric cancer patients. Ann. Oncol. 29, 1037–1048 (2018).
T.C. designed and constructed the cDx interactive webportal. S.W.C., D.-S.K., D.-H.N.,
138. The Cancer Genome Atlas Research Network.. Comprehensive genomic
S.T.K. and J.L. interpreted the clinical data. J.-K.L., S.S., J.-W.O., M.S., H.J.K., S.H.K.,
characterization defines human glioblastoma genes and core pathways.
G.H.R. and Y.-J.K. organized and analyzed the drug-screening experiments. Y.J.Shin,
Nature 455, 1061–1068 (2008).
H.J.K., Y.J.Seo, M.L., S.Y.K., M.-H.S., J.K., T.L., S.-Y.S., K.-M.K., M.K., J.O.P. and Y.Y.
139. Fernandez-Cuesta, L. & Thomas, R. K. Molecular pathways: targeting NRG1
organized and processed the specimens for patient-derived cultures and genome analysis.
fusions in lung cancer. Clin. Cancer Res. 21, 1989–1994 (2015).
D.K. and M.L. conducted the animal experiments. J.K.S., H.J.C., I.-H.L., H.S., N.K.D.K.,
140. Wu, H. et al. Ibrutinib selectively and irreversibly targets EGFR (L858R,
J.S.B. and W.-Y.P. analyzed the genomic profiling. D.-S.K., J.W.C., H.J.S., J.-I.L., J.-W.L.,
Del19) mutant but is moderately resistant to EGFR (T790M) mutant
H.-C.K., J.E.L., M.G.C., S.W.S., Y.M.S., J.I.Z. and B.C.J. provided surgical specimens.
NSCLC cells. Oncotarget 6, 31313–31322 (2015).
J.-K.L., Z.L., J.K.S., S.S. and J.W. wrote the manuscript with the feedback from J.L.,
141. Bernard, S. et al. Activity of ibrutinib in mantle cell lymphoma patients
R.G.W.V., A.I., J.L., R.R. and D.-H.N. J.L., R.R. and D.-H.N. designed and supervised the
with central nervous system relapse. Blood 126, 1695–1698 (2015).
entire project.
142. Jain, P. et al. Early responses predict better outcomes in patients with newly
diagnosed chronic myeloid leukemia: results with four tyrosine kinase
inhibitor modalities. Blood 121, 4867–4874 (2013). Competing interests
143. Louvet, C. et al. Correlation between progression free survival and response The authors declare no competing interests.
rate in patients with metastatic colorectal carcinoma. Cancer 91,
2033–2038 (2001).
144. Tsimberidou, A. M. & Kurzrock, R. Precision medicine: lessons learned Additional information
from the SHIVA trial. Lancet Oncol. 16, e579–e580 (2015). Supplementary information is available for this paper at https://doi.org/10.1038/
s41588-018-0209-6.
Acknowledgements Reprints and permissions information is available at www.nature.com/reprints.
This research was supported by a grant of the Korea Health Technology Research Correspondence and requests for materials should be addressed to J.L. or R.R.
and Development project through the Korea Health Industry Development or D.-H.N.
Institute (KHIDI), funded by the Ministry of Health and Welfare, Republic of Korea
Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in
(HI14C3418). This work has been funded by NIH grants (R01 CA185486, R01
published maps and institutional affiliations.
CA179044, U54 CA193313 and U54 209997) and NSF/SU2C/V Foundation Ideas Lab
Multidisciplinary Team (PHY-1545805) and Hong Kong RGC grants (N_HKUST601/17 © The Author(s), under exclusive licence to Springer Nature America, Inc. 2018

Nature Genetics | www.nature.com/naturegenetics


Articles NaTuRE GEnETICs

Methods accessed using an adenosine triphosphate monitoring system based on firefly


Tumor specimens and their derivative cells. After receiving informed consent, luciferase (ATPlite 1step; PerkinElmer) and estimated by EnVision Multilabel
tumor specimens or malignant ascites with corresponding clinical records were Reader (PerkinElmer)13. The relative cell viability for each dose was obtained
obtained from patients undergoing surgery or paracentesis at the Samsung by normalization with dimethyl sulfoxide (DMSO) per each plate. Screening
Medical Center (SMC) in accordance with its Institutional Review Board (IRB plates were subjected to quality control measurement using z-factor scores,
file numbers 201004004, 201107089, 201310017, and 201310072). This work was comparing both negative and positive control wells150. DRCs were fitted using
performed in compliance with all relevant ethical regulations for research using GraphPad Prism 6 (GraphPad): best-fit lines and the resulting IC50 values were
human specimens. Cells from malignant effusions were collected by centrifugation calculated using GraphPad: (log[inhibitor] versus response – variable slope (four
at 300 g for 10 min, followed by washing with Dulbecco’s phosphate-buffered parameters)). The equation for modeling log (inhibitor) vs. response curve is
saline (Thermo Scientific). Portions of the surgical samples were enzymatically Y =​  Bottom  +​  (Top  −​ Bottom) / (1 +​ 10(log[IC50 −x])× HillSlope)). Top and Bottom are
dissociated using Liberase TM (Roche). PDCs were cultured in neurobasal plateaus in the units of the y axis and IC50 is the concentration of drug that provides
or DMEM/F12 medium with N2 and B27 supplements (0.5×​each; Thermo a response halfway between Bottom and Top. HillSlope describes the steepness of
Scientific) and growth factors based on human recombinant basic fibroblast growth the curves. The AUC for each DRC was calculated using the trapezoidal method,
factor and epidermal growth factor (EGF; 20 ng ml−1 each; R&D Systems)25,27. ignoring regions defined by fewer than two peaks55. The AUC values from non-
Short tandem repeat analysis (AmpFlSTR Identifiler; Applied Biosystems) was convergent or ambiguous DRCs were excluded in all analyses.
performed to verify the corresponding normal blood. A Universal Mycoplasma
Detection Kit (9ATCC 30-1012K) was used to confirm the absence of Mycoplasma Transcriptome expression comparison between primary tumors and patient-
contamination. derived cells. We selected 24 GBM samples with available RNA-Seq data on both
primary tissues and matching PDCs. In addition to the normalization described
WES or targeted exome sequencing. An Agilent SureSelect kit was used to above, we further calculated z-scores for each gene within the tissue/PDC
capture the exonic DNA fragments. An Illumina HiSeq 2000 instrument was used cohort to eliminate batch effects. Then, for each gene, we applied a Spearman’s
for sequencing and generation of 2×​100-base pair paired-end reads. FASTQ files correlation as a score to evaluate its expression similarity. Based on its cut-offs, we
were aligned to the human genome assembly (hg19) using the Burrows–Wheeler generated different lists of genes for comparison (Supplementary Fig. 4). Finally,
Aligner (version 0.6.2)145. Before further analysis, the initially aligned BAM files we calculated pairwise Spearman’s correlation coefficients between the normalized
were subjected to preprocessing that sorted, removed duplicated reads, locally expression values obtained for primary tissues and PDCs.
realigned reads around potential small indels, and recalibrated base quality scores
using SAMtools, Picard (version 1) and the Genome Analysis ToolKit (GATK, Molecular profile comparisons among PDCs, tumor tissue, and conventional
version 2.5.2)146. In addition to WES, we also performed targeted sequencing cancer cell lines on glioma samples. To evaluate whether PDCs could provide a
of full coding exons of 80 commonly mutated cancer genes (CancerSCAN; better representation of primary tumor tissue, we compared recurrent mutation
Supplementary Tables 1 and 2)31,33. Tissues or PDCs from malignant gliomas were and gene expression profiles of our PDC system, TCGA, and two publicly available
analyzed with a massive parallel targeted sequencing platform, covering exons pharmacogenomics resources using traditional cancer cell line models: the CCLE
of cancer-driven and/or glioma-associated genes (GliomaSCAN; Supplementary and Genomics of Drug Sensitivity in Cancer datasets. The comparison was
Tables 1 and 3)1,9,35. Read alignments and conventional preprocessing were conducted on GBM samples only.
conducted similar to the WES data. For somatic mutation profiles, 33 GBM PDCs with available GliomaScan data
and normal controls were selected to call somatic mutations. Mutation reports
Mutation calling. To identify somatic mutations from WES data for tumor samples on the TCGA GBM cohort (n =​ 287) were downloaded from cBioPortal (see
with matched blood control, we applied the variance-calling software SAVI2, based URLs). For GDSC, mutation information on GBM cell lines (n =​ 34) was obtained
on the empirical Bayesian method36. Somatic mutations were identified based on from the supplementary table of Iorio et al.23. Binary calls for mutation data were
the report of SAVI2, and following four additional criteria: (1) not annotated as a downloaded from the CCLE website (see URLs), and 68 GBM cell lines were
synonymous variant, intragenic variant, or intron variant; (2) not annotated as a extracted. We selected a list of 27 marker genes, which had a mutation occurrence
common SNP; (3) a mutation allele frequency of >​5% in the tumor sample; and of >​5 patients on TCGA’s report and were also covered by the protocol of
(4) an altered read depth of <​2 in the matched normal control. For targeted DNA GliomaScan. Next, we calculated and compared the mutation frequency of selected
sequencing, we only focused on variant callings within the genomic region covered genes in each cohort.
by the protocol. For samples without a matching normal control, only the variants To compare the transcriptome similarities, 24 GBM PDCs with available RNA-
with cosmic identification were considered. Seq data were selected. For GBM samples from TCGA, we used the pan-cancer
normalized gene expression data (n =​ 172) downloaded from the University of
CNA. Excavator was used to generate estimated CNAs in a tumor specimen California Santa Cruz (UCSC) Xena (see URLs). Normalized expression data
compared with its matching blood. For each gene, we calculated the copy number of 35 GBM cell lines were downloaded from the Sanger Institute’s website (see
as 2x+1, where x is the segmentation mean, which is the log2 (fold-change) value URLs). For CCLE, expression data of 69 GBM cell lines were downloaded from the
for each tumor divided by the normal control. The gene was labeled as ‘amplified’ CCLE website (see URLs). Then, expression matrixes of four cohorts were merged
when the x value was above 0 and ‘deleted’ when it was below 0. The copy-number on overlapped genes, and the values were quantile-normalized at the sample
estimation was only conducted on WES data. level. Next, to evaluate the transcriptome similarities of two different cohorts,
we calculated the Spearman’s correlations of mRNA expression for every pair of
RNA-Seq. Reads were aligned to the human reference using STAR 2.4.0b147. samples coming from the two different cohorts. Lastly, P values were generated
The genome index was generated using an annotation file of the human genome from Kruskal–Wallis tests to estimate the statistical differences of the correlation
(GRCh37), version 19 (Ensembl 74) from GENCODE. For gene expression distributions among multiple comparisons.
analysis, featureCounts from the package ‘Subread’ was adopted to calculate RPKM
values, followed by log2 transformation, and quantile normalization. Genes with Lineage-specific drug identification. Drug-response profiles were directly
low expression levels across the cohort were removed. compared between a given tissue type and all other PDCs. The Wilcoxon rank-sum
test was adopted to show the statistical significance, followed by the Benjamini–
Exon skipping and gene fusion analysis. Chimerascan was used to generate a Hochberg procedure for multiple tests correction. We also applied TDA for the
list of candidate gene fusions using RNA-Seq data148. To reduce the false positive identification of tissue-specific drug sensitivity.
rate and nominate potential driver events, we applied the Pegasus annotation
and prediction pipeline38. Only previously reported in-frame, highly expressed TDA. Topological representations were constructed using the ‘Mapper’ algorithm,
fusions are reported in this manuscript. To examine the rearrangement of EGFR, as implemented by Ayasdi (see URLs). Open-source implementations of this
we applied prada-guess-if from the PRADA package. PRADA is an RNA-Seq algorithm are also available. The output of ‘Mapper’ is a low-dimensional network
analysis pipeline developed at MD Anderson149. The transcribed allelic fraction of representation of the data, where nodes represent sets of drugs or PDCs. In the
EGFRvIII was defined as the fraction of junction reads joining exon 1 to exon 8. manuscript, we have constructed three topological networks, as detailed below.
A list of major alterations, including copy-number changes, point mutations, The TDA network of 462 PDCs was constructed using Pearson’s correlation
and fusion genes is provided in Supplementary Table 2. as a similarity measure between AUC profiles across 60 drugs. As ‘Mapper’
parameters, we used two-dimensional scaling projections as lenses with a
PDC-based chemical screening and analysis. Tumorsphere-forming PDCs, resolution of 30 and a gain of 3 (equalized). Each node of the network represents
cultured in serum-free medium, were dissociated into single cells and seeded into a set of PDCs with similar AUC profiles. A PDC can appear in several nodes, and
384-well plates (500 cells well−1) with technical duplicates. PDCs were treated with two nodes are connected by an edge if they have at least one PDC in common.
a 60-drug library, targeting major oncogenic signaling molecules (SelleckChem), We then computed mean adjusted AUCs for each node, and for each drug: AUCs
in a fourfold and seven-point serial dilution series from 4.88 nM to 20 μ​M. Two larger than 300 for a PDC were taken to be 300, and AUCs smaller than 200
of the same PDCs and two of the same cell lines were screened every month to were taken to be 0 to emphasize strong drug sensitivity. For each node, we also
confirm the preservation of chemical activities of our drug library. After 6 days computed the fraction of PDCs derived from a particular cancer type. The mean
of incubation at 37 °C in a 5% CO2 humidified incubator, cell viability was adjusted AUC and cancer type fraction were then used both to color nodes and to

Nature Genetics | www.nature.com/naturegenetics


NaTuRE GEnETICs Articles
compute the Pearson’s correlation between the drug effect and cancer type. P values that indicates the degree of propagation of information flow, and H is the negative
of these correlations are used to identify significant relationships between cancer Laplacian matrix defined on the adjacency matrix from the STRING network. We
type and drug effect. only selected the top 500 closest genes as candidate features. The hypothesis is that
The TDA network of 44 EGFR-altered PDCs was constructed based on the correlation-based sensitivity features might be close to the known targets in terms
Euclidean (L2) metric as a similarity measure between gene expression profiles. of sharing a common pathway or carrying out the same function.
Multidimensional scaling projection lenses with a resolution of 12 and a gain of 4 Notably, one parameter, γ, was involved in the above diffusion kernel process.
(equalized) were used as ‘Mapper’ parameters. Each node represents a set of EGFR- Next, we applied an elastic net regression algorithm154 using only mRNA expression
altered PDCs with similar gene expression profiles. A PDC can appear in several of genes filtered by the network propagation. Then, tenfold cross-validations were
nodes, and two nodes are connected by an edge if they have at least one PDC in used to optimize the parameter γ. The final list of gene expression features was
common. We then computed mean AUCs for each drug and mean gene expression determined by a diffusion kernel process with optimized γ. γ was screened using 30
for every gene in every node of the network. The mean AUC and gene expression values of γ =​  0.75n with n =​  1, 2, …​, 30.
were then used both to color nodes and to compute the Pearson’s correlation Next, we trained the standard elastic net regression (R package: glmnet_2.0-5),
between drug effect and gene expression. P values of these correlations are used to combing all alterations, tumor subtyping, and pre-selected mRNA expression
identify significant relationships between gene expression and drug effect. as the input features. Given one drug and n PDCs, we used y =​  (y1, y2, …​ yn)
The TDA network of 60 drugs was constructed based on the Pearson’s to represent the drug sensitivity of n PDCs, and xi =​  (xi1,…​, xip), (i =​  1,…​, p) to
correlation as a similarity measure between their AUC profiles across 462 PDCs. indicate p features of ith PDC; the elastic net regression is to solve the following
Two-dimensional scaling projection lenses with a resolution of 15 and a gain of 3 optimization problem:
(equalized) were used as ‘Mapper’ algorithm parameters. Each node represents a
set of drugs with similar AUC profiles. A drug can appear in several nodes, and  n  p  2 
 1   1 
two nodes are connected by an edge if they have at least one drug in common. min β , β 
0  2n
∑ yi −β0− ∑ (xijβj)  + λ (1−α) βl2 + αβl 
 2 2 1
Every drug family is assigned a specific color. Mean RGB values are computed for  i =1  j =1  
every node to identify the drug family that dominates the node.  

α controls the relative balance of the L1 and L2 penalty terms, and β is the fitting
Oncogenic signaling pathways activation assessment. To determine the relative
coefficients to be estimated, while λ controls the overall penalty level of the
activity of cancer pathways between two cohorts, ssGSEA (version gsea2-2.2.1)
regularized term. Similar to γ, α was optimized by tenfold cross-validations, but
was applied using patients’ gene expression profiles. To eliminate batch effects, we
using all features. α was screened using 50 values with α ∈​[0,1] equally spaced.
normalized gene expression by calculating the z-score within each cohort. Then,
For each model fitting, we used the function cv.glmnet, with its optimized λ value
for each sample, we ranked all genes on the basis of their expression values to
provided from the function.
create a .rnk file as input for the software GSEAPreranked. The enrichment score
After parameter optimization, we adopted a bootstrapping strategy for 100
was computed for oncogenic pathways as the assessment. Cancer-related pathway
times to obtain a robust evaluation of the predictive power of features. During each
signatures were downloaded from the Molecular Signatures Database (MSigDB;
bootstrapping, we randomly selected 80% of the PDCs with 80% of the features to
see URLs). The PAM pathway was compared between gastric cancers and
fit the elastic net with above-optimized α. For each feature, the time of appearances
glioblastomas of TCGA patients, using the pan-cancer normalized gene expression
(non-zero fitting coefficient) out of the 100 bootstrappings, together with
data downloaded from UCSC Xena. We randomly selected 100 TCGA patients
the average of its non-zero weights, were used as its final assessment of
from each cohort for ssGSEA. We also compared the activation of signaling
predictive ability.
pathways between two EGFR-altered GBM subgroups, which demonstrated two
drastic responses to EGFR therapy. First, we selected 49 GBM PDCs with EGFR
Limiting dilution assay. Single-cell suspensions were plated into 96-well plates
alterations. Then, for each EGFR inhibitor, we ranked the drug AUC values for
at 1–250 cells per well. Cells were incubated at 37 °C for one to two weeks under
these 49 PDCs, and selected the top and bottom ten ranked PDCs as the resistance
serum-free conditions supplemented with EGF and basic firbroblast growth factor
and sensitivity subgroups, respectively, for comparison.
(20 ng ml−1 each), with or without drug treatment. Each well was examined for the
formation of tumor spheres. Statistical significance was accessed using extreme
Pan-cancer analyses uncover statistically significant associations between drug
limiting dilution analysis (Walter and Eliza Hall Bioinformatics)89.
sensitivity and genomic alteration. A pre-selected list of major cancer-driver
alterations, including copy-number changes, somatic mutations, and gene fusions,
shRNA-mediated knockdown. shRNA lentiviral clones for SIK2
was considered to identify drug-response interactions (Supplementary Table 2).
(TRCN0000037495 and TRCN0000037497 for clones 1 and 2, respectively),
For each drug, sensitivity data were compared between pan-cancer subgroups
HDAC4 (TRCN0000004832, TRCN0000314667, and TRCN0000314665 for clones
based on the presence or absence of the selected genomic alteration using a
1, 2, and 3, respectively) and NRG1 (TRCN0000058303) were purchased from
Wilcoxon rank-sum test. Samples with unknown status of given alteration were
Sigma–Aldrich. Lentiviruses were produced from 293FT cells with packaging mix
excluded from the analysis.
(ViraPower Lentiviral Expression Systems; Thermo Fisher) and concentrated by
ultracentrifugation. Lentiviral particles were transduced into corresponding tumor
Pharmacogenomic modeling of drug sensitivity. To predict drug sensitivity
cells and subjected for puromycin selection.
profiles based on cooperative interactions among multiple layers of variables, we
applied the dNetFS—a regression model-based algorithm (Supplementary Fig. 14)90.
PDX animal models and drug treatment. All animal experiments were approved
Generally, our dNetFS pipeline consists of two parts: (1) network propagation
by the Institutional Review Board of the SMC and performed according to the
to select potential predictive features with a strong network connection to known
guidelines of the Animal Use and Care Committees at SMC. A GBM orthotopic
drug targets; and (2) fitting an elastic-net model to generate a robust list of
xenograft model was established as reported previously26. In brief, patient GBM
molecular predictors.
cells (N464; 2 ×​ 105 per mouse) were dissociated, resuspended in 5 μ​l of Hank’s
The input variables of the dNetFS include the pre-selected list of cancer-
balanced salt solution (Thermo Scientific), and stereotactically (2 mm left and
driver alterations (mutations, CNVs, and fusions), tumor subtyping and mRNA
1 mm anterior of the bregma; 2 mm deep from the dura) injected into the brains
expression. Here, we only employed the expression of 7,262 genes, which showed
of BALB/c nude mice (~6–8-week-old females; Orient Bio). After randomization,
high consistency between primary tissue and matched PDCs with a correlation
vehicle (0.5% methyl cellulose; Sigma–Aldrich) or ibrutinib (50 mg kg−1 day−1;
of greater than 0.4 (Supplementary Fig. 5). However, this number is still much
Chemietek) were treated orally for five consecutive days per cycle155. Kaplan–Meier
larger than the total feature number of driver alterations, leading to unbalanced
survival and immunohistochemical analysis were performed under blinded
feature pools. Also, it falls into the standard sparsely selection problem, where the
inspection.
number of features is far greater than the number of observations. Thus, before any
regression model-based shrinkage, the dNetFS performed a pre-selection on the
Immunohistochemistry. Paraformaldehyde-fixed tissues were embedded in
expression features based on a protein–protein interaction network. We adopted
paraffin and sectioned to a thickness of 4 μ​m. Sections from the brains of xenograft
the STRING human network (version 10.0) and only kept the most confident 10%
tumor-bearing mice were stained with hematoxylin and eosin (Sigma–Aldrich).
of interactions151. For a given drug, we first obtained the known drug-targeted genes
Paraffin-embedded slides were subjected to antigen retrieval in 10 mM sodium
from the Drug Gene Interaction Database (DGIdb), which is a comprehensive
citrate buffer. For immunohistochemistry, tissue sections were incubated with
collection of available information on the ‘actionable’ genome152. Then, a diffusion
primary antibodies against pEGFR (2234), pAKT (4060), and pERK (9102; all from
kernel was applied as a measure to capture information flow on the network to
Cell Signaling Technology), followed by the appropriate biotinylated secondary
evaluate the closeness between known drug targets and other proteins153:
antibody. Sections were then stained with 3,3-diaminobenzidine substrate and
γ2 2 counterstained with hematoxylin and eosin solution.
K = e γH = I + γH + H +…
2! Radiological determination of tumor responses. All complete and partial
responses were confirmed by repeated radiological evaluation, including computed
This represents the continuous time limit of a lazy random walk, where K is the
tomography scanning or magnetic resonance imaging per response evaluation
diffusion kernel, e is the nature base, I is the identity matrix, γ is the one parameter
criteria in solid tumors (RECIST) guideline version 1.1 (ref. 126).

Nature Genetics | www.nature.com/naturegenetics


Articles NaTuRE GEnETICs
Statistical analysis. Statistical analyses were conducted by either Pearson’s 147. Dobin, A. et al. STAR: ultrafast universal RNA-Seq aligner. Bioinformatics
correlation coefficient test, t-test (two-tailed), Wilcoxon rank-sum test (two- 29, (15–21 (2013).
sided), or binomial exact test (two-sided). Survival curves were estimated with the 148. Iyer, M. K., Chinnaiyan, A. M. & Maher, C. A. ChimeraScan: a tool for
Kaplan–Meier method. For in vitro results, two to three independent experiments identifying chimeric transcription in sequencing data. Bioinformatics 27,
were conducted. Results are expressed as means ±​ s.d. for the indicated number of 2903–2904 (2011).
observations. All statistical analyses were conducted and obtained using GraphPad 149. Torres-Garcia, W. et al. PRADA: pipeline for RNA sequencing data analysis.
Prism software 6 or the R software (https://www.r-project.org). Bioinformatics 30, 2224–2226 (2014).
150. Zhang, J. H., Chung, T. D. & Oldenburg, K. R. A simple statistical
Reporting Summary. Further information on research design is available in the parameter for use in evaluation and validation of high throughput
Nature Research Reporting Summary linked to this article. screening assays. J. Biomol. Screen. 4, 67–73 (1999).
151. Franceschini, A. et al. STRINGv9.1: protein–protein interaction networks,
Data availability with increased coverage and integration. Nucleic Acids Res. 41,
All sequenced data have been deposited in the European Genome-phenome D808–D815 (2013).
Archive (EGA) under accession EGAS00001002515. Processed data and basic 152. Griffith, M. et al. DGIdb: mining the druggable genome. Nat. Methods 10,
association analysis are publicly available through an interactive web portal (the 1209–1210 (2013).
Cancer-Drug eXplorer (cDx); see URLs). 153. Kondor, R. I. & Lafferty, J. Diffusion kernels on graphs and other discrete
structures. In Proc. 19th International Conference on Machine Learning 8
References (Morgan Kaufmann, 2002).
145. Baras, A., Yu, Y., Filtz, M., Kim, B. & Moskaluk, C. A. Combined genomic 154. Zou, H. & Hastie, T. Regularization and variable selection via the elastic
and gene expression microarray profiling identifies ECOP as an upregulated net. J. R. Stat. Soc. B 67, 20 (2005).
gene in squamous cell carcinomas independent of DNA amplification. 155. Honigberg, L. A. et al. The Bruton tyrosine kinase inhibitor PCI-32765
Oncogene 28, 2919–2924 (2009). blocks B-cell activation and is efficacious in models of autoimmune
146. DePristo, M. A. et al. A framework for variation discovery and genotyping disease and B-cell malignancy. Proc. Natl Acad. Sci. USA 107,
using next-generation DNA sequencing data. Nat. Genet. 43, 491–498 (2011). 13075–13080 (2010).

Nature Genetics | www.nature.com/naturegenetics


nature research | life sciences reporting summary
Corresponding Author: Do-Hyun Nam, Raul Rabadan, Jeeyun Lee

Date: June 7, 2018

Life Sciences Reporting Summary


Nature Research wishes to improve the reproducibility of the work we publish. This form is published with all life science papers and is intended to
promote consistency and transparency in reporting. All life sciences submissions use this form; while some list items might not apply to an individual
manuscript, all fields must be completed for clarity.
For further information on the points included in this form, see Reporting Life Sciences Research. For further information on Nature Research policies,
including our data availability policy, see Authors & Referees and the Editorial Policy Checklist.

` Experimental design
1. Sample size
Describe how sample size was determined. The sample size of pharmacological data from 462 patient-derived cells
(PDCs) were as large as that from Cancer Cell Line Encyclopedia (CCLE,
n=479), which were proven to be sufficient to extract statistically
significant lineage-specific and gene-drug associations.
2. Data exclusions
Describe any data exclusions. Drug sensitivity data (Area Under Curve, AUC) from non-converged or non-
fitted dose-response curve according to the 4 parameter logistics were
excluded.
3. Replication
Describe whether the experimental findings were reliably reproduced. All drug screening data were obtained from technical replicates and
reliably reproduced.
4. Randomization
Describe how samples/organisms/participants were allocated into Tumor specimens had been randomly collected from patients with
experimental groups. informed consents, and there has been no intended bias of allocation into
experimental design. For in vivo experiments, mice were allocated to each
group based on weight to ensure that all groups were within similar weight
distribution.
5. Blinding
Describe whether the investigators were blinded to group allocation The investigators and authors have been consistently blinded to the group
during data collection and/or analysis. allocation during data collection and/or analysis.
Note: all studies involving animals and/or human research participants must disclose whether blinding and randomization were used.

June 2017

1
6. Statistical parameters

nature research | life sciences reporting summary


For all figures and tables that use statistical methods, confirm that the following items are present in relevant figure legends (or the Methods
section if additional space is needed).

n/a Confirmed

The exact sample size (n) for each experimental group/condition, given as a discrete number and unit of measurement (animals, litters, cultures, etc.)
A description of how samples were collected, noting whether measurements were taken from distinct samples or whether the same sample
was measured repeatedly.

A statement indicating how many times each experiment was replicated

The statistical test(s) used and whether they are one- or two-sided (note: only common tests should be described solely by name; more
complex techniques should be described in the Methods section)
A description of any assumptions or corrections, such as an adjustment for multiple comparisons

The test results (e.g. p values) given as exact values whenever possible and with confidence intervals noted

A summary of the descriptive statistics, including central tendency (e.g. median, mean) and variation (e.g. standard deviation, interquartile range)

Clearly defined error bars

See the web collection on statistics for biologists for further resources and guidance.

` Software
Policy information about availability of computer code
7. Software
Describe the software used to analyze the data in this study. R version 3.2.1, python 2.7.11, Matlab 2014b, Ayasdi 7.3.0.2, Burrows-
Wheeler Aligner 0.6.2, Picard 1, Genome Analysis ToolKit 2.5.2, SAVI2,
SAMtools 0.1.18, Excavator 2.2, STAR 2.4.0b, Chimerascan v0.4.3, Pegasus,
PRADA 1.1, GraphPad Prism 6, ssGSEA2-2.2.1, dNetFS 1.0, STRING 10.0,
ELDA
For all studies, we encourage code deposition in a community repository (e.g. GitHub). Authors must make computer code available to editors and reviewers upon
request. The Nature Methods guidance for providing algorithms and software for publication may be useful for any submission.

` Materials and reagents


Policy information about availability of materials
8. Materials availability
Indicate whether there are restrictions on availability of unique The patient-derived cell/mouse models were restricted on availability, and
materials or if these materials are only available for distribution by a can be distributed only with appropriate material transfer agreements.
for-profit company.
9. Antibodies
Describe the antibodies used and how they were validated for use in Immunoblot: p-EGFR 1:1000 (Cell Signaling, #3777), EGFR 1:1000 (abcam,
the system under study (i.e. assay and species). ab52894), p-STAT3 1:1000 (Cell Signaling, #9145), STAT3 1:1000 (Cell
Signaling, #8768), pERK 1:2000 (Cell Signaling, #9101), ERK 1:2000 (Cell
Signaling, #9102), pAKT 1:1000 (Cell Signaling, #4060), AKT 1:1000 (Cell
Signaling, #9272), pS6K 1:1000 (abcam, ab126818), S6K 1:1000 (abcam,
ab131440), Beta-actin 1:5000 (Santa Cruz, sc-81178), HDAC4 1:1000 (R&D
Systems, AF6205), SIK2 1:1000 (R&D Systems, AF5737), NRG1 1:1000 (R&D
Systems, AF396NA), GAPDH 1:5000 (Cell Signaling, #2118).

Immunofluorescence: EGFR 1:250 (abcam, ab52894), pERK 1:250 (Cell


Signaling, #9101).

Immunohistochemistry: pEGFR 1:400 (Cell Signaling, #3777), pAKT 1:100


(Cell Signaling, #4060), pERK 1:100 (Cell Signaling, #9101).

All antibodies were validated for usage on human and the corresponding
assays by the distributor.
June 2017

2
10. Eukaryotic cell lines

nature research | life sciences reporting summary


a. State the source of each eukaryotic cell line used. SW480 cell line was obtained from ATCC

b. Describe the method of cell line authentication used. We performed short tandem repeat (STR) analysis (AmpFISTR Identifier,
Applied Biosystems) and matched the resulting data to ATCC STR database.

c. Report whether the cell lines were tested for mycoplasma We used Universal Mycoplasma Detection Kit (ATCC 30-1012K) to verify
contamination. absence of mycoplasma contamination in our cell line.

d. If any of the cell lines used in the paper are listed in the database No commonly misidentified cell lines were used in this study.
of commonly misidentified cell lines maintained by ICLAC,
provide a scientific rationale for their use.

` Animals and human research participants


Policy information about studies involving animals; when reporting animal research, follow the ARRIVE guidelines
11. Description of research animals
Provide details on animals and/or animal-derived materials used in Balb/c nude mice (6~8 week-old female)
the study.

Policy information about studies involving human research participants


12. Description of human research participants
Describe the covariate-relevant population characteristics of the Biospecimens from 429 patients were used in this study: 229 males, 199
human research participants. females, and 1 unknown. Median age at diagnosis was 56 years, ranging
from 18 to 83.

June 2017

You might also like