24
Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer Immunity In Vivo Graphical Abstract Highlights d Genome-scale in vivo CRISPR screens identified tumor- intrinsic factors of oncogenesis d Prkar1a is a strong genetic driver of oncogenesis d Prkar1a loss altered the transcriptome and proteome in immune responses d Prkar1a mutation in cancer cells remodels the tumor microenvironment Authors Adan Codina, Paul A. Renauer, Guangchuan Wang, ..., Jesse Rinehart, Rong Fan, Sidi Chen Correspondence [email protected] In Brief Codina et al. performed genome-scale in vivo CRISPR screens that robustly identified multiple regulators of tumor- intrinsic factors that alter the ability of cells to grow as tumors across different levels of immunocompetence. Characterization of Prkar1a, a convergent hit from these screens, showed that its knockdown leads to drastic alterations in the phosphoproteome and transcriptome, corresponding to changes in the tumor microenvironment. Codina et al., 2019, Cell Systems 8, 136–151 February 27, 2019 ª 2019 Elsevier Inc. https://doi.org/10.1016/j.cels.2019.01.004

Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

  • Upload
    others

  • View
    7

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

Article

Convergent Identification

and Interrogation ofTumor-Intrinsic Factors that Modulate CancerImmunity In Vivo

Graphical Abstract

Highlights

d Genome-scale in vivo CRISPR screens identified tumor-

intrinsic factors of oncogenesis

d Prkar1a is a strong genetic driver of oncogenesis

d Prkar1a loss altered the transcriptome and proteome in

immune responses

d Prkar1a mutation in cancer cells remodels the tumor

microenvironment

Codina et al., 2019, Cell Systems 8, 136–151February 27, 2019 ª 2019 Elsevier Inc.https://doi.org/10.1016/j.cels.2019.01.004

Authors

Adan Codina, Paul A. Renauer,

GuangchuanWang, ..., Jesse Rinehart,

Rong Fan, Sidi Chen

[email protected]

In Brief

Codina et al. performed genome-scale

in vivo CRISPR screens that robustly

identified multiple regulators of tumor-

intrinsic factors that alter the ability of

cells to grow as tumors across different

levels of immunocompetence.

Characterization ofPrkar1a, a convergent

hit from these screens, showed that its

knockdown leads to drastic alterations in

the phosphoproteome and

transcriptome, corresponding to changes

in the tumor microenvironment.

Page 2: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

Cell Systems

Article

Convergent Identification and Interrogationof Tumor-Intrinsic Factors that ModulateCancer Immunity In VivoAdan Codina,1,2,3,4,16 Paul A. Renauer,1,2,3,4,16 Guangchuan Wang,1,2,3,16 Ryan D. Chow,1,2,3,5,16 Jonathan J. Park,1,2,3,5,17

Hanghui Ye,1,2,3,17 Kerou Zhang,3,6,17 Matthew B. Dong,1,2,3,5,17 Brandon Gassaway,1,3,13 Lupeng Ye,1,2,3

Youssef Errami,1,2,3 Li Shen,1,2,3 Alan Chang,1,2,3 Dhanpat Jain,7,8 Roy S. Herbst,8,9,10,11 Marcus Bosenberg,7,11,12

Jesse Rinehart,1,3,13 Rong Fan,3,6 and Sidi Chen1,2,3,4,5,11,14,15,18,*1System Biology Institute, Integrated Science & Technology Center, 850 West Campus Drive, West Haven, CT 06516, USA2Department of Genetics, Yale University School of Medicine, 333 Cedar Street, New Haven, CT 06510, USA3Center for Cancer Systems Biology, Integrated Science & Technology Center, 850 West Campus Drive, West Haven, CT 06516, USA4MCGD Program, Yale University, 333 Cedar Street, New Haven, CT 06510, USA5Yale M.D.-Ph.D. Program, 367 Cedar Street, New Haven, CT 06510, USA6Department of Biomedical Engineering, Yale University, 55 Prospect Street, New Haven, CT 06511, USA7Department of Pathology, Yale University School of Medicine, New Haven, CT 06510, USA8Department of Medicine, Yale University School of Medicine, New Haven, CT 06510, USA9Department of Pharmacology, Yale University School of Medicine, New Haven, CT 06510, USA10Smilow Cancer Hospital, 35 Park St, New Haven, CT 06510, USA11Yale Comprehensive Cancer Center, 20 York Street, Ste North Pavilion 4, New Haven, CT 06510, USA12Department of Dermatology, Yale University School of Medicine, New Haven, CT 06510, USA13Department of Cellular and Molecular Physiology, Yale University, 333 Cedar St., New Haven, CT 06520, USA14Immunobiology Program, The Anlyan Center, 300 Cedar Street, New Haven, CT 06520, USA15Yale Stem Cell Center, Yale University School of Medicine, New Haven, CT 06510, USA16These authors contributed equally17These authors contributed equally18Lead Contact

*Correspondence: [email protected]://doi.org/10.1016/j.cels.2019.01.004

SUMMARY

The genetic makeup of cancer cells directs oncogen-esis and influences the tumor microenvironment. Inthis study, we massively profiled genes that function-ally drive tumorigenesis using genome-scale in vivoCRISPR screens in hosts with different levels of immu-nocompetence. As a convergent hit from thesescreens, Prkar1a mutant cells are able to robustlyoutgrow as tumors in fully immunocompetent hosts.Functional interrogation showed that Prkar1a lossgreatly altered the transcriptome and proteomeinvolved in inflammatory and immune responses aswell as extracellular protein production. Single-celltranscriptomic profiling and flow cytometry analysismapped the tumor microenvironment of Prkar1amutant tumors and revealed the transcriptomic alter-ations in host myeloid cells. Taken together, our datasuggest that tumor-intrinsic mutations in Prkar1a leadto drastic alterations in the genetic program of cancercells, thereby remodeling the tumormicroenvironment.

INTRODUCTION

As a key hallmark of cancer (Hanahan andWeinberg, 2011), can-

cer cells have the ability to escape immune surveillance (Kim

136 Cell Systems 8, 136–151, February 27, 2019 ª 2019 Elsevier Inc.

et al., 2007) using mechanisms such as immuno-editing (Dunn

et al., 2002). Immune escape leads to resistance to various forms

of cancer immunotherapy, especially checkpoint blockade

immunotherapy (CBI), by primary, adaptive, or acquired resis-

tance. A greater understanding of this process is therefore of

central importance for improving clinical outcomes (Sharma

et al., 2017). Complex genetic, cellular, and immunological fac-

tors influence oncogenesis under a fully functional immune sys-

tem (Chen and Mellman, 2013). The interactions between the

immune system and cancer cells are continuous, dynamic, and

evolving. These interactions span from the initial establishment

of a cancer cell to the development of metastatic disease, which

is dependent on immune evasion, alteration of molecular fea-

tures, cellular composition, and the tumor-immune microenvi-

ronment (Chen and Mellman, 2017). Our understanding of

this phenomenon has just begun to be elucidated by pioneering

studies, which have shown both tumor-extrinsic and tumor-

intrinsic mechanisms. Tumor-extrinsic mechanisms include

non-tumor components within the tumor microenvironment

(TME), including regulatory T cells (Tregs), myeloid-derived sup-

pressor cells (MDSCs), M2 macrophages, and other inhibitory

immune checkpoints. All of these are able to contribute to the in-

hibition of anti-tumor immune responses and are thus also of

fundamental importance to CBI efficacy (Sharma et al., 2017).

Tumor-intrinsic mechanisms of immune evasion include a lack

of antigenic proteins, low mutational burden, impaired antigen

presentation by the silencing or deletion of key molecules such

as Transporter Associated with Antigen Processing (TAP),

Page 3: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

Beta-2-Microglobulin (B2M), and Human Leukocyte Antigen

(HLA) (Sharma et al., 2017). Additionally, constitutive expression

of immunosuppressive ligands such as Programmed Cell Death

Ligand 1 (PD-L1), either via transcriptional activation, epigenetic

regulation, or truncation of the 30 UTR has been shown to be im-

mune suppressive (Kataoka et al., 2016). Immune evasion can

also arise from oncogenic pathways that directly influence the

malignant nature of cancer cells, resulting in the alteration of

the tumor-immune microenvironment. Examples include onco-

genic signaling through the MAPK pathway resulting in the inhi-

bition of T cell recruitment and function (Liu et al., 2013), loss of

PTEN that enhances PI3K or ATK signaling (Lastwika et al., 2016;

Parsa et al., 2007), EGFR mutations that activate PD-1 pathway

(Akbay et al., 2013), overexpression of MYC that enforces the

expression of CD47 or PD-L1 (Casey et al., 2016), expression

of CDK5 to dampen the ability of T cells to reject tumors (Dorand

et al., 2016), stabilization of b-catenin resulting in constitutive

WNT signaling that induces T cell exclusion (Spranger et al.,

2015), modulation of interferon gamma (IFNg) receptor signaling

molecule Janus kinase 2 (JAK2) (Zaretsky et al., 2016), direct

changes in the expression of INFg receptors 1 and 2 (IFNGR1

and IFNGR2) and interferon regulatory factor 1 (IRF1) (Gao

et al., 2016), as well as epigenetic modifications of immune

genes (Heninger et al., 2015).

Several screening systems in human cells and mouse models

have previously demonstrated success in identifying essential

genes in cancer immunotherapy (Manguso et al., 2017; Pan

et al., 2018; Patel et al., 2017). We reasoned that performing

in vivo tumorigenesis screens in hosts with different levels of

immunocompetence would provide an approach to discover

genetic factors in cancer cells that mediate tumor progression

and immune response. Rag1�/� mice lack mature T and

B cells because of the deficiency of Rag1.Nu/Nu are Foxn3 defi-

cient and athymic. C57BL/6 mice are commonly used wild-type

mice that are fully immunocompetent. Thus, these mice present

with different levels of immunocompetence. In this study,

through robust genetic screening, we have convergently identi-

fied Prkar1a as a potent suppressor of oncogenesis in multiple

tiers of immunocompetence. We functionally interrogated its

role in regulating gene expression and cellular properties for

modulating the immune response and remodeling the TME.

RESULTS

Convergent In Vivo Genome Scale Knockout ScreensIdentify Prkar1a as a Suppressor of Oncogenesis acrossMultiple Hosts with Different Immune CompetenceWe used a genome-scale in vivo CRISPR screening approach

to identify oncogenic drivers inmultiple hostswith different levels

of immune competence. Specifically, we utilized a p53�/�;Myc-

primed mouse hepatocyte cell line (denoted IM, or IMC9 after

transduction with Cas9-GFP), which is immortalized but not

transformed. We screened this line in multiple genotypes,

including Rag1�/� mice, Nu/Nu mice, and fully immunocompe-

tent C57BL/6 mice as a syngeneic host (Figure 1A). The hepato-

cytes were capable of forming tumors in Rag1�/� mice (10/10,

100%) but not Nu/Nu mice up to an injection dose of 10 million

cells (0/10, 0%), suggesting that the immune system in Nu/Nu

mice can keep tumor growth in check. In sharp contrast,

when we mutagenized the cells with a genome-scale CRISPR

knockout library (mBrie) (Doench et al., 2016) (STAR Methods),

all (21/21) injections led to rapid tumor formation (Kolmogorov-

Smirnov [KS] test, p = 7e�4) (Figures 1A, 1B, and S1A;

Table S1). Neither the parental line (0/4) nor themBrie library mu-

tagenized pool (0/20) was able to generate tumor in C57BL/6

mice observed for at least 90 days. These data together sug-

gested that the loss-of-function mutations in a subset of genes

enabled the cells to robustly grow tumors within the Nu/Nu host.

To determine which mutants comprised the tumors that grew

inNu/Numice, we deep-sequenced the sgRNA library represen-

tations within the tumors, plasmid, and transduced cells prior to

transplantation (STAR Methods). Library representation ana-

lyses revealed high coverage in three infection control replicates

(97.8%on average) and recurrent patterns of sgRNA representa-

tion in tumors (correlation test among tumors, p < 2.2e�16) (Da-

taset S1; Figures S1B and S1C). Analyses of gene-targeting and

non-targeting sgRNA distributions revealed high consistency

between replicates and the overall shifts among different sample

types (Figures S1D and S1E), providing a high-quality genome-

scale in vivo screen dataset for analysis of enriched sgRNAs in

the outgrown tumors. When comparing the mean library repre-

sentation of in vivo tumor cells to that in pre-injection cells, we

detected a population of sgRNAs that were highly enriched in

the tumors,many of which targeted the same genewith indepen-

dent sgRNAs, whereas non-targeting control (NTC) sgRNAs

rarely conferred tumor outgrowth (Figures 1C, S2A, and S2B).

By ranking the genes by the number of independent sgRNAs

across all 21 tumors, we found that 2,335 genes have at least

one significant sgRNAs passing a false discovery rate (FDR) cut-

off of 0.2% (Figure 1D). Out of those, 120 genes had more than

two significant sgRNAs, and 12 genes had all four of their target-

ing sgRNAs significantly enriched, serving as independent evi-

dence of the knockout effect (Figure 1D). We then analyzed

sgRNA enrichment using criteria for sgRNA abundance and sig-

nificant incidence rate across independent tumors (Figure 1E)

and independent sgRNAs. This approach identified genetic

mutants that represent the strength of selection, prevalence of

phenotype, and robustness of constructs. Our analyses revealed

a three-way convergence of robust hits, identifying 11 genes that

were significantly enriched over all statistical criteria, including

Apc, Csnk1a1, Nf2, Prkar1a, Pten, Rbsn, Serpinb7, Tsc1, Tsc2,

Vps45, and Zbtb40 (Figure 1E). Additional analyses using RIGER

(STARMethods), an established CRISPR screen analytical pipe-

line that utilizes the second best performing sgRNAs, revealed a

highly similar set of genes, with 16 genes at themaximum level of

statistical significance (Tables S2 and S3; Figures 1F and 2A).

Further analysis using MAGeCK, a model-based approach with

consideration of the empirical null distribution of NTCs (STAR

Methods), also identified a similar set of genes, with 9 genes

having a top-level statistical significance of positive selection

signature (Tables S4 and S5; Figures 1G and 2A). Therefore,

this screen identified high-confidence candidate genes that

were strongly enriched in tumors that escaped the immune sys-

tem of the Nu/Nu host.

To further enhance the rigor and stringency of identification, we

repeated the screen with an independent CRISPR knockout li-

brary (mGeCKO) using the same optimized screening strategy

(Dataset S2). This screen identified five top genes that passed

Cell Systems 8, 136–151, February 27, 2019 137

Page 4: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

A

B C D

E F G

Figure 1. Genome-Scale In Vivo CRISPR Screens in Different Mouse Hosts Identified Genes Regulating Oncogenesis

(A) Schematics of the experiment. Left, p53�/�;Mycmutated non-malignant hepatocyte cell lines are capable of growing as tumors in immunodeficient Rag1�/�

mice with deficiency in both innate and adaptive immune system; however, they are rejected byNu/Numice. Right, genome-scale CRISPR screen and validation

for mutants that can grow as tumors. Genome-scale CRISPR libraries (mBrie or mGeCKO) mutagenized cell pools have robust tumor growth in Nu/Nu host.

Validation of individual gene mutant cells with top hits in mouse hosts with different levels of immune competence.

(B) Tumor growth curves of mBrie library transduced hepatocytes (n = 21), as compared to various injections using the parental cell line (n = 3 each) over a 30-day

period (Kolmogorov-Smirnov [KS] test, mBrie versus all controls, p = 7e�4). Error bars are SEM.

(C) A color-coded scatterplot showing hit identification for genes whose loss of function lead to consistent tumor growth. The screen was performed using a

second-generation genome-scale CRISPR knockout library (mBrie) transforming hepatocytes that can grow in fully immunodeficient Rag1�/� mice but not in

Nu/Numice that have competent innate immunity although lacking mature T cells. Library-transduced cells were capable of escaping the innate immune system

and grew as tumors at 100% penetrance. Average sgRNA library representation across all Nu/Nu tumors (n = 21) was plotted against average sgRNA library

representation across all cell replicates (n = 3). The 11 top hits that pass all three statistical criteria were plotted as color-coded dots, with each dot representing

one sgRNAwhere the same color represents independent sgRNAs for the same gene. Non-targeting control sgRNAs (NTCs) were shown as dark gray dots. Other

gene-targeting sgRNAs (GTSs) were shown as light gray dots.

(D) A barplot of the number of independent scoring sgRNAs in themBrie screen. False discovery rate (FDR) of 0.2%was used as a cutoff for scoring in each tumor.

Four sgRNAs per gene were contained in the mBrie library. Highly consistent hits withR3 scoring sgRNAs (14 genes with 3/4 scoring sgRNAs, 12 genes with 4/4)

were shown.

(E) A Venn diagram showing the overlap of hits identified using three selection criteria, revealing 11 genes significant regardless of criteria. The first criterion is by

abundance (i.e., a gene with one or more sgRNA(s) comprising ofR2% total reads in a tumor); the second criterion is by phenotypic penetrance (i.e., a gene with

the frequency of one or more sgRNA(s) passing FDR 0.2% cutoff across 25% of all independent tumor samples); the third criteria is by independent construct

(i.e., a gene with R2 sgRNA(s) passing FDR 0.2% cutoff).

(F) RIGER analysis of the genome-scale screen showing top hits.

(G) MAGeCK analysis of the genome-scale screen showing top hits.

See also Figures S1 and S2.

138 Cell Systems 8, 136–151, February 27, 2019

Page 5: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

Figure 2. Convergent Screen Analysis and Validation Revealed Prkar1a Mutant Cancer Cell Robustly Generate Tumors in a Fully Immuno-

competent Host

(A) A Venn diagram showing the convergence of two genome-scale in vivo CRISPR screens, with hits identified using the custom 3-Way Hit Caller pipeline. (Left)

The screen was repeated with a genome-scale CRISPR library (mGeCKO), and 5 genes were identified as top hits passing all statistical criteria. ThemBrie screen

andmGeCKO screen shared 3 hits:Prkar1a,Apc, and Tsc2. (Middle) Convergence of two genome-scale in vivoCRISPR screens, with hits identified using RIGER.

(Right) Convergence of mBrie in vivo CRISPR screens, showing consistent hits using three different hit calling algorithms (custom 3-Way HitCaller, MAGeCK,

and RIGER).

(B) Quantification of tumor sizes of the top screen hit mutants. (Top) tumor volume measurement of mutants in C57BL/6J host 2 weeks post injection. Prkar1a

mutant cells grew rapidly in this host, whereas NTC, Apc, Tsc2, or Csnk1a1 mutants did not (n = 10 each); (Bottom) tumor volume measurement of mutants in

Nu/Nu host 2.5 weeks post injection. Two-sided unpaired Mann-Whitney test was used to assess statistical significance for Prkar1a, Apc, Tsc2, or Csnk1a1

mutants, relative to NTC (ns, not significant, i.e., p > 0.05; *, p < 0.05; **, p < 0.01; ***, p < 0.001).

(C) The rate of tumor incidence of the top screen hit mutants. (Top) The tumor incidence rate of mutants in C57BL/6J host. Two-sided Fisher’s exact test was used

to assess statistical significance for Prkar1a, Apc, Tsc2, or Csnk1a1 mutants, relative to NTC (ns, not significant, i.e., p > 0.05; *, p < 0.05; **, p < 0.01;

***, p < 0.001).

(D) Representative H&E images of tumors in (C). Scale bar, 250 mm.

See also Figures S2 and S3.

Cell Systems 8, 136–151, February 27, 2019 139

Page 6: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

A

B C D

E

F G

(legend on next page)

140 Cell Systems 8, 136–151, February 27, 2019

Page 7: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

the stringent statistical criteria using our custom 3-way hit caller

pipeline (FDR < 0.2%, R2 sgRNAs, R2% read abundance, and

R25% of all tumors), of which three (Apc, Prkar1a, and Tsc2)

overlap with the most stringent hits from the mBrie screen

(Figure 2A). Similarly, four genes were identified as overlapping,

stringent hits from both screens using RIGER and MAGeCK

(Figure 2A). Multiple independent sgRNAs targeting these genes

were consistently found at high levels across tumor samples

relative to the infection-control cell samples, upon verification

of single genes (Figures S2C–S2F). Apc, Prkar1a, and Tsc2 are

the three genes that robustly emerged independent of statistical

criteria, computational algorithm, and screening library (Figures

2A and S2C–S2F).

To validate Prkar1a and other top candidates, we performed

tumor challenges that used single sgRNAs to target these genes

in immortalized hepatocyte cells, followed by transplantation

into Nu/Nu mice or fully immunocompetent C57BL/6 mice.

Compared to NTC controls, all four top mutants identified from

the screens were validated as potent tumor generators when

transplanted into Nu/Nu mice (Figures 2B–2D and S3A–S3E).

In sharp contrast, only the Prkar1a mutant is capable of gener-

ating tumors in immunocompetent C57BL/6 mice (Figures 2B–

2D), whereas mice challenged by cells with strong drivers

mutated, including Apc, Tsc2, and Csnk1a1, remained tumor-

free for the whole duration of the experiment (60 days) (Figures

2B–2D). These data indicated that Prkar1a knockout endowed

the immortalized hepatocytes with the ability to develop tumors

in fully immunocompetent hosts.

Phosphoproteomic Characterization of Prkar1aMutants Identified Its Downstream Signaling TargetsBecause of the strong phenotype of Prkar1a knockout, we set

out to interrogate the molecular regulation of Prkar1a and its

downstream pathways. Prkar1a encodes the alpha regulatory

subunit of the cAMP-dependent type I protein kinase (Protein

kinase A, PKA). As a result of the potential direct regulatory roles

of Prkar1a in phosphorylation signaling, we investigated the

changes in kinase activity and the phosphoproteome upon

Figure 3. Phosphoproteomic Characterization of Prkar1a Mutant Cells

(A) Schematics of Prkar1a phosphoproteomic experiments. Two independent set

proteomics approach that included stable isotope labeling using amino acids

enrichment of phosphopeptides, cation exchange chromatography (CEX) fracti

quently quantified, identified, and assembled into groups of proteins, based on s

(B) Waterfall plots of the top upregulated and downregulated proteins as MS-ide

mutants (Prkar1a/NTC log2 normalized-protein ratio [nr]) are plotted. As expecte

(C) Color-coded scatterplot of differential phosphorylation between independen

mutant sample (Prkar1a/NTC log2 nr) are plotted for each phosphorylation site (PS

red dots, and top differential PS (DPS, >3-fold change) are labeled with protein s

(D) Western of STAT3 and phospho-STAT3 (pSTAT3) in Prkar1a mutant and NTC

(E) Waterfall plots of the top hyper-phosphorylation site (PS) and hypo-PS in Prk

mutants (Prkar1a/NTC log2 nr) are labeled with the gene symbol (bold), UniProt a

peptide sequence in a 31 residue window, centered on the PS residue.

(F) Waterfall plot of the Ingenuity Pathway Analysis of protein groups with PS

pathways are ranked by significance (�log10 p value), and the bar colors represe

(G) Identification of hyper- and hypo-phosphorylation motifs in Prkar1amutant ce

hypo-PS (bottom) motifs that are over-represented in Prkar1amutant cells. Each

score for the enrichment below the motif. (Right) Logos are shown for the top kin

(bottom) motifs in Prkar1amutant cells. Up to four of the top kinase-substrate mo

substrate motifs are shown with the significance of the motif match (upper right of

(upper left of each Logo).

Prkar1a loss via stable isotope labeling using amino acids in

cell culture (SILAC) (Figure 3A). Because of the amount of

materials needed for SILAC experiments, we performed the

proteomics experiments in culture outside a host context, with

the goal to map out the cancer cells’ intrinsic properties.

We labeled the Prkar1a mutant cells (sgPrkar1a) with Lys6/

Arg6 ‘‘heavy’’ stable isotope and control cells (NTC) with Lys0/

Arg0 ‘‘light’’ stable isotope, and proteins were extracted,

mixed, trypsin-digested, TiO2-enriched for phosphopeptides,

fractionated by cation exchange chromatography (CEX) and

analyzed by liquid chromatography-tandem mass spectrometry

(LC-MS/MS) using an LTQ Orbitrap XL (STAR Methods). Two

independent SILAC-pair replicates were used to compare

Prkar1a mutants to control cells at the peptide, protein, and

post-translational modification levels, at which we identified a

respective 19,979 unique peptides (peptide spectrum matches,

PSMs), 2,940 protein groups and 2,994 phosphorylation sites

(PSs) after strict quality control filtering (Datasets S3, S4, and

S5). From this dataset, we identified differentially regulated pro-

teins, as well as a set of differentially phosphorylated peptides

upon Prkar1a loss, based on phosphorylation fold change at a

peptide PS and by the number of differentially phosphorylated

peptides in a protein group (Figures 3B, 3C, and 3E). As a robust-

ness check, PRKAR1A protein level was confirmed to be one of

the most downregulated proteins in the sgPrkar1a group (Fig-

ure 3B), validating the proteomics measurements. Additional

validation of the data was assessed by the western blot quanti-

fication of STAT3, one of the most upregulated proteins in the

dataset and a key player in immune and inflammation signaling

(Figures 3B and 3D). The proteins with the most significant hy-

per-phosphorylation upon Prkar1a loss included IWS1, LIMCH1,

and GOLGB1 (Figure 3E). The phospho-signaling pathway

changes in Prkar1a mutants, based on differentially phosphory-

lated proteins, showed that the most significantly enriched

pathway change in Prkar1a mutants was integrin-linked kinase

(ILK) signaling, Rho GTPase signaling, PKA signaling, mTOR

signaling, JAK/STAT signaling, and others, all of which modulate

oncogenesis and extracellular matrix (ECM) remodeling

Identified Its Downstream Signaling Targets

s of Prkar1amutant and NTC cells were compared using a multiplexed shotgun

in cell culture (SILAC), followed by protein extraction, trypsin digestion, TiO2

onation, and LC-MS/MS analysis by LTQ Orbitrap XL. Peptides were subse-

hared peptide sequences.

ntified protein groups in Prkar1a mutants. The protein level changes in Prkar1a

d, the PRKAR1A protein is among the most downregulated.

t phosphoproteomic experiments. Phosphorylation changes in each Prkar1a

). Hypo-PS and hyper-PS (>1.5-fold change) are shown as respective blue and

ymbols.

cell samples.

ar1a mutants (Left and right panels, respectively). The PS changes in Prkar1a

ccession number, and phosphorylated residue (in parentheses), as well as the

changes (DPS; >1.5-fold change) in Prkar1a mutants. The top ten canonical

nt the activation Z score, which is presented to the right of each bar.

lls. (Left) Sequence log-odds diagrams (Logos) are shown for hyper-PS (top) or

motif is presented above the Logo with an asterisk following the PS and a motif

ase-interacting motifs that match to the identified hyper-PS (top) or hypo-PS

tifs (Bonferroni-adjusted p < 1e�3) are presented for each PS motif. All kinase-

each Logo) and the kinase name with phosphorylated residues in parentheses

Cell Systems 8, 136–151, February 27, 2019 141

Page 8: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

A

−2

−1

0

1

2

sgPrkar1a

z-score

sgNTCsg1 sg1 sg1 sg1sg1 sg1 sg2 sg2 sg2sg3 sg3 sg3 sg3sg3sg3 -6 -4 -2 0 2 4 6

sgPrkar1a/sgNTC Expression(log2 fold change)

0

40

80

120260

300

pvalue(-log

10)

●●●●●●●●●●●●

Extracellular regionExtracellular spaceProteinaceous extracellular matrixExtracellular exosomecAMP catabolic processResponse to drug3',5'-cyclic-AMP phosphodiesterase activityExtracellular matrix

Molecular mediator of immune responseRibonucleotide catabolic processFatty acid metabolic process

GSEA: Upregulated with sgPrkar1a

responseMolecular mediator of

immune responseRibonucleotide

catabolic process

-log10 FDR q-value

Fatty acidmetabolic process

0.0 0.5 1.0 1.5 2.0 2.5

DAVID: Upregulated with sgPrkar1a

-log10 q value0 2 4 6 8 10 12

3',5'-cyclic-AMPphosphodiesterase activity

Response to drug

cAMP catabolic processExtracellular exosome

Proteinaceous extracellularmatrix

Extracellular space

Extracellular region

0.3

Enrichmentscore

0.20.1

0.40.50.60.7

0.0

sgPrkar1ahigh

sgNTChigh

B

C F G

H

D

E

sgPrkar1asgNTCsg1 sg1 sg1 sg1sg1 sg1 sg2 sg2 sg3sg3 sg3 sg3 sg3sg2sg3

FmodCol6a3Col5a2Col8a2PrelpCol15a1Mamdc2CrtapCcdc80Ccdc80CtgfCtgfCyr61Adamts20Mmp8HpseSpon2VcanEcm1Ccbe1Ecm1Col24a1Lama4Ecm1Sfrp1Adamts9Wnt11VcanThbs4EpycHmcn1

-2 -1 0 1 2

Proteinaceous extracellular matrixAcute inflammatory response

-2 -1 0 1 2

Il20rbPhb2OsmrAsh1lC2Adam8C2Ash1lCfhCfhCfpRhbdd3Phb2Il20rbOsmrCfpRhbdd3PhbRhbdd3CfhIl20rbPhb2PhbC1qbpPhb2Il20rbCd55Tnfrsf11aCd55Il6Ccr7PpargPtgs2Tnfrsf11aPla2g4aPhb2C1qbpPhbOsmrCd55Cd55PhbOsmrPhb2C8gRhbdd3Cd46Ash1lRhbdd3SpnCd55CfhSusd4C2Il6stPros1Il6stGstp1Gstp1Masp1Il6stC3Tnfsf11

Serping1C4bPhbOsmrSerping1Rhbdd3Rhbdd3Adora1C1qbpAdam8CfhOsmrIl6stSpnC8aIl6st

sgPrkar1asgNTCsg1 sg2 sg2 sg2sg1 sg1 sg3 sg3 sg1sg3 sg3 sg3 sg1sg1sg3

z Score

sgPrkar1a

sgNTC

0

200

400

600

800

1000

ELISA: Secreted IL-6

pg/mlIL-6

sgPrkar1a

sgNTC

* below detection

*

p <0.0001

(legend on next page)

142 Cell Systems 8, 136–151, February 27, 2019

Page 9: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

(Figure 3F). Comparative analysis of all PSs as a set showed a

shift in the kinase activity of Prkar1amutants, emphasized by al-

terations in the targetmotifs of differentially phosphorylated pep-

tides, for which the hypo-phosphorylated peptides are enriched

in the S*Pmotif, whereas the hyper-phosphorylated peptides are

enriched in RRxS* and RxxS* motifs (Figure 3G). These data

together revealed the proteome and phospho-proteome level al-

terations in Prkar1a mutant cells.

Whole-Transcriptome Profiling Identified DeregulatedGene Expression Programs in Prkar1a MutantsTo further dissect the gene product changes downstream of

phosphorylation, we utilized mRNA sequencing (mRNA-seq) to

profile the transcriptome of Prkar1a mutant cells together with

matched NTCs in culture, using three independent sgRNAs for

sgPrkar1a and two independent NTCs, with three biological

replicates for each sample group (STARMethods). We collected

total RNA from IMC9 cells at 7 days post-infection in biological

triplicates and performed strand-specific, poly-A enriched

mRNA-seq (Dataset S6). Using a differential expression analysis,

we identified a set of 371 transcripts that are differentially ex-

pressed between Prkar1a mutant and wild-type cells, with 119

transcripts being downregulated and 252 upregulated upon

Prkar1a loss (Dataset S7; Figures 4A and 4B). These differentially

expressed genes belonged to a distinct set of pathways with the

most significantly upregulated genes centered on immune, in-

flammatory, and secreted protein signatures (Figure 4B). Gene

Ontology analysis by the Database for Annotation, Visualization

and Integrated Discovery (DAVID) revealed that extracellular

proteins were among the top categories of Prkar1a-upregulated

genes, along with the anticipated on-target cAMP catabolic

process signature mediated by the feedback regulation of

PKA signaling via the Phosphodiesterases (Pde) gene family

(Tables S6 and S7; Figure S4A). The immune, inflammatory,

and cAMP-catabolism signatures were recapitulated with gene

set enrichment analysis (GSEA) (Tables S8 and S9; Figures 4D

and 4E). A closer look of specific transcripts revealed that

Figure 4. Whole Transcriptome Profiling of Prkar1aMutants Identified

Extracellular Protein Production

(A) An overall heatmap of all differentially expressed genes between Prkar1a mu

pendent NTCs were subjected to whole transcriptome analysis by mRNA-seq (n

gulated or downregulated in Prkar1a mutants) is shown by Z scores.

(B) A color-coded volcano plot of all differentially expressed transcripts betwee

corrected p value) was plotted against the log2 fold-change of normalized gene

most highly enriched gene sets, described in the legend.

(C) Waterfall plots of most significantly enriched gene ontologies for genes that

by the Database of Annotation, Visualization and Integrated Discovery (DAVID), a

sets including extracellular or secreted proteins emerged on top as most enric

phosphodiesterase activity, indicated the on-target activity of Prkar1a-regulated

(D) A waterfall plot of most significantly enriched gene sets of all Prkar1a upregula

levels considered during enrichment calculation (methods), and the top gene set

inflammatory response and molecular mediator of immune response emerged o

catabolic process indicated the on-target activity of Prkar1a-regulated PKA path

(E) A GSEA individual pathway analysis plot showing the enrichment of acute infl

(F) A heatmap of all differentially expressed transcripts between Prkar1a mutant

Relative gene expression levels were shown as Z scores.

(G) A heatmap of all differentially expressed transcripts between Prkar1amutant a

Relative gene expression levels were shown as Z scores.

(H) A barplot of IL-6 ELISA between Prkar1a mutant and control cells.

See also Figure S4.

Prkar1a significantly upregulated genes encode inflammatory

and secretory factors, such as cell surface proteins, cytokines,

ECM proteins, and molecular mediators of immune response

(Figures 4F, 4G, and S4B). Using ELISA, we further confirmed

that Prkar1a mutant cells turned on the protein production

switch of interleukin (IL)-6, a pro-inflammatory cytokine with

important roles in promoting tumor progression (Coussens

and Werb, 2002) (Prkar1a versus NTC, >800-fold change;

p < 0.0001; Figure 4H). Moreover, STAT3, the transcription factor

associated with IL-6 expression was also found to be upregu-

lated at the mRNA level, showing consistency between our

proteomic and transcriptomic data sets (p = 6.1e�25). Echoing

the phenotypes, these data suggest that cancer-cell-intrinsic

Prkar1a-loss leads to upregulated production of secreted cyto-

kines and extracellular proteins implicated in the inflammatory

response, which may influence the host TME.

Single-Cell Transcriptome Profiling Mapped the TumorMicroenvironment ofPrkar1aMutant Tumors in Hosts ofVarying ImmunocompetenceTo directly interrogate the host TME, we performed single-cell

RNA sequencing (scRNA-seq) (STAR Methods), an approach

that can elucidate both the gene expression signature and

cell population structure in a high-throughput manner. We re-

challenged Rag1�/�, Nu/Nu, and C57BL/6 mice with Prkar1a

and NTC cells (Figure 5A). As expected, Prkar1a cells grew tu-

mors in all three mice genotypes, whereas NTC cells only grew

in Rag1�/� (Figures 5A and 5B). We then sought to profile the

tumor-infiltrating immune repertoire and simultaneously profile

the cancer cells in the same tumor. However, since the vast

majority of cells in the tumor single-cell suspensions are

cancer cells, we performed fluorescence-activated cell sorting

(FACS) to purify both CD45+ immune cells and GFP+ cancer

cells. Using these enriched immune cell populations, we

subsequently spiked-in a smaller fraction of cancer cells taken

from the same tumor. The single cells were barcoded and had

cDNA libraries constructed in a single-cell manner, based on

an Upregulated Gene Expression Program of Immune Response and

tant and control cells. Three independent sgRNAs for Prkar1a and two inde-

= 3 biological replicates for each sgRNA). Differential gene expression (upre-

n Prkar1a mutant and control cells. The statistical significance (�log10 FDR-

expression levels. Each dot represents one gene, which is color-coded by the

are upregulated in Prkar1a mutants. The enrichment analysis was performed

nd the top gene sets are shown, as ranked by FDR-corrected q values. Gene

hed; Gene sets including cAMP catabolic process and response, as well as

PKA pathway.

ted genes. Pathway analysis was performed using GSEA with gene expression

s are shown, as ranked by FDR-corrected q values. Gene sets including acute

n top as the most significantly enriched ones; the gene set of ribonucleotide

way; gene sets are not mutually exclusive.

ammatory response genes in the Prkar1a mutant cells as compared to NTC.

and control cells for the genes in the acute inflammatory response category.

nd control cells for the genes in the proteinaceous extracellular matrix category.

Cell Systems 8, 136–151, February 27, 2019 143

Page 10: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

A

C

B

Figure 5. Single-Cell Profiling of Prkar1a Mutant Tumors in Hosts of Different Immune Competence

(A) Schematic for single-cell RNA-seq (scRNA-seq) experiments. IMC9 tumor cells were transfected with sgPrkar1a or NTC to generate Prkar1a knockout or

control cells, respectively. Transfected IMC9 cells were then subcutaneously injected into immunocompromised (Rag1�/�), partially immunocompromised

(Nu/Nu), and immunocompetent (C57BL/6) mice as shown, and the resulting tumors were extracted, and then sorted to CD45+ tumor-infiltrating immune cells

and tumor cells that were mixed in a 1:10 ratio for scRNA-seq. One representative mouse in each group was sequenced.

(B) H&E staining of tumors subjected to tumor-immune scRNA-seq. Scale bar, 250 mm.

(C) Cellular populations within the TME, visualized by t-distributed stochastic neighbor embedding (t-SNE) plot of clustered, dimensionality-reduced single-cell

transcriptional profiles. Each dot represents a single cell that is arranged relative to other cells based on transcriptional profiles, such that similar cells are grouped

together.

See also Figures S5 and S6.

previously established methods (Gierahn et al., 2017; Macosko

et al., 2015) (STAR Methods) (Figure 5A) (Datasets S8, S9, S10,

S11, S12, S13, and S14; Tables S10, S11, S12, S13, and S14).

The single-cell transcriptional data were dimensionality-reduced

by principal-component analysis (PCA), visualized by t-distrib-

uted stochastic neighbor embedding (t-SNE) and clustered by

shared nearest neighbor (SNN) methods. The single-cell clus-

tering identified distinct different subpopulations of the

sequenced TME cellular repertoire for the Prkar1a and NTC tu-

mors in Rag1�/� mice as well as the Prkar1a tumors in Nu/Nu

and C57BL/6 mice (Figure 5C).

By merging the transcriptome data from the single-cell popu-

lations of sgPrkar1a- and sgNTC-generated tumors in Rag1�/�

mice, we identified 3 clusters (Figure 6A), based on the optimal

cell separation in the t-SNE embedding, canonical immune

144 Cell Systems 8, 136–151, February 27, 2019

gene expression patterns, and optimal rank selection using

non-negative matrix factorization (NMF) (Brunet et al., 2004;

Gaujoux and Seoighe, 2010). We distinguished cluster 1 and 2

as immune cells, largely comprising myeloid populations, based

on the significantly upregulated expression of Cd14, Fcgr3

(Cd16), and Fcgr2b (Cd32), as well as increased expression of

Itgam (Cd11b) (MAST analysis; adjusted p < 1e�9) (Figures 6C

and S6), whereas cluster 3 cells are primarily cancer cells, based

on expression of tumor-specific marker genes, Mgp, Fkbp11,

Cdkn2a, andMrpl33, identified by comparing themost highly ex-

pressed genes in the IMC9 cells from bulk RNA-seq analysis to a

reference panel of immune cells comprising Immunological

GenomeProject data (ImmGen version 1) (Heng et al., 2008) (Fig-

ures 6C, S5B, and S6). In addition, we classified the clusters into

stromal and innate immune cell types by comparing the mean

Page 11: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

Figure 6. Differential Single-Cell Expression Analysis of the Immune Microenvironment in the Tumors Induced by Prkar1a Mutant and

Control Cancer Cells in Rag1–/– Mice

(A) t-distributed stochastic neighbor embedding (t-SNE) visualization of aligned single-cell transcriptional data generated from Prkar1a�/� or NTC tumors in

Rag1�/� mice. Cells of the t-SNE plot are labeled by the tumor dataset (top) and by unsupervised clusters that were identified (bottom).

(legend continued on next page)

Cell Systems 8, 136–151, February 27, 2019 145

Page 12: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

expression of cluster-specific differentially expressed genes

(MAST analysis of each cluster relative to all other cells; adjusted

p < 1e�3, 95th-percentile of jlog-scale fold-change expressionj)to an ImmGen reference panel, comprising 85 cell types (Figures

6B and S5D). This transcriptional comparison shows moderate

correlation (Kendall correlation coefficient t > 0.3) between

clusters 1–2 andmyeloid cell types, including macrophage, den-

dritic cell (DC), andmonocyte cells, with cluster 1 also correlating

with granulocytic neutrophils (Figure 6B). In contrast, the clus-

ter 3 population moderately correlated (t > 0.3) with fibroblast

cells and weakly (t > 0.2) with endothelial and epithelial cells

(Figure 6B).

Further categorization suggested that cluster 1 is best

described as a population of tumor-associated macrophages

(TAMs), with significantly increased expression (adjusted p <

1e�8) of key TAM markers, including Cd68, Cd163, and Csf1r

(Cd115) (Bronte et al., 2016). In addition, our data suggest that

both immune cell clusters contain MDSCs with notable expres-

sion of multiple MDSC markers (Bronte et al., 2016), including

significantly increased expression of Arg1, Cybb (Nox2), and

Il4ra markers, relative to tumor cells of cluster 3 (Figures 6C

and S6). Once the clusters were characterized, we assessed

the transcriptional differences related to Prkar1a loss, as it af-

fects each cell cluster. These differential expression analyses

show that Prkar1a loss is significantly associated with the

increased expression of the inflammatory marker Saa3 in all

three clusters (adjusted p < 1e�5), indicating that the inflamma-

tory phenotype observed by RNA-seq is conferred to the TME.

Decreased expression of MHCII genes in both immune cell

clusters was also observed, including Cd74, H2-Aa, H2-Ab1,

and H2-Eb1 (adjusted p < 1e�9; Figure 6D). In addition, the

chemokine gene Cxcl13 is significantly increased in cluster 1

cells (adjusted p < 4.23e�22; Figure 6D). Together, these data

mapped the TME of Prkar1a mutant tumors in Rag1�/� mice

and revealed the cellular repertoire as well as transcriptional

changes in immune genes.

Next, we investigated whether the same immune infiltrates are

found within immunocompetent C57BL/6 tumors generated by

the loss of Prkar1a. Wild-type tumors were not assessed in the

immunocompetent background, as NTC tumors did not form in

C57BL/6; therefore, only Prkar1a mutant tumors can be profiled

in this host to serve as a refinement but not comparison. We

repeated the single-cell transcriptomics analysis of eGFP+

IMC9 and CD45+ immune cells of C57BL/6 tumors as a 1:10

mixed cell ratio, with the 10X Genomics platform for scRNA-

(B) Identification of cell cluster populations based on expression correlation to kno

differentially expressed genes (MAST analysis; FDR-adjusted p < 0.001) with

expression of these marker genes in each cluster was compared by Kendall c

innate immune and stromal cells). The scaled correlation coefficients for the clus

unsupervised hierarchical clustering, performed separately within 6 broad cell ty

(C) Classification of cell clusters by cell-type-specific characteristic gene express

identified as myeloid cell populations by the gene expression of canonical myelo

(Cd11b). These clusters also express known markers of tumor-associated mac

represents the IMC9 cancer cells based on the expression of the previously defin

(D) Volcano plots showing differentially expressed genes in the TME between the

Each panel shows a cluster-specific differential expression analysis, for which the

fold-changes are shown in the x axes. Differentially expressed genes were defined

knockout tumor model. Overexpressed and underexpressed genes are depicted

See also Figure S7.

146 Cell Systems 8, 136–151, February 27, 2019

seq library prep prior to sequencing. After pre-processing, we

applied the Potential of Heat-diffusion for Affinity-based Trajec-

tory Embedding (PHATE) technique (Figure S5C Left) (Moon

et al., 2017). The ensuing population structure of the data

showed a notable split into two discrete groups of cells, which

represented immune and tumor subsets, based on the expres-

sion of Ptprc (Cd45) and Mgp genes, respectively (Figure S5C

Right). As the expression of these genes was not detected in

all cells, we partitioned the population structure into two clusters

using unsupervised SNN clustering. Next, we investigated the

immune cell populations in finer detail by subsetting the Ptprc+

cluster and visualizing the subpopulation structure by PHATE

dimensionality reduction, followed by unsupervised clustering

(Figure S5D). We next sought to categorize the five immune sub-

populations by comparing the expression of cluster-specific

genes (STAR Methods) in each cluster to 107 known immune

cells of a reference panel comprising neutrophils, macrophages,

monocytes, DCs, natural killer cells, and T cells (Figure S5E). The

correlation heatmap shows a macrophage signature in all im-

mune subclusters, with the strongest macrophage correlations

to subclusters 1, 2, and 4 (t > 0.5). These three clusters also

have moderately high correlations to DCs and monocytes

(t > 0.4), whereas subcluster 3 correlates moderately high with

neutrophil and macrophage cell types (t > 0.4). In support of

this cell-type categorization, we show significantly increased

expression of TAM-specific genes Adgre1 (F4/80), Csf1r

(Cd115), and H2-Aa (MHC-II) in subcluster 4 (adjusted p <

1e�5; Figure S7). We also found significantly increased expres-

sion of various myeloid markers in the adjacent subclusters of

1 and 3, which might represent populations of MDSCs. We

also show significantly upregulated expression of MDSC regula-

tory genes (Ddit3 and S100a8) and M-MDSC enzymes (Arg1 and

Nos2) in subcluster 3 cells (Figure S7). Furthermore, the data

show increased trends in PMN-MDSC-related genes, Cybb

(Nox2) and Cd244 (Figure S7). We see the expression of these

immune cell type genes follow gradient-like expression patterns

across cells in the PHATE plot (Figure S7 Bottom). As the PHATE

dimensionality reduction technique has been shown to preserve

both local and global relationships between cells (Moon et al.,

2017), these gradient-like transcriptional trends may reflect a

spectrum of the phenotype in these different myeloid cell types,

which will be interesting to explore in future single-cell analyses.

These data further corroborated the refined TME of Prkar1a

mutant tumors in the fully immunocompetent C57BL/6 wild-

type host.

wn immune cell populations. Cluster-specific marker genes were identified as

the top 95% fold-change between clusters, and the log-normalized mean

orrelation analysis to ImmGen V1 transcriptional data (85 cell populations of

ter-specific markers are shown for each ImmGen cell population, arranged by

pe categories of the ImmGen data.

ion patterns, as shown by expression-labeled t-SNE plots. Clusters 1 and 2 are

id cell surface markers such as Cd14, Fcgr3 (Cd16), Fcgr2b (Cd32), and Itgam

rophages (TAMs) and myeloid-derived suppressor cells (MDSCs). Cluster 3

ed IMC9-specific gene markers including Mgp, Fkbp11, Cdkn2a, and Mrpl33.

myeloid-like immune cells in tumors induced by Prkar1a�/� and control cells.

FDR-adjusted p values are shown in the y axes and the log-scale expression

as those with an adjusted p value < 0.01 and a fold-change > 1.8 in the Prkar1a

by red and blue points, respectively.

Page 13: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

CB

A

63.6%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%66666666666666666363336333333333333 666.66666666666666666666666666663666333333333333 63 63 633 63 63 666666666666666666666666666666666663333333 633 63 666666666666666666666666666666666666666663663636333633333 63 6666333.63.663 666666666666666666666666666666666636636333333333333333333.63 63 63 6666666666666666666666666363336333333.63 6.63.6663 63 663 63 66666666666666666633333 6663.633 66666666666666666666666666663633333333 63 6333 63 63 6333 66666666666666666666666666666633333333.666633 63.63 666666666666666666666666666666666666666666666666666666636336333 633 63 633 63666666666666666666666666666666636666333333333.63.6.63333333.36666666666666666666363333.33.3.33..3.3666666666666666636333333.3.3.33.666666666666663333.33.3366666666666633333333

0 104

105

0

-103

103

104

105

0 103

104

105

106

0

102

103

104

105

106

M-MDSCs

PMN-MDSCs

sgNTC TME

sgNTC

sgPrkar1a

DCs

TAMs

M-MDSCs

PMN-MDSCs

0

10

20

30

40

Immune cell populations

%ofCD45+Cells

**

**

0 104

105

PE/Cy7 CD45

APC/Cy7CD11b

FITC Gr-1

APCLy-6C

0

-103

103

104

105

0 103

104

105

106

0

102

103

104

105

106

M-MDSCs

PMN-MDSCs

sgPrkar1a TME

0 104

105

0

103

104

105

APC/Cy7 CD45.2

SSC-A

APC Ly6C

FITC

CD11b

PerCP/Cy5.5 MHC-II

SSC-A

PE F4/80

PE/Cy7CD24

59.4%

0 104

105

106

0

103

104

105

0 103

104

105

0

103

104

105

0 104

105

106

0

-103

103

104

105

106

DC

TAM

0 104

105

0

103

104

105

0 104

105

106

0

103

104

105

0 103

104

105

0

103

104

105

0 104

105

106

0

-103

103

104

105

106

DC

TAM

sgNTC TME

sgPrkar1a TME

APC/Cy7 CD45.2

SSC-A

APC Ly6C

FITC

CD11b

PerCP/Cy5.5 MHC-II

SSC-A

PE F4/80

PE/Cy7CD24

PE/Cy7 CD45

APC/Cy7CD11b

FITC Gr-1

APCLy-6C

Figure 7. Flow Cytometry Analysis of the

Immune Microenvironment in the Tumors

Induced by Prkar1aMutant and Control Can-

cer Cells

(A and B) Flow cytometry plots present the differ-

ences in the quantities of specific immune cell

subsets from tumors that were generated with or

without Prkar1a loss. (A) The gating schemes for the

detection of dendritic cells (DCs) and tumor-

associated macrophages (TAMs) are presented

using composite flow cytometry data from five

sgPrkar1a-generated tumors or six sgNTC-gener-

ated tumors in Rag1�/� mice. (B) Plots show the

flow cytometry gating scheme for the quantifi-

cation of monocytic and polymorphonuclear

myeloid-derived suppressor cells (M-MDSCs and

PMN-MDSCs, respectively) in tumors that were

generated with or without Prkar1a loss. The com-

posite flow cytometry data are shown for eight

sgPrkar1a-generated tumors or six sgNTC-gener-

ated tumors in Rag1�/� mice.

(C) Quantitative analysis of immune cell infiltration

differences between Prkar1a�/� and NTC tumors.

Immune cells were quantified from the flow

cytometry data in (A) and (B) and normalized to the

proportion of CD45/CD45.2+ cells for each tumor.

The immune cell populations were compared

between Prkar1a�/� and NTC tumors by unpaired

t test with Holm-Sidak correction for multiple-

testing. Significant population changes are de-

picted with asterisks (**, adjusted p < 1e�5).

Flow Cytometry Confirms the Alterations of TumorMicroenvironment in Prkar1a Mutant TumorsTo directly quantify Prkar1a knockout-mediated differences in

tumor-infiltrating immune cell populations, we performed flow cy-

tometry on eight tumors with Prkar1a loss and six tumors without.

First, we quantified DCs and TAMs (Figures 7A and 7C); however,

there are no significant changes between Prkar1a knockout and

wild-type tumors, yet there was a noteworthy decrease in TAMs

in Prkar1a knockout tumors (Figures 7A and 7C). Subsequently,

we inspected MDSC populations to show a significant increase

of PMN-MDSCs and a significant decrease of M-MDSCs in

Prkar1a knockout tumors (adjusted p = 4.99e�7 and 1.88e�6,

respectively; Figures 7B and 7C). Taken together, these data

show that the loss of Prkar1a leads to a polarized recruitment of

PMN-MDSCs, as well as a decrease in TAMs and the M-MDSCs

from which they are derived, representing a substantial alteration

of the host TME.

DISCUSSION

Malignant transformation, oncogenesis, and evading immune

destruction are all key hallmarks of cancer with fundamental roles

in thecancer immunecyclewithdirect consequenceson immuno-

Cell S

therapy outcomes (Chen and Mellman,

2013; Hanahan and Weinberg, 2011;

Sharma et al., 2017). Using multiple

whole-genomeknockout screens,we iden-

tified several candidate factors whose loss

of function consistently led to tumor

outgrowth in mouse hosts with different levels of immunocompe-

tence. Knocking out several hits enabled the cells to grow out in

Nu/Nu mice as tumors, while only Prkar1a knockout cells are

able to grow in fully immunocompetent C57BL/6mice. This is sur-

prising given that under the same genetic context, knockouts of

well-known tumor suppressors such as Tsc2 and Apc failed to

grow in the C57BL/6 host. Prkar1a encodes a regulatory subunit

of the PKA holoenzyme complex. Prkar1a is the only subunit of

the four whose knockout leads to pronounced disease-related

phenotypes. Prkar1a mutations were associated with disease in

the Carney complex (CNC), a disease syndrome characterized

by hyper-pigmented skin and sporadic neoplasias (Akin et al.,

2017; Alband and Korbonits, 2014; Stratakis, 2016). In this

context, hereditaryorgermlineLOFmutations inPrkar1aweresuf-

ficient to produce the disease in a dominant manner. In addition,

Prkar1a mutations are associated with several tumor types such

as cardiac, endocrine, thyroid, and lung adenocarcinoma (Kondo

et al., 2017; Wang et al., 2016) (Saloustros et al., 2017). Prkar1a

overexpression is also associatedwith tumor formation inmultiple

contexts, indicating that altering Prkar1a function leads to differ-

ential phenotypesacrosscellular and genetic backgrounds. Addi-

tionally, recent clinical findings show that patientswith a rare form

of liver cancer, known as fibrolamellar HCC, harbor the classical

ystems 8, 136–151, February 27, 2019 147

Page 14: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

DNAJB1-PRKACA fusion, which was recapitulated in a mouse

model of cancer (Kastenhuber et al., 2017; Rondell et al., 2017).

This fusion protein results in hyperactive PKA signaling, similar

to Prkar1a knockouts. Supporting the similarity, a subset of

CNC patients also develop fibrolamellar HCC, indicating signifi-

cant overlap between the phenotypes (Rondell et al., 2017). Given

the increased likelihood with which patients with CNC develop

various types of tumors,Prkar1awas speculated to be a suppres-

sor for early oncogenesis, where the immune system plays a

fundamental role in clearing mutated cells with aberrant growth

to preserve tissue homeostasis (Brown et al., 2017).

While its role as a tumor suppressor is now supported, the

molecular changes associated with Prkar1a�/�-mediated

tumorigenesis are less clear. Reports have shown that Prkar1a

knockout increases PKA activity, as well as resistance to

apoptosis via mTOR activation (Robinson-White et al., 2006;

Pringle et al., 2014). Additionally, it has also been shown that

Prkar1a regulates the ERK, Snail and E-cadherin pathways

(Wang et al., 2016). We investigated the mechanistic mecha-

nisms of Prkar1a in cancer cells, showing that its genetic pertur-

bation leads to drastic alterations in the proteome, phosphopro-

teome, and the transcriptome. These changes center on acute

inflammatory responses, mediators of immune response, and

production of secreted ECM proteins and cytokines. For

example, IL-6 production is dramatically upregulated, potentially

enhancing Prkar1a�/� cells’ ability to drive inflammation in the

absence of other cell types, as well as to drive early malignant

transformation in both an autocrine and paracrine manner.

Notably, an miRNA overexpressed in various cancers, miRNA-

1246, targets PKA pathway genes leading to an inflammatory

phenotype (Bott et al., 2017), which further supports our findings.

Single-cell transcriptome profiling and comparative analysis

of Prkar1a mutant and wild-type tumor cells in Rag1�/� mice re-

vealed the TME alterations in this host. The phenotypic effect of

Prkar1a knockout is too strong to have comparable tumors in the

hosts with more potent immune systems (Nu/Nu and C57BL/6,

where Prkar1a wild-type cells cannot grow as tumors). The

downregulation of MHC-II in the immune cells of Prkar1a tumors

on Rag1�/� mice may suggest gene expression changes in the

TAM populations or recruitment of myeloid-derived suppressive

cells. Alternatively, the downregulation of antigen presentation

in resident macrophages is in line with the reduced MHCII

expression. Furthermore, characterization of the immune cell

infiltrates by flow cytometry in Rag1�/� mice showed a signifi-

cant increase of PMN-MDSCs relative to controls. PMN-MDSCs

are antigen-specific immunosuppressive cells, known to down-

regulate cytotoxic T cell activity resulting in tumor-specific

T cell tolerance (Kumar et al., 2016). While additional studies

will need to be performed to understand the significance of

this finding, it raises the possibility that the PMN-MDSCs could

play an outsized role in regulating the growth of Prkar1a�/�

tumors in hosts with intact adaptive immune system. Future

molecular characterization of Prkar1a may further unveil the

detailed molecular, cellular, and immunological mechanisms

underlying TME changes driven by this gene or other tumor-

intrinsic factors. In summary, by convergent genome-scale

screens and multiple sets of functional interrogation, we have

identified Prkar1a as a gene regulating a tumor-intrinsic genetic

program for remodeling the TME.

148 Cell Systems 8, 136–151, February 27, 2019

STAR+METHODS

Detailed methods are provided in the online version of this paper

and include the following:

d KEY RESOURCES TABLE

d CONTACT FOR REAGENT AND RESOURCE SHARING

d EXPERIMENTAL MODEL AND SUBJECT DETAILS

B Institutional Approval

d METHOD DETAILS

B Cell Culture

B Mice

B Virus Production of CRISPR Library

B Cellular CRISPR Library Production

B Cell Library Transplantation

B Mouse Organ Dissection, Imaging, and Histology

B Mouse Tissue Collection for Molecular Biology

B Genomic DNA Extraction

B Demultiplexing and Read Preprocessing

B Mapping of sgRNA Spacers and Quantitation of

sgRNAs

B Normalization and Summary-Level Analysis of sgRNA

Abundances

B Determining Significantly Enriched sgRNAs (Custom 3-

Way Hit Caller)

B Individual Gene Validation in Nu/Nu and

C57BL/6J Mice

B SILAC Proteomics Sample Preparation

B Proteomics Analysis

B Phosphorylation Site Motif Analysis

B Western Blot

B RNA-seq Differential Expression Analysis

B Pathway Enrichment Analysis

B ELISA

B TIL Isolation for scRNA-seq

B Pre-processing the Single-Cell RNA-seq Data

B Filtering and Visualization of scRNA-seq Data

B Cluster Analysis of scRNA-seq Data

B Flow Cytometry

B Blinding Statement

d QUANTIFICATION AND STATISTICAL ANALYSIS

d DATA AND SOFTWARE AVAILABILITY

B Reagent Availability

B Code Availability

B Data Availability and Accessions

SUPPLEMENTAL INFORMATION

Supplemental Information includes 7 figures, 14 tables, and14 datasets andcan

be found with this article online at https://doi.org/10.1016/j.cels.2019.01.004.

ACKNOWLEDGMENTS

We thank S. Zhang, M. Martinez, Y. Cheng, L. Guo, B. Fitzgerald, G. Gabel, A.

Ristic, A. Araki, G.L. Wang, C. Castaldi, D. Rimm, and A. Brooks for helping in

parts of the study design, experiments, or analyses. We thank all members in

Chen laboratory, as well as various colleagues in the Department of Genetics,

Systems Biology Institute, Cancer Systems Biology Center, MCGD Program,

Immunobiology Program, BBS Program, Cancer Center, and StemCell Center

at Yale for assistance and/or discussion. We thank the Center for Genome

Analysis, Center for Molecular Discovery, Pathology Tissue Services,

Page 15: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

Histology Services, High Performance Computing Center, West Campus

Analytical Chemistry Core and West Campus Imaging Core, and Keck

Biotechnology Resource Laboratory at Yale for technical support.

S.C. is supported by Yale SBI/Genetics Startup Fund, NIH/NCI

(1DP2CA238295-01, 1R01CA231112-01, 1U54CA209992-8697, 5P50CA196530-

A10805, and 4P50CA121974-A08306), DamonRunyonDale Frey Award (DFS-

13-15), Melanoma Research Alliance (412806 and 16-003524), St-Baldrick’s

Foundation (426685), Breast Cancer Alliance, Cancer Research Institute

(CLIP), AACR (499395 and 17-20-01-CHEN), The Mary Kay Foundation

(017-81), The V Foundation (V2017-022), Ludwig Family Foundation, DoD

(W81XWH-17-1-0235), Sontag Foundation, and Chenevert Family Foundation.

R.S.H. is supported by Yale SPORE in lung cancer (5P50CA196530). R.F. is

supported by NIH/NCI (1U54CA209992). A. Codina and P.A.R. are supported

by Yale Ph.D. training grant from NIH (T32GM007499). R.D.C., J.J.P., and

M.B.D. are supported by the Yale MSTP training grant from NIH

(T32GM007205). G.W. is supported by Cancer Research Institute Irvington

and RJ Anderson Postdoctoral Fellowships. H.Y. and K.Z. are supported by

USTC student scholarships.

AUTHOR CONTRIBUTIONS

A. Codina designed the experiments with S.C. and initiated the study. A. Co-

dina., G.W., and P.A.R. performed most experiments. P.A.R. and R.D.C. per-

formed most data analyses. J.J.P. and A. Chang assisted data analysis. H.Y.,

K.Z., M.D., L.Y., Y.E., L.S. and B.G. assisted experiments. D.J., R.S.H.,

M.B.D., J.R., and R.F. provided conceptual advice. S.C. conceived the study,

secured funding, and supervised the work. A. Codina., P.A.R., R.D.C., and

S.C. prepared the manuscript.

DECLARATION OF INTERESTS

The authors declare no competing interests.

Received: May 24, 2018

Revised: November 3, 2018

Accepted: January 22, 2019

Published: February 20, 2019

REFERENCES

Akbay, E.A., Koyama, S., Carretero, J., Altabef, A., Tchaicha, J.H.,

Christensen, C.L., Mikse, O.R., Cherniack, A.D., Beauchamp, E.M., Pugh,

T.J., et al. (2013). Activation of the PD-1 pathway contributes to immune

escape in EGFR-driven lung tumors. Cancer Discov. 3, 1355–1363.

Akin, S., Noyan, S., Dagdelen, S., Pasaoglu, I., Kaynaroglu, V., Askun, M.M.,

Bilen, C.Y., Kiratli, H., Baydar, D.E., Onder, S., et al. (2017). Unusual presenta-

tions of Carney complex in patient with a novel PRKAR1A mutation. Neuro

Endocrinol. Lett. 38, 248–254.

Alband, N., and Korbonits, M. (2014). Familial pituitary tumors. Handb. Clin.

Neurol. 124, 339–360.

Bailey, T.L., Boden, M., Buske, F.A., Frith, M., Grant, C.E., Clementi, L., Ren,

J., Li, W.W., and Noble, W.S. (2009). Meme SUITE: tools for motif discovery

and searching. Nucleic Acids Res. 37, W202–W208.

Bayona-Bafaluy, M.P., Acın-Perez, R., Mullikin, J.C., Park, J.S., Moreno-

Loshuertos, R., Hu, P., Perez-Martos, A., Fernandez-Silva, P., Bai, Y., and

Enrıquez, J.A. (2003). Revisiting the mouse mitochondrial DNA sequence.

Nucleic Acids Res. 31, 5349–5355.

Bott, A., Erdem, N., Lerrer, S., Hotz-Wagenblatt, A., Breunig, C., Abnaof, K.,

Worner, A., Wilhelm, H., M€unstermann, E., Ben-Baruch, A., and Wiemann, S.

(2017). miRNA-1246 induces pro-inflammatory responses in mesenchymal

stem/stromal cells by regulating PKA and PP2A. Oncotarget. 8, 43897–43914.

Bray, N.L., Pimentel, H., Melsted, P., and Pachter, L. (2016). Near-optimal

probabilistic RNA-seq quantification. Nat. Biotechnol. 34, 525–527.

Bronte, V., Brandau, S., Chen, S.H., Colombo, M.P., Frey, A.B., Greten, T.F.,

Mandruzzato, S., Murray, P.J., Ochoa, A., Ostrand-Rosenberg, S., et al.

(2016). Recommendations for myeloid-derived suppressor cell nomenclature

and characterization standards. Nat. Commun. 7, 12150.

Brown, S., Pineda, C.M., Xin, T., Boucher, J., Suozzi, K.C., Park, S., Matte-

Martone, C., Gonzalez, D.G., Rytlewski, J., Beronja, S., et al. (2017). Correction

of aberrant growth preserves tissue homeostasis. Nature 548, 334–337.

Brunet, J.P., Tamayo, P., Golub, T.R., and Mesirov, J.P. (2004). Metagenes

and molecular pattern discovery using matrix factorization. Proc. Natl. Acad.

Sci. USA 101, 4164–4169.

Butler, A., Hoffman, P., Smibert, P., Papalexi, E., and Satija, R. (2018).

Integrating single-cell transcriptomic data across different conditions, technol-

ogies, and species. Nat. Biotechnol. 36, 411–420.

Casey, S.C., Tong, L., Li, Y., Do, R., Walz, S., Fitzgerald, K.N., Gouw, A.M.,

Baylot, V., G€utgemann, I., Eilers, M., et al. (2016). MYC regulates the antitumor

immune response through CD47 and PD-L1. Science 352, 227–231.

Chen, D.S., and Mellman, I. (2013). Oncology meets immunology: the cancer-

immunity cycle. Immunity 39, 1–10.

Chen, D.S., and Mellman, I. (2017). Elements of cancer immunity and the

cancer-immune set point. Nature 541, 321–330.

Chen, S., Sanjana, N.E., Zheng, K., Shalem, O., Lee, K., Shi, X., Scott, D.A.,

Song, J., Pan, J.Q.,Weissleder, R., et al. (2015). Genome-wide CRISPR screen

in a mouse model of tumor growth and metastasis. Cell 160, 1246–1260.

Church, D.M., Goodstadt, L., Hillier, L.W., Zody, M.C., Goldstein, S., She, X.,

Bult, C.J., Agarwala, R., Cherry, J.L., DiCuccio, M., et al. (2009). Lineage-

specific biology revealed by a finished genome assembly of the mouse.

PLoS Biol. 7, e1000112.

Coussens, L.M., and Werb, Z. (2002). Inflammation and cancer. Nature 420,

860–867.

Dobin, A., Davis, C.A., Schlesinger, F., Drenkow, J., Zaleski, C., Jha, S., Batut,

P., Chaisson,M., andGingeras, T.R. (2013). STAR: ultrafast universal RNA-seq

aligner. Bioinformatics 29, 15–21.

Doench, J.G., Fusi, N., Sullender, M., Hegde, M., Vaimberg, E.W., Donovan,

K.F., Smith, I., Tothova, Z., Wilen, C., Orchard, R., et al. (2016). Optimized

sgRNA design to maximize activity and minimize off-target effects of

CRISPR-Cas9. Nat. Biotechnol. 34, 184–191.

Doench, J.G., Hartenian, E., Graham, D.B., Tothova, Z., Hegde, M., Smith, I.,

Sullender, M., Ebert, B.L., Xavier, R.J., and Root, D.E. (2014). Rational design

of highly active sgRNAs for CRISPR-Cas9-mediated gene inactivation. Nat

Biotechnol. 32, 1262–1267.

Dorand, R.D., Nthale, J., Myers, J.T., Barkauskas, D.S., Avril, S., Chirieleison,

S.M., Pareek, T.K., Abbott, D.W., Stearns, D.S., Letterio, J.J., et al. (2016).

Cdk5 disruption attenuates tumor PD-L1 expression and promotes antitumor

immunity. Science 353, 399–403.

Dunn, G.P., Bruce, A.T., Ikeda, H., Old, L.J., and Schreiber, R.D. (2002).

Cancer immunoediting: from immunosurveillance to tumor escape. Nat.

Immunol. 3, 991–998.

Finak, G., McDavid, A., Yajima, M., Deng, J., Gersuk, V., Shalek, A.K., Slichter,

C.K., Miller, H.W., McElrath, M.J., Prlic, M., et al. (2015). MAST: a flexible sta-

tistical framework for assessing transcriptional changes and characterizing

heterogeneity in single-cell RNA sequencing data. Genome Biol. 16, 278.

Gao, J., Shi, L.Z., Zhao, H., Chen, J., Xiong, L., He, Q., Chen, T., Roszik, J.,

Bernatchez, C., Woodman, S.E., et al. (2016). Loss of IFN-gamma pathway

genes in tumor cells as a mechanism of resistance to anti-CTLA-4 therapy.

Cell 167, 397–404.e9.

Gaujoux, R., and Seoighe, C. (2010). A flexible R package for nonnegative ma-

trix factorization. BMC Bioinformatics 11, 367.

Gierahn, T.M., Wadsworth, M.H., 2nd, Hughes, T.K., Bryson, B.D., Butler, A.,

Satija, R., Fortune, S., Love, J.C., and Shalek, A.K. (2017). Seq-Well: portable,

low-cost RNA sequencing of single cells at high throughput. Nat. Methods 14,

395–398.

Gupta, S., Stamatoyannopoulos, J.A., Bailey, T.L., and Noble, W.S. (2007).

Quantifying similarity between motifs. Genome Biol. 8, R24.

Hanahan, D., and Weinberg, R.A. (2011). Hallmarks of cancer: the next gener-

ation. Cell 144, 646–674.

Heng, T.S., and Painter, M.W.; Immunological Genome Project (2008). The

Immunological Genome Project: networks of gene expression in immune cells.

Nat. Immunol. 9, 1091–1094.

Cell Systems 8, 136–151, February 27, 2019 149

Page 16: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

Heninger, E., Krueger, T.E., and Lang, J.M. (2015). Augmenting antitumor im-

mune responses with epigenetic modifying agents. Front. Immunol. 6, 29.

Hu, J., Rho, H.S., Newman, R.H., Zhang, J., Zhu, H., and Qian, J. (2014a).

PhosphoNetworks: a database for human phosphorylation networks.

Bioinformatics 30, 141–142.

Hu, J.F., Rho, H.S., Newman, R.H., Hwang, W.C., Neiswinger, J., Zhu, H.,

Zhang, J., and Qian, J. (2014b). Global analysis of phosphorylation networks

in humans. Biochim. Biophys. Acta 1844, 224–231.

Huang, da W., Sherman, B.T., and Lempicki, R.A. (2009a). Bioinformatics

enrichment tools: paths toward the comprehensive functional analysis of large

gene lists. Nucleic Acids Res. 37, 1–13.

Huang, da W., Sherman, B.T., and Lempicki, R.A. (2009b). Systematic and

integrative analysis of large gene lists using David bioinformatics resources.

Nat. Protoc. 4, 44–57.

Kastenhuber, E.R., Lalazar, G., Houlihan, S.L., Tschaharganeh, D.F., Baslan,

T., Chen, C.C., Requena, D., Tian, S., Bosbach, B., Wilkinson, J.E., et al.

(2017). DNAJB1-PRKACA fusion kinase interacts with beta-catenin and the

liver regenerative response to drive fibrolamellar hepatocellular carcinoma.

Proc. Natl. Acad. Sci. USA 114, 13076–13084.

Kataoka, K., Shiraishi, Y., Takeda, Y., Sakata, S., Matsumoto, M., Nagano, S.,

Maeda, T., Nagata, Y., Kitanaka, A., Mizuno, S., et al. (2016). Aberrant PD-L1

expression through 3’-UTR disruption inmultiple cancers. Nature 534, 402–406.

Kim, R., Emi, M., and Tanabe, K. (2007). Cancer immunoediting from immune

surveillance to immune escape. Immunology 121, 1–14.

Kolde, R. 2019. Package ‘pheatmap’. https://cran.r-project.org/web/packages/

pheatmap/pheatmap.pdf.

Kondo, K., Harada, M., Konomoto, T., Hatanaka, M., and Nunoi, H. (2017).

Novel PRKAR1A mutation in Carney complex with cardiac myxoma. Pediatr.

Int. 59, 840–841.

Kumar, M.S., Pester, R.E., Chen, C.Y., Lane, K., Chin, C., Lu, J., Kirsch, D.G.,

Golub, T.R., and Jacks, T. (2009). Dicer1 functions as a haploinsufficient tumor

suppressor. Genes Dev. 23, 2700–2704.

Kumar, V., Patel, S., Tcyganov, E., and Gabrilovich, D.I. (2016). The nature of

myeloid-derived suppressor cells in the tumor microenvironment. Trends

Immunol. 37, 208–220.

Langmead, B., Trapnell, C., Pop, M., and Salzberg, S.L. (2009). Ultrafast and

memory-efficient alignment of short DNA sequences to the human genome.

Genome Biol. 10, R25.

Lastwika, K.J., Wilson, W., 3rd, Li, Q.K., Norris, J., Xu, H., Ghazarian, S.R.,

Kitagawa, H., Kawabata, S., Taube, J.M., Yao, S., et al. (2016). Control of

PD-L1 expression by oncogenic activation of the AKT-mTOR pathway in

non-small cell lung cancer. Cancer Res. 76, 227–238.

Li, W., Xu, H., Xiao, T., Cong, L., Love, M.I., Zhang, F., Irizarry, R.A., Liu, J.S.,

Brown, M., and Liu, X.S. (2014). MAGeCK enables robust identification of

essential genes from genome-scale CRISPR/Cas9 knockout screens.

Genome Biol. 15, 554.

Liu, C., Peng, W., Xu, C., Lou, Y., Zhang, M., Wargo, J.A., Chen, J.Q., Li, H.S.,

Watowich, S.S., Yang, Y., et al. (2013). BRAF inhibition increases tumor infiltra-

tion by T cells and enhances the antitumor activity of adoptive immunotherapy

in mice. Clin. Cancer Res. 19, 393–403.

Lun, A.T., McCarthy, D.J., and Marioni, J.C. (2016). A step-by-step workflow

for low-level analysis of single-cell RNA-seq data with Bioconductor.

Version 2. F1000Res, 2122.

Luo, B., Cheung, H.W., Subramanian, A., Sharifnia, T., Okamoto, M., Yang, X.,

Hinkle, G., Boehm, J.S., Beroukhim, R., Weir, B.A., et al. (2008). Highly parallel

identification of essential genes in cancer cells. Proc Natl Acad Sci U S A. 105,

20380–20385.

Macosko, E.Z., Basu, A., Satija, R., Nemesh, J., Shekhar, K., Goldman, M.,

Tirosh, I., Bialas, A.R., Kamitaki, N., Martersteck, E.M., et al. (2015). Highly par-

allel genome-wide expression profiling of individual cells using nanoliter drop-

lets. Cell 161, 1202–1214.

Manguso, R.T., Pope, H.W., Zimmer, M.D., Brown, F.D., Yates, K.B., Miller,

B.C., Collins, N.B., Bi, K., LaFleur, M.W., Juneja, V.R., et al. (2017). In vivo

150 Cell Systems 8, 136–151, February 27, 2019

CRISPR screening identifies Ptpn2 as a cancer immunotherapy target.

Nature 547, 413–418.

Martin, M. (2011). Cutadapt removes adapter sequences From high-

throughput sequencing reads. EMBnet J. 17, 10–12.

Mi, H., Muruganujan, A., Casagrande, J.T., and Thomas, P.D. (2013). Large-

scale gene function analysis with the PANTHER classification system. Nat

Protoc. 8, 1551–1566.

Moon, K.R., van Dijk, D., Wang, Z., Chen, W., Hirn, M.J., Coifman, R.R.,

Ivanova, N.B., Wolf, G., and Krishnaswamy, S. (2017). PHATE: A dimension-

ality reduction method for visualizing trajectory structures in high-dimensional

biological data. bioRxiv. https://doi.org/10.1101/120378v3.

Newman, R.H., Hu, J.F., Rho, H.S., Xie, Z., Woodard, C., Neiswinger, J.,

Cooper, C., Shirley, M., Clark, H.M., Hu, S.H., et al. (2013). Construction of hu-

man activity-based phosphorylation networks. Mol. Syst. Biol. 9, 655.

Pan, D., Kobayashi, A., Jiang, P., Ferrari de Andrade, L., Tay, R.E., Luoma,

A.M., Tsoucas, D., Qiu, X., Lim, K., Rao, P., et al. (2018). A major chromatin

regulator determines resistance of tumor cells to T cell-mediated killing.

Science 359, 770–775.

Parsa, A.T., Waldron, J.S., Panner, A., Crane, C.A., Parney, I.F., Barry, J.J.,

Cachola, K.E., Murray, J.C., Tihan, T., Jensen,M.C., et al. (2007). Loss of tumor

suppressor PTEN function increases B7-H1 expression and immunoresist-

ance in glioma. Nat. Med. 13, 84–88.

Patel, S.J., Sanjana, N.E., Kishton, R.J., Eidizadeh, A., Vodnala, S.K., Cam, M.,

Gartner, J.J., Jia, L., Steinberg, S.M., Yamamoto, T.N., et al. (2017). Identification

of essential genes for cancer immunotherapy. Nature 548, 537–542.

Pimentel, H., Bray, N.L., Puente, S., Melsted, P., and Pachter, L. (2017).

Differential analysis of RNA-seq incorporating quantification uncertainty.

Nat. Methods 14, 687–690.

Pringle, D.R., Vasko, V.V., Yu, L., Manchanda, P.K., Lee, A.A., Zhang, X.,

Kirschner, J.M., Parlow, A.F., Saji, M., Jarjoura, D., et al. (2014). Follicular thy-

roid cancers demonstrate dual activation of PKA and mTOR as modeled by

thyroid-specific deletion of Prkar1a and Pten in mice. J Clin Endocrinol

Metab. 99, E804–E812.

Robinson-White, A.J., Leitner, W.W., Aleem, E., Kaldis, P., Bossis, I., and

Stratakis, C.A. (2006). PRKAR1A inactivation leads to increased proliferation

and decreased apoptosis in human B lymphocytes. Cancer Res. 66,

10603–10612.

Robinson, M.D., McCarthy, D.J., and Smyth, G.K. (2010). edgeR: a

Bioconductor package for differential expression analysis of digital gene

expression data. Bioinformatics 26, 139–140.

Rondell, P.G., Lackner, K., Terracciano, L., Gonzalez-Cantu, Y., Maleszewski,

J.J., Greipp, P.T., Simon, S.M., and Torbenson, M.S. (2017). Fibrolamellar car-

cinoma in the Carney complex: PRKAR1A loss instead of the classic DNAJB1-

PRKACA fusion. Hepatology 68, 1441–1447.

Saloustros, E., Salpea, P., Starost, M., Liu, S., Faucz, F.R., London, E., Szarek,

E., Song, W.J., Hussain, M., and Stratakis, C.A. (2017). Prkar1a gene knockout

in the pancreas leads to neuroendocrine tumorigenesis. Endocr. Relat. Cancer

24, 31–40.

Sanjana, N.E., Shalem, O., and Zhang, F. (2014). Improved vectors and

genome-wide libraries for CRISPR screening. Nat. Methods 11, 783–784.

Satija, R., Farrell, J.A., Gennert, D., Schier, A.F., and Regev, A. (2015). Spatial

reconstruction of single-cell gene expression data. Nat. Biotechnol. 33,

495–502.

Schwartz, D., and Gygi, S.P. (2005). An iterative statistical approach to the

identification of protein phosphorylation motifs from large-scale data sets.

Nat. Biotechnol. 23, 1391–1398.

Shalem, O., Sanjana, N.E., Hartenian, E., Shi, X., Scott, D.A., Mikkelson, T.S.,

Heckl, D., Ebert, B.L., Root, D.E., Doench, J.G., et al. (2014). Genome-scale

CRISPR-Cas9 knockout screening in human cells. Science 343, 84–87.

Sharma, P., Hu-Lieskovan, S., Wargo, J.A., and Ribas, A. (2017). Primary, adap-

tive, and acquired resistance to cancer immunotherapy. Cell 168, 707–723.

Spranger, S., Bao, R., and Gajewski, T.F. (2015). Melanoma-intrinsic beta-

catenin signalling prevents anti-tumour immunity. Nature 523, 231–235.

Page 17: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

Stratakis, C.A. (2016). Carney complex: A familial lentiginosis predisposing to a

variety of tumors. Rev. Endocr. Metab. Disord. 17, 367–371.

Subramanian, A., Tamayo, P., Mootha, V.K., Mukherjee, S., Ebert, B.L.,

Gillette, M.A., Paulovich, A., Pomeroy, S.L., Golub, T.R., Lander, E.S., et al.

(2005). Gene set enrichment analysis: a knowledge-based approach for inter-

preting genome-wide expression profiles. Proc. Natl. Acad. Sci. USA 102,

15545–15550.

Tyanova, S., Temu, T., and Cox, J. (2016). The MaxQuant computational plat-

form for mass spectrometry-based shotgun proteomics. Nat. Protoc. 11,

2301–2319.

van der Maaten, L.J.P. (2014). Accelerating t-SNE using Tree-Based

Algorithms. J. Mach. Learn. Res. 15, 3221–3245.

van der Maaten, L.J.P., and Hinton, G.E. (2008). Visualizing High-Dimensional

Data Using t-SNE. J. Mach. Learn. Res. 9, 2579–2605.

Waltman, L., and van Eck, N.J. (2013). A smart local moving algorithm for

large-scale modularity-based community detection. Eur. Phys. J. B 86.

Wang, S., Cheng, Y., Zheng, Y., He, Z., Chen, W., Zhou, W., Duan, C., and

Zhang, C. (2016). PRKAR1A is a functional tumor suppressor inhibiting

ERK/Snail/E-cadherin pathway in lung adenocarcinoma. Sci. Rep.

6, 39630.

Zaretsky, J.M., Garcia-Diaz, A., Shin, D.S., Escuin-Ordinas, H., Hugo, W., Hu-

Lieskovan, S., Torrejon, D.Y., Abril-Rodriguez, G., Sandoval, S., Barthly, L.,

et al. (2016). Mutations associated with acquired resistance to PD-1 blockade

in melanoma. N. Engl. J. Med. 375, 819–829.

Zender, L., Spector, M.S., Xue, W., Flemming, P., Cordon-Cardo, C., Silke, J.,

Fan, S.T., Luk, J.M., Wigler, M., Hannon, G.J., et al. (2006). Identification and

validation of oncogenes in liver cancer using an integrative oncogenomic

approach. Cell 125, 1253–1267.

Cell Systems 8, 136–151, February 27, 2019 151

Page 18: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

STAR+METHODS

KEY RESOURCES TABLE

REAGENT or RESOURCE SOURCE IDENTIFIER

Antibodies

Phospho-Stat3 (Tyr705) (D3A7) XP Cell Signaling Technologies RRID:AB_2491009

Stat3a (D1A5) XP Rabbit mAb #8768 Cell Signaling Technologies RRID:AB_2722529

GAPDH antibody Santa Cruz Biotech RRID:AB_627679

Ultra-LEAF Purified anti-mouse CD3 3Antibody

(clone: 145-2C11)

Biolegend RRID:AB_11149115

APC anti-mouse CD3 3Antibody (clone: 145-2C11) Biolegend RRID:AB_312677

FITC anti-mouse Ly-6G/Ly-6C (Gr-1) Antibody

(clone: RB6-8C5)

Biolegend RRID:AB_313370

FITC anti-mouse/human CD11b Antibody (clone: M1/70) Biolegend RRID:AB_312789

PE/Dazzle594 anti-mouse CD11c Antibody (clone: N418) Biolegend RRID:AB_2563655

PerCP/Cy5.5 anti-mouse I-A/I-E Antibody (clone: m5/114.15.2) Biolegend RRID:AB_2191071

APC anti-mouse Ly-6C Antibody (clone: HK1.4) Biolegend RRID:AB_1732076

PE anti-mouse F4/80 Antibody (clone: BM8) Biolegend RRID:AB_893486

APC/Cy7 anti-mouse CD45.2 Antibody (clone: 104) Biolegend RRID:AB_830789

APC/Cy7 anti-mouse/human CD11b Antibody (clone: M1/70) Biolegend RRID:AB_830641

PE/Cy7 anti-mouse CD24 Antibody (clone: M1/69) Biolegend RRID:AB_756048

PE/Cy7 anti-mouse CD45 Antibody (clone: 30-F11) Biolegend RRID:AB_312978

Bacterial and Virus Strains

One Shot Stbl3 Chemical Competent E. Coli ThermoFisher Catalog Number: C737303

Endura ElectroCompetent Cells Lucigen Catalog Number: 60242-2

Critical Commercial Assays

Qubit dsDNA HS Assay Kit ThermoFisher Catalog Number: Q32854

Chromium Single Cell 10x Genomics N/A

Gibson Assembly Master Mix NEB Catalog Number: E2611

Collagenase, Type IV, powder ThermoFisher Catalog Number: 17104019

QuickExtract DNA Extraction Solution Epicenter Catalog Number: QE09050

Proteinase K Sigma-Aldrich Catalog Number: 3115879001

RNase A Qiagen Catalog Number: 19101

Phusion Flash High-Fidelity PCR Master Mix ThermoFisher Catalog Number: F548L

DreamTaq Green PCR Master Mix (2X) ThermoFisher Catalog Number: K1082

E-Gel Low Range Quantitative DNA Ladder ThermoFisher Catalog Number: 12373031

T7E1 NEB Catalog Number: M0302L

TRIzol Plus RNA Purification Kit ThermoFisher Catalog Number: 12183555

QIAquick Gel Extraction Kit Qiagen Catalog Number: Q28706

Deposited Data

UniProt C57BL6J UP000000589 proteome

(last modified: May 16, 2017)

Bayona-Bafaluy et al.,

(2003); Church et al., (2009)

http://www.uniprot.org/proteomes/

UP000000589

PhosphoNetworks kinase phosphorylation motif data Hu et al., (2014a), (2014b);

Newman et al., (2013)

http://www.phosphonetworks.org/

download/motifMatrix.csv

Immunological Genome Project expression data (version 1) Heng et al., (2008) GEO: GSE15907

Mouse genome (mm10) Macosko et al., (2015) GEO: GSE63472

Prkar1a mutant mRNA-seq This study GEO: PRJNA506858

Prkar1a mutant proteomics This study PRIDE: PXD011837

Prkar1a mutant single-cell RNA-seq This study GEO: PRJNA509929

(Continued on next page)

e1 Cell Systems 8, 136–151.e1–e7, February 27, 2019

Page 19: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

Continued

REAGENT or RESOURCE SOURCE IDENTIFIER

Experimental Models: Cell Lines

HEK293FT ThermoFisher Catalog Number: R70007

Myc;p53-/- hepatocyte Zender et al. (2006) Clone IM

Myc;p53-/- hepatocyte Zender et al. (2006) Clone 172

Experimental Models: Organisms/Strains

C57BL/6J Jackson Laboratory RRID:IMSR_JAX:000664

Nu/Nu Jackson Laboratory RRID:IMSR_JAX:002019

Rag1-/- Jackson Laboratory RRID:IMSR_JAX:002216

Recombinant DNA

mGeCKOa Sanjana et al., (2014) https://www.addgene.org/pooled-library/

zhang-mouse-gecko-v2/

mBrie Doench et al. (2014) https://www.addgene.org/pooled-library/

broadgpp-mouse-knockout-brie/

Software and Algorithms

FlowJo software 9.9.6 FlowJo, LLC

Cutadapt 1.12 Martin, (2011) https://github.com/marcelm/cutadapt/

Bowtie 1.1.2 Langmead et al., (2009) https://sourceforge.net/projects/bowtie-bio/

files/bowtie/1.1.2

GSEA Subramanian et al., (2005) http://software.broadinstitute.org/gsea/

DAVID Huang et al., (2009b) https://david.ncifcrf.gov/

Scran R package Lun et al., (2016) https://github.com/MarioniLab/scran/

Rtsne R package van der Maaten, (2014);

van der Maaten

and Hinton, (2008)

https://github.com/jkrijthe/Rtsne/

Edge R package Robinson et al., (2010) https://github.com/StoreyLab/edge/

Panther classification system Mi et al., (2013) http://pantherdb.org/

RIGER Luo et al. (2008) https://software.broadinstitute.org/GENE-E/

extensions/RIGER.jar

MAGeCK Li et al. (2014) http://mageck.sourceforge.net

Dropseq_tools-1.12 Macosko et al., (2015) http://mccarrolllab.com/dropseq/

Seurat R package Satija et al., (2015) http://satijalab.org/seurat/

STAR 2.5.3a Dobin et al., (2013) https://github.com/alexdobin/STAR/

R: The R Project for Statistical Computing The R Foundation https://www.r-project.org/

MaxQuant version 1.6.0.1 Tyanova et al., (2016) https://www.biochem.mpg.de/5111795/

maxquant

MEME software suite version 4.12.0 Bailey et al., (2009) http://meme-suite.org/

Tomtom motif comparison Gupta et al., (2007) http://meme-suite.org/

Kallisto Bray et al., 2016 https://pachterlab.github.io/kallisto/

Sleuth Pimentel et al., (2017) http://pachterlab.github.io/sleuth/

Picard 2.9.0 Broad Institute, 2018 http://broadinstitute.github.io/picard/

Modification Motifs (MoMo) Cheng, and Mellman,

(2017) (BioRx)

http://meme-suite.org/

Ingenuity Pathway Analysis Qiagen https://analysis.ingenuity.com

phateR R package Moon et al., (2017) (BioRx) https://github.com/KrishnaswamyLab/phateR

MAST R package Finak et al., 2015 https://github.com/RGLab/MAST/

pheatmap R package Kolde, (2019) https://github.com/raivokolde/pheatmap

CONTACT FOR REAGENT AND RESOURCE SHARING

All resources related to this work are available upon reasonable request to the lead contact, Sidi Chen ([email protected]).

Cell Systems 8, 136–151.e1–e7, February 27, 2019 e2

Page 20: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

EXPERIMENTAL MODEL AND SUBJECT DETAILS

Institutional ApprovalAll experimental work involving recombinant DNA was performed under the guidelines of Yale University Environmental Health and

Safety (EHS) committee with an approved protocol (Chen-rDNA-15-45). All animal work was performed under the guidelines of Yale

University Institutional Animal Care and Use Committee (IACUC) with approved protocols (Chen-2015-20068; Chen-2018-20068)

and was consistent with the Guide for Care and Use of Laboratory Animals, National Research Council, 1996 (Institutional Animal

Welfare Assurance No. A-3125-01).

METHOD DETAILS

Cell CultureImmortalized hepatocyte cell lines with the background of p53-/-;cMyc (IM and 172) were from (Zender et al., 2006). Lung cancer cells

were from (Kumar et al., 2009). Cells were transduced with lenti-Cas9 or lenti-Cas9-GFP to generate Cas9-positive cells. All cell lines

were grown under standard conditions using DMEM containing 5% FBS & 1% Penstrep in a CO2 incubator unless otherwise noted.

Cell culture for SILAC experiment was described below.

MiceSeveral mouse strains with various levels of immunocompetence were used as transplant hosts, including C57BL/6, Nu/Nu and

Rag1-/-. Mice, both sexes, between the ages of 6-12 weeks of age were used for the study. All animals were housed in standard

individually ventilated, pathogen-free conditions, with 12h:12h or 13h:11h light cycle, room temperature (21-23�C) and 40-60% rela-

tive humidity. When a cohort of animals was receiving multiple treatments, animals were randomized by 1) randomly assign

animals to different groups using littermates, 2) random mixing of females prior to treatment, maximizing the evenness or represen-

tation of mice from different cages in each group, and/or 3) random assignment of mice to each group, in order to minimize the

effect of gender, litter, small difference in age, cage, housing position, where applicable. All mouse experiments follow ARRIVE

guidelines.

Virus Production of CRISPR LibraryFirst-generation and second-generation mouse genome-scale CRISPR sgRNA libraries, mGeCKO and mBrie, respectively, from

(Sanjana et al., 2014) and (Doench et al., 2016). Libraries were amplified and re-sequenced by barcoded Illumina sequencing with

customized primers to ensure proper representation. For virus production, envelope plasmid pMD2.G, packaging plasmid psPAX2,

and library plasmid were added at ratios of 1:1:2.5, and then polyethyleneimine (PEI) was added and mixed well by vortexing. The

solution was stand at room temperature for 10-20min, and then themixture was dropwisely added into 80-90% confluent HEK293FT

cells and mixed well by gently agitating the plates. Six hours post-transfection, fresh DMEM supplemented with 10% FBS and 1%

Pen/Strep was added to replace the transfection media. Virus-containing supernatant was collected at 48 h and 72 h post-

transfection, and was centrifuged at 1500 g for 10 min to remove the cell debris; aliquoted and stored at -80�C. Virus was titrated

by infecting IM cells at a number of different concentrations, followed by the addition of 2 mg/mL puromycin at 24 h post-infection

to select the transduced cells. The virus titers were determined by calculating the ratios of surviving cells 48 or 72 h post infection

and the cell count at infection.

Cellular CRISPR Library ProductionCells were transduced at 500x or higher coverage and at MOI between 0.2 to 0.4 to ensure sufficient coverage of the library and low

percentage of multiple infectants, using a method similar to (Chen et al., 2015; Shalem et al., 2014).

Cell Library TransplantationUn-transduced and library-transduced cells were injected subcutaneously into the right and left flanks ofNu/Numice at 2e7 cells per

flank. The mGeCKO- or mBrie- transduced IMC9 cells were treated in a similar manner, with the exception that 2.5e7 cells were

injected into the right flank only. Tumors were measured every 2 to 3 days by caliper, and their sizes were estimated as spheres.

Statistical significance was assessed by Welch’s t-test, given the unequal sample numbers and variances for each treatment

condition.

Mouse Organ Dissection, Imaging, and HistologyMicewere sacrificed by carbon dioxide asphyxiation or deep anesthesia with isoflurane followed by cervical dislocation. Mouse livers

and other organs were manually dissected and examined under a fluorescent stereoscope (Zeiss). Organs were then fixed in 4%

formaldehyde or 10% formalin for 48 to 96 hours, embedded in paraffin, sectioned at 6 mm and stained with hematoxylin and eosin

(H&E) for pathology. For tumor size quantification, H&E slides were scanned using an Aperio digital slidescanner (Leica). Tumor sizes

in area were measured using ImageScope (Leica). Immunohistochemistry (IHC) was done with well-established markers using

standard protocols (see Key Resources Table for reagent information). IHC staining quantification was done manually using

ImageScope (Leica), where applicable.

e3 Cell Systems 8, 136–151.e1–e7, February 27, 2019

Page 21: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

Mouse Tissue Collection for Molecular BiologyMouse organs were dissected and collected manually under a dissection scope where applicable. For molecular biology, tissues

were flash frozen with liquid nitrogen, ground in 24 Well Polyethylene Vials with metal beads in a GenoGrinder machine (OPS diag-

nostics). Homogenized tissues were used for DNA/RNA/protein extractions using standard molecular biology protocols.

Genomic DNA ExtractionA total of 3e7 cells per replicate, or 200-800 mg of frozen ground tissue each sample, was used for gDNA extraction for library

readout. Samples were re-suspended in 6 mL of NK Lysis Buffer (50 mM Tris, 50 mM EDTA, 1% SDS, pH 8.0) supplemented with

30 mL of 20 mg/mL Proteinase K (Qiagen) in 15 mL conical tubes, and incubated at 55�C bath for 2 h up to overnight. After all the

tissues have been lysed, 30 mL of 10 mg/mL RNAse A (Qiagen) was added, mixed well and incubated at 37�C for 30 min. Samples

were chilled on ice and then 2 mL of pre-chilled 7.5 M ammonium acetate (Sigma) was added to precipitate proteins. The samples

were inverted and vortexed for 15-30 s and then centrifuged atR 4,000 g for 10 min. The supernatant was carefully decanted into a

new 15mL conical tube, followed by the addition of 6 mL 100% isopropanol (at a ratio of� 0.7), inverted 30-50 times and centrifuged

atR 4,000 g for 10 minutes. Genomic DNA should be visible as a small white pellet. After discarding the supernatant, 6 mL of freshly

prepared 70% ethanol was added, mixed well, and then centrifuged at R 4,000 g for 10 min. The supernatant was discarded by

pouring; and remaining residues was removed using a pipette. After air-drying for 10-30 min, DNA was re-suspended by adding

200-500 mL of Nuclease-Free H2O. The genomic DNA concentration was measured using a Nanodrop (Thermo Scientific), and

normalized to 1000 ng/mL for the following readout PCR.

Demultiplexing and Read PreprocessingRaw single-end fastq read files were filtered and demultiplexed using Cutadapt (Martin, 2011). To remove extra sequences down-

stream (i.e. 3’ end) of the sgRNA spacer sequences, we used the following settings: cutadapt –discard-untrimmed -aGTTTTAGAGCT

AGAAATGGC. As the forward PCR primers used to readout sgRNA representation were designed to have a variety of barcodes to

facilitate multiplexed sequencing, we then demultiplexed these filtered reads with the following settings: cutadapt -g file:fbc.fasta

–no-trim, where fbc.fasta contained the 12 possible barcode sequences within the forward primers. Finally, to remove extra se-

quences upstream (i.e. 5’ end) of the sgRNA spacers, we used the following settings: cutadapt –discard-untrimmed –gGTGGAAAGG

ACGAAACACCG. Through this procedure, the raw fastq read files could be pared down to the 20 bp sgRNA spacer sequences.

Mapping of sgRNA Spacers and Quantitation of sgRNAsHaving extracted the 20 bp sgRNA spacer sequences from each demulitplexed sample, we then mapped the sgRNA spacers to the

library from which they originated (either mGeCKO or mBrie). To do so, we first generated a bowtie index of either sgRNA library

using the bowtie-build command in Bowtie 1.1.2 (Langmead et al., 2009). Using these bowtie indexes, we mapped the filtered fastq

read files using the following settings: bowtie -v 1 –suppress 4,5,6,7 –chunkmbs 2000 –best. Using the resultant mapping output, we

quantitated the number of reads that had mapped to each sgRNA within the library. To generate sgRNA representation bar plots, we

set a detection threshold of 1 read and counted the number of unique sgRNAs present in each sample.

Normalization and Summary-Level Analysis of sgRNA AbundancesWe normalized the number of reads in each sample by converting raw sgRNA counts to reads per million (rpm). The rpm values

were then subject to log2 transformation for certain analyses. To generate correlation heatmaps, we used the NMF R package

(Gaujoux and Seoighe, 2010) and calculated the Pearson correlations between individual samples using log2 rpm counts. To calcu-

late the cumulative distribution function for each sample group, we first averaged the normalized sgRNA counts across all

samples within a given group. We then used the ecdfplot function in the latticeExtra R package to generate empirical cumulative

distribution plots.

Determining Significantly Enriched sgRNAs (Custom 3-Way Hit Caller)Three criteria were used to identify enriched sgRNAs / genes: 1) if an sgRNA comprisedR 2% of the total reads in at least one tumor

sample; 2) if an sgRNAwas deemed statistically significantly enriched inR 25%of all tumor samples (formBrie) orR 50%of all tumor

samples (for mGeCKO) using a false-discovery rate (FDR) threshold of 2%based on the abundances of all nontargeting controls; or 3)

ifR 2 independent sgRNAs targeting the same gene were each found to be statistically significant at FDR < 2% in at least one tumor

sample. For the first and second criteria, individual sgRNA hits were collapsed to genes to facilitate comparisons with the hits from

the third criteria.

Individual Gene Validation in Nu/Nu and C57BL/6J MiceTwo to three sgRNAswere chosen for each validation target, corresponding to themost highly enriched sgRNAs foundwithin tumors.

Additionally, 3 NTCs sgRNAswere chosen randomly from themGeCKO library to serve as negative controls. All sgRNA spacers were

cloned in to lentiviral backbone, and virus was produced in HEK293FT cells as described above at a lower scale. IM.C9 cells were

infected with validation constructs and selected for via 5 mg/mL puromycin treatment to produce KO cell lines. For each cell line 4e5

cells were injected into both flanks of a Nu/Nu or C57BL/6 mouse 5-7 days post viral infection. Tumors were measured by caliper

every 2-3 days to assess growth rate.

Cell Systems 8, 136–151.e1–e7, February 27, 2019 e4

Page 22: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

SILAC Proteomics Sample PreparationBoth sgPrkar1a and NTC transduced IMC9 cells were grown in standard tissue culture conditions as described previously, with the

exception of the media used. sgPrkar1a cells were cultured with media containing lys6 and arg6 isotopes, deficient in the regular

forms of the amino acids. NTC cells were grown in standard media. Cells were cultured for one week, and > 5 passages to promote

full incorporation of the heavy isotopes. Cells were counted using a countess II machine and trypan blue staining. Protein was sub-

sequently extracted with RIPA buffer on ice for 30 minutes, using 200ml 1x buffer per million cells. Lysate was then spun down for

15 min at 14,000 g. Protease inhibitor cocktail (Roche) was then added to the extracted supernatant to prevent degradation. Protein

concentration was measured and then normalized by Bradford assay with BSA standard. Following Bradford lysates from heavy

sgPrkar1a cells and light NTC cells were mixed in a 1:1 ratio, followed by phosphopeptide enrichment. The flowthrough fractions

contain unmodified peptides, and the eluate fractions contain phosphopeptides, which were subjected to MassSpec analysis.

Proteomics AnalysisProteomic quantification and analysis of SILAC-labeled samples was performed using MaxQuant version 1.6.0.1 (Tyanova et al.,

2016). Specifically, raw LC-MS/MS data files were analyzed using the default Orbitrap and MS/MS instrument parameters for

Arg6/Lys6 ‘‘heavy’’ and Arg0/Lys0 ‘‘light’’ double SILAC-labeled peptides that had no more than three labeled residues. Peptide

searches were then performed using a reference panel, derived from the in silico trypsin digest of the UniProtC57BL/6Jmouse refer-

ence proteome (Proteome ID: UP000000589; last modified: May 16, 2017) (Bayona-Bafaluy et al., 2003; Church et al., 2009) with a

maximum of 2 missed cleavage sites. These peptide searches included carbamidomethyl (C) as a fixed modification, as well as

oxidation (M), acetyl (N-terminal), Gln/Glu -> pyro-Glu, phospho (STY) and deamination (NQ) as variable modifications. In addition,

each peptide was searched for a maximum of 6 variable modifications, a maximum of 200 modification combinations, and the

searches were set to include peptides with >= 7 residues, <= 4600 Da, and unspecific searches required a peptide length between

8 and 25 residues. Co-fragmented peptides from a MS/MS spectrum were identified and quantified. Missing patterns of incomplete

isotope pattern SILAC labeled pairs were re-quantified, based on the identified isotope pattern. The ‘‘Match between runs’ algorithm

was used to identify peptide features with amissingMS/MS spectrum or peptide sequence from another run, based on identifiedm/z

and RT information. For the protein-level analysis, protein group identification required >=1 razor/unique peptide, and protein group

quantification required at least 2 quantified peptide feature ratios between SILAC pairs. Protein ratio calculations included both razor

and unique peptides, as well as unmodified peptides and peptides with oxidation (M) and acetyl (N-terminal) modifications; however,

the unmodified counterparts/versions of these modified peptides were omitted from quantification. The protein group ratios were

calculated using the ‘‘advanced ratio estimation’’ method, which reports the median of the peptide feature ratios or the result of a

regression model for the intensities and log ratios if there is a correlation. Results from the peptide spectrum match (PSM), protein

group and phospho-site analyses had reverse and contaminant sequences removed and were filtered based on a 1% false-

discovery rate (FDR) that was calculated using the target-decoy strategy, in which decoy sequences were derived from reverse pro-

tein sequences. The PSM were also filtered based on the best spectrum match of a PSM (Andromeda score > 40), the difference

between the Andromeda score of the best hit with that of the second best sequence hit (delta-score > 8), and the posterior error

probability (PEP) of a false match (PEP < 0.01). The phospho-site level data were filtered as the PSM with the addition of a

localization probability threshold for post-translational modifications (probability > 0.75). Protein groups were also filtered based

on FDR (q value < 0.01).

Phosphorylation Site Motif AnalysisHypo- and hyper-phosphorylation site (hypo-PS and hyper-PS, respectively) motifs were identified by a motif discovery analysis

performed using the Modification Motifs (MoMo) tool from the MEME software suite version 4.12.0 (Bailey et al., 2009). The motif

discovery analysis used default settings with the Motif-X algorithm (Schwartz and Gygi, 2005) and included PS sequences, adjusted

to a 15 residue width with the PS of interest at the center, and the background sequence database was the UniProt mouse proteome,

which was used as the initial proteomics reference panel. The identified PS motifs were then compared to kinase phosphorylation

motifs using the Tomtom motif comparison tool (Gupta et al., 2007) from the MEME software suite. The motif comparison analysis

was performed using the default settings and kinase phosphorylation motif data from the PhosphoNetworks database (Hu et al.,

2014a; Hu et al., 2014b; Newman et al., 2013).

Western BlotNTC and sgPrkar1a cells were cultured in regular media as above for 48 hours. Total protein samples from were collected, and pro-

tease inhibitor cocktail was added to prevent degradation. Protein concentration was measured and then normalized by Bradford

assay with BSA standard. Western was performed using a standard molecular biology protocol with antibodies detailed in Key

Resources Table.

RNA-seq Differential Expression AnalysisStrand-specific single-end RNA-seq read files were analyzed to obtain transcript level counts using Kallisto (Bray et al., 2016), with

the settings –rf-stranded -b 100. Differential expression analysis between sg-NTC and sg-Prkar1a treated cells were then performed

using Sleuth (Pimentel et al., 2017), with default settings.

e5 Cell Systems 8, 136–151.e1–e7, February 27, 2019

Page 23: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

Pathway Enrichment AnalysisFor GSEA (Subramanian et al., 2005) we considered the entire set of genes in the differential expression analysis and sorted the genes

by their fold change values without regard to p-values. This ranked gene list was used for GSEA pre-ranked analysis, in reference to

the C5-all pathway database. We considered a pathway as statistically significant if the q-value was less than 0.1. Using an adjusted

p-value cutoff of 0.05, and a fold change threshold of ±1 (the ‘‘b’’ value from Sleuth output), we determined the set of transcripts

that were significantly upregulated or downregulated with Prkar1a deletion. Converting these transcripts to the gene level, we

then used the resultant gene sets for DAVID functional annotation analysis (Huang et al., 2009a, 2009b), where we only considered

GO categories. We considered a GO category as statistically significant if the Benjamini-Hochberg adjusted p-value was less

than 0.05.

ELISANTCand sgPrkar1a cells were cultured in regularmedia as above for 48 hours. Supernatant protein samples fromwere collected, and

protease inhibitor cocktail was added to prevent degradation. IL-6 ELISA was performed using a commercial kit (RnD).

TIL Isolation for scRNA-seqCells were cultured and prepared for subcutaneous injection as above. Five million cells were injected into each mouse, and tumors

was harvested for scRNA-seq experiments. Tumor tissue was immediately placed in PBS containing 2% FBS to maintain cell

viability. Tumor tissue was then manually cut into small pieces using a razor in a 10cm dish containing 5ml PBS containing 2%

FBS. Tumor tissue was then suspended into 15-20ml PBS and loaded into gentleMACs tubes. Collagenase IV was added in a

1:100 dilution to each sample, prior to loading into the gentleMACs tissue dissociation machine. Tissue was then dissociated for

30’ at 37 C. After dissociation samples were passed through a 100uM filter and re-suspended into 40mL of PBS. Samples were

then spun down at 400g for 5 min, and supernatant was discarded. Cell pellets then were washed with 40mL PBS, and spun

down again at 400g for 5 min and supernatant was discarded. 3mL ACK lysis buffer was added to each cell pellet, gently pipetting

to mix, then incubated for 3-5 min. After incubation samples were re-suspended in 40mL PBS and passed through a 40uM filter to

remove debris. Cells were then washed again, and stained with anti-CD45.2 antibody for 30’ using a 1:100 dilution (1e6 cells per mL).

Prior to FACS, cells were washed 2x in PBS. Cells were then sorted based on CD45.2 (pan-immune cells) andGFP (cancer cells), and

mixed as 90%:10% for single cell transcriptomics analysis. Unique molecular identifier (UMI) and single cell barcoded library prep-

aration and sequencing was done using a slightly modified approach based on previously established methods (Gierahn et al., 2017;

Macosko et al., 2015).

Pre-processing the Single-Cell RNA-seq DataDrop-seq (version 1.12) (Macosko et al., 2015) tools were used to preprocessing all Seq-Well (Gierahn et al., 2017) single-cell RNA-

seq data unless stated otherwise. First, the paired-end read fastq files were merged and converted to Bam files by Picard-

FastqToSam (Picard version 2.9.0). From the Bam files, cellular and unique molecular identifier barcodes were extracted, and reads

were filtered to exclude those with single-base quality scores below 10. The sequences were then trimmed to remove adapters and

polyA tails, and the files were converted back to fastq format by Picard-SamToFastq. The sequences were then mapped by STAR

aligner (version 2.5.3a) (Dobin et al., 2013) to the mm10 mouse reference genome, provided by the McCarroll lab (GEO dataset:

GSE63472) (Macosko et al., 2015). The aligned data were sorted by Picard-SortSam, and the cell and UMI barcode metadata

were attached using Picard-MergeBamAlignment. Gencode annotations were added to the aligned reads, and digital gene expres-

sion (DGE) files were prepared for the 1000 most frequently occurring cell barcodes. For data generated by Chromium Single-cell

3’ RNA-sequencing (10X Genomics), preprocessing was performed using the Cellranger software count-pipeline (10X Genomics),

aligning to the mm10 mouse genome reference with default settings. The pre-processed 10X and DGE matrices were then imported

using the Seurat package (Satija et al., 2015) for the R programming language (version 3.4.4), and the data were filtered to include

genes detected in >5 cells, and cells with >200 detected genes and <5% mitochondrial genes. The data were then normalized to

10,000 transcripts/cell and log-transformed, using a pseudo-count of 1 (ln[x + 1]).

Filtering and Visualization of scRNA-seq DataAll single cell RNA-seq data analysis pipelines were performed using the Seurat package for R (Macosko et al., 2015) (Satija et al.,

2015) (Butler et al., 2018). For single-dataset analyses of NTC-Rag1-/-, sgPrkar1a-Rag1-/-, sgPrkar1a-Nu/Nu and sgPrkar1a-

C57BL/6 tumor micro-environments, variable genes were identified using the FindVariableGenes Seurat function, in which genes

are split into 20 equal-width bins by mean expression, then filtered for those with high dispersion (>0.5 standard deviations from

the average dispersion (log variance/mean ratio) within a bin) and high/low mean expression (>8 or <0.0125) (Satija et al., 2015).

The variance in these variably expressed genes was then assessed by principal components analysis (PCA), and principal

components were dimensionality reduced by t-distributed stochastic neighbor embedding (t-SNE) (van der Maaten and Hinton,

2008) via the Barnes-Hut approximation implementation (van der Maaten, 2014) or PHATE (Moon et al., 2017). Next, cell clustering

was performed using shared nearest neighbor (SNN) algorithm (Waltman and van Eck, 2013). The t-SNE, PHATE and SNN tech-

niques were each performed using the number of principal components that explains majority of the data variance, based on the

‘‘elbow’’ in plots of the partial variance described by each PC (Not shown). For the integrated analysis of sg-Prkar1a and sg-NTC-

treated Rag1-/- cell samples, dimensionality reduction was performed using canonical correlation analysis (CCA). The correlation

Cell Systems 8, 136–151.e1–e7, February 27, 2019 e6

Page 24: Convergent Identification and Interrogation of … › fanlab › Rong_Fan_Group...Article Convergent Identification and Interrogation of Tumor-Intrinsic Factors that Modulate Cancer

structure of the two datasets was then aligned using the dynamic time warping method, as described previously (Butler et al., 2018).

The aligned, integrated data were visualized by t-SNE and clustered using SNN, as before.

Cluster Analysis of scRNA-seq DataCluster-specific marker genes were selected by differential expression analyses for each cluster, relative to all other clusters, using

the Model-based Analysis of Single-cell Transcriptomics (MAST) (Finak et al., 2015) with the Surat R package (Satija et al., 2015),

filtering the genes to include those with an adjusted-p < 0.01 and within the top absolute fold-change (95th percentile). The

cluster-specific markers were then used to compare each cluster population with known immune cells from the Immunological

Genome Project data (ImmGen version 1; GEO dataset: GSE15907) (Heng et al., 2008). More specifically, the marker expression

was averaged among cluster/cell type, log-normalized (as described above), and compared between datasets by Kendall

correlation analysis (fast approximationmethod by pcaPPRpackage), after which, the correlation coefficients from each comparison

were scaled and visualized by heatmap (pheatmap R package). In addition, this correlation heatmap data were grouped by ImmGen

(Heng et al., 2008) cell categories, and the specific cell types were arranged within these categories by unsupervised hierarchical

clustering.

Tumor cell markers were determined by comparing the NTC-treatedRag1-/- tumor cell bulk RNA-seq transcriptional data with that

of the ImmGen innate immune populations, using log-normalized, group-mean, gene-level expression values. The tumor markers

were selected as the top 50 genes, based on the tumor/immune cell transcriptional fold-change, that were detected in each

scRNA-seq dataset. Four representative tumor cell markers were chosen, after filtering for the genes, based on the detected

expression in single-cell data clusters (>= 0.50 detection) and by cluster-specific fold-change (fold-change >= 1.5). Note that these

representative genes were among the top 20 tumor markers, and were only selected for t-SNE/PHATE visualization purposes.

Lastly, all differential expression analyses were performed by MAST (Finak et al., 2015).

Flow CytometryFlow cytometry analysis of immune populations was performed using a BD FACSAria cytometer. The cell surface antibody staining

protocol for the analysis involved blocking Fc receptors with TruStain FcX anti-mouse CD16/32 for 10 min at 4C (1ug/1e6 cells

in 100uL staining buffer (0.5% (w/v) BSA, 2mM EDTA in PBS, pH 7.4); Biolegend), washing cells, staining cells with the appropriate

fluorophore-conjugated antibodies for 30 minutes at 4C (1uL antibody/1e6 cells in a total volume of 100uL staining buffer), washing

cells, and then resuspending in PBS, prior to FACS. Antibodies are listed in Key Resources Table.

Blinding StatementInvestigators were blinded for tumor sizemeasurement and histology scoring, but not blinded for screen data analysis, proteomics or

RNA-seq.

QUANTIFICATION AND STATISTICAL ANALYSIS

Standard data analysis (Non-NGS) were performed using regular statistics, where NGS data were analyzed with specific pipelines

described in methods under separate sub-headlines. Data comparison between two groups was performed using a two-tailed un-

paired t-test or non-parametric Wilcox test. p values and statistical significance were estimated for all analyses. Prism (GraphPad

Software Inc.) and RStudio were used for these analyses.

DATA AND SOFTWARE AVAILABILITY

Reagent AvailabilityReagents including plasmids and libraries are available to the academic community through Addgene. Other reagents used in this

study are described in Key Resources Table.

Code AvailabilityAll custom scripts used to process and analyze the data are available upon reasonable request.

Data Availability and AccessionsPhosphoproteomics data were deposited into PRIDE database under PRIDE: PXD011837. The mRNA-seq data were deposited into

NCBI SRA under GEO: PRJNA506858. Single-cell RNA-seq were deposited into GEO under GEO: PRJNA509929.

e7 Cell Systems 8, 136–151.e1–e7, February 27, 2019