13
Small Molecule Therapeutics Dissecting Therapeutic Resistance to ERK Inhibition Sharda Jha 1 , Erick J. Morris 1 , Alan Hruza 2 , My Sam Mansueto 1 , Gottfried K. Schroeder 1 , Jaren Arbanas 1 , Daniel McMasters 1 , Clifford R. Restaino 1 , Priya Dayananth 2 , Stuart Black 2 , Nathaniel L. Elsen 2 , Anthony Mannarino 1 , Alan Cooper 2 , Stephen Fawell 1 , Leigh Zawel 1 , Lata Jayaraman 1 , and Ahmed A. Samatar 1 Abstract The MAPK pathway is frequently activated in many human cancers, particularly melanomas. A single-nucleotide mutation in BRAF resulting in the substitution of glutamic acid for valine (V 600E ) causes constitutive activation of the downstream MAPK pathway. Selective BRAF and MEK inhibitor therapies have dem- onstrated remarkable antitumor responses in BRAF V600E -mutant melanoma patients. However, initial tumor shrinkage is transient and the vast majority of patients develop resistance. We previously reported that SCH772984, an ERK 1/2 inhibitor, effectively sup- pressed MAPK pathway signaling and cell proliferation in BRAF, MEK, and concurrent BRAF/MEK inhibitor-resistant tumor models. ERK inhibitors are currently being evaluated in clinical trials and, in anticipation of the likelihood of clinical resistance, we sought to prospectively model acquired resistance to SCH772984. Our data show that long-term exposure of cells to SCH772984 leads to acquired resistance, attributable to a mutation of glycine to aspartic acid (G 186D ) in the DFG motif of ERK1. Structural and biophysical studies demonstrated specic defects in SCH772984 binding to mutant ERK. Taken together, these studies describe the interaction of SCH772984 with ERK and identify a novel mechanism of ERK inhibitor resistance through mutation of a single residue within the DFG motif. Mol Cancer Ther; 15(4); 54859. Ó2016 AACR. Introduction The MAPK pathway is a key driver of tumorigenesis, and mutations in the RAS and BRAF genes are among the most common in human cancers. Mutations in RAS occur in a variety of tumor types and are important in driving MAPK and PI3K pathway activation (1, 2). Mutations in BRAF occur in approxi- mately 50% of metastatic melanomas, frequently at the V 600 residue, making this the most common oncogenic event in this disease and the primary contributor to MEK and ERK activation (3). Interestingly, recent whole-exome studies have shown that most melanomas harbor at least one mutation that would be predicted to activate the MAPK pathway (4). The clinical management of melanomas has rapidly evolved with the discovery and development of potent and selective inhibitors of the MAPK pathway. Recent clinical trials with RAF inhibitors (vemurafenib and dabrafenib) and to a lesser extent MEK inhibitors (trametinib), have demonstrated rapid, robust, and reproducible objective antitumor responses in a majority of treated melanoma patients. However, the durability of response is limited by the acquisition of drug resistance, result- ing in disease progression (59). Multiple mechanisms have been reported for BRAF and MEK inhibitor resistance leading to MAPK pathway reactivation. These include upregulation of receptor tyrosine kinase (RTK) signaling, amplication of BRAF, alternative splicing of mutant BRAF, emergence of muta- tions in RAS or MEK occurring concurrently with mutant BRAF, activation of EGFRSFKSTAT3 signaling, and activation of NRAS signaling (1018). MEK inhibitor resistance has been reported to arise as a result of mutations in the allosteric drug- binding pocket or amino-terminal negative inhibitory domain (19, 20). Relative contribution of these varied mechanisms to disease progression in patients is presently unclear although most appear to lead to reactivation of ERK. Combined BRAF/ MEK inhibition as a strategy to mitigate acquired BRAF inhib- itor resistance is clinically effective and was recently approved by the FDA for BRAF-mutant melanoma. This approach, while successful in delaying tumor progression, still does not alleviate acquired drug resistance, suggesting escape routes that may or may not be MAPK pathway dependent (21, 22). Indeed MAPK pathway-independent mechanisms have already been reported in patients in BRAF inhibitor resistance (12, 22). 1 Early Development and Discovery Sciences, Merck Research Labo- ratories, Boston, Massachusetts. 2 Early Development and Discovery Sciences, Merck Research Laboratories, Kenilworth, New Jersey. Note: Supplementary data for this article are available at Molecular Cancer Therapeutics Online (http://mct.aacrjournals.org/). Current address for E.J. Morris: Novartis Oncology Translational Research, Cam- bridge, MA; current address for A. Cooper: Cooper Medchem Consulting, LLC, NJ; current address for S. Fawell: Oncology iScience, Astrazeneca, Waltham, MA; current address for L. Zawel: Centers for Therapeutic Innovation, Pzer, Boston, MA; and current address for L. Jayaraman: Cerulean Pharma Inc, Waltham, MA; current address for A.A. Samatar: TheraMet Biosciences, Princeton Junction, NJ. S. Jha and E.J. Morris are co-rst authors and contributed equally to this article. A.A. Samatar and L. Jayaraman are co-senior authors and contributed equally to this article. Corresponding Authors: Ahmed A. Samatar, TheraMet Biosciences, 6 Jacob Drive, Princeton, NJ 08550. Phone: 609-902-6387; Fax: 609-716-6358; E-mail: [email protected]; and Lata Jayaraman, Cerulean Pharma Inc, 35 Gatehouse Drive, Waltham, MA 02451. Phone: 781-209-6384; Fax: 844-894-2407; E-mail: [email protected] doi: 10.1158/1535-7163.MCT-15-0172 Ó2016 American Association for Cancer Research. Molecular Cancer Therapeutics Mol Cancer Ther; 15(4) April 2016 548 on January 18, 2021. © 2016 American Association for Cancer Research. mct.aacrjournals.org Downloaded from Published OnlineFirst February 1, 2016; DOI: 10.1158/1535-7163.MCT-15-0172

Dissecting Therapeutic Resistance to ERK InhibitionJaren Arbanas1, Daniel McMasters1, Clifford R. Restaino1, Priya Dayananth2, Stuart Black2, Nathaniel L. Elsen 2 , Anthony Mannarino

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Dissecting Therapeutic Resistance to ERK InhibitionJaren Arbanas1, Daniel McMasters1, Clifford R. Restaino1, Priya Dayananth2, Stuart Black2, Nathaniel L. Elsen 2 , Anthony Mannarino

Small Molecule Therapeutics

Dissecting Therapeutic Resistance to ERKInhibitionSharda Jha1, Erick J. Morris1, Alan Hruza2, My Sam Mansueto1, Gottfried K. Schroeder1,Jaren Arbanas1, Daniel McMasters1, Clifford R. Restaino1, Priya Dayananth2, Stuart Black2,Nathaniel L. Elsen2, Anthony Mannarino1, Alan Cooper2, Stephen Fawell1, Leigh Zawel1,Lata Jayaraman1, and Ahmed A. Samatar1

Abstract

The MAPK pathway is frequently activated in many humancancers, particularly melanomas. A single-nucleotide mutation inBRAF resulting in the substitution of glutamic acid for valine(V600E) causes constitutive activation of the downstream MAPKpathway. Selective BRAF and MEK inhibitor therapies have dem-onstrated remarkable antitumor responses in BRAFV600E-mutantmelanoma patients. However, initial tumor shrinkage is transientand the vast majority of patients develop resistance. We previouslyreported that SCH772984, an ERK 1/2 inhibitor, effectively sup-pressed MAPK pathway signaling and cell proliferation in BRAF,MEK, and concurrent BRAF/MEK inhibitor-resistant tumormodels.

ERK inhibitors are currently being evaluated in clinical trials and, inanticipation of the likelihood of clinical resistance, we sought toprospectively model acquired resistance to SCH772984. Our datashow that long-term exposure of cells to SCH772984 leads toacquired resistance, attributable to amutation of glycine to asparticacid (G186D) in the DFGmotif of ERK1. Structural and biophysicalstudies demonstrated specific defects in SCH772984 binding tomutant ERK. Taken together, these studies describe the interactionof SCH772984 with ERK and identify a novel mechanism of ERKinhibitor resistance throughmutation of a single residuewithin theDFG motif. Mol Cancer Ther; 15(4); 548–59. �2016 AACR.

IntroductionThe MAPK pathway is a key driver of tumorigenesis, and

mutations in the RAS and BRAF genes are among the mostcommon in human cancers. Mutations in RAS occur in a varietyof tumor types and are important in driving MAPK and PI3Kpathway activation (1, 2). Mutations in BRAF occur in approxi-mately 50% of metastatic melanomas, frequently at the V600

residue, making this the most common oncogenic event in thisdisease and the primary contributor to MEK and ERK activation

(3). Interestingly, recent whole-exome studies have shown thatmost melanomas harbor at least one mutation that would bepredicted to activate the MAPK pathway (4).

The clinical management of melanomas has rapidly evolvedwith the discovery and development of potent and selectiveinhibitors of the MAPK pathway. Recent clinical trials with RAFinhibitors (vemurafenib and dabrafenib) and to a lesser extentMEK inhibitors (trametinib), have demonstrated rapid, robust,and reproducible objective antitumor responses in a majorityof treated melanoma patients. However, the durability ofresponse is limited by the acquisition of drug resistance, result-ing in disease progression (5–9). Multiple mechanisms havebeen reported for BRAF and MEK inhibitor resistance leading toMAPK pathway reactivation. These include upregulation ofreceptor tyrosine kinase (RTK) signaling, amplification ofBRAF, alternative splicing of mutant BRAF, emergence of muta-tions in RAS or MEK occurring concurrently with mutant BRAF,activation of EGFR–SFK–STAT3 signaling, and activation ofNRAS signaling (10–18). MEK inhibitor resistance has beenreported to arise as a result of mutations in the allosteric drug-binding pocket or amino-terminal negative inhibitory domain(19, 20). Relative contribution of these varied mechanisms todisease progression in patients is presently unclear althoughmost appear to lead to reactivation of ERK. Combined BRAF/MEK inhibition as a strategy to mitigate acquired BRAF inhib-itor resistance is clinically effective and was recently approvedby the FDA for BRAF-mutant melanoma. This approach, whilesuccessful in delaying tumor progression, still does not alleviateacquired drug resistance, suggesting escape routes that may ormay not be MAPK pathway dependent (21, 22). Indeed MAPKpathway-independent mechanisms have already been reportedin patients in BRAF inhibitor resistance (12, 22).

1Early Development and Discovery Sciences, Merck Research Labo-ratories, Boston, Massachusetts. 2Early Development and DiscoverySciences, Merck Research Laboratories, Kenilworth, New Jersey.

Note: Supplementary data for this article are available at Molecular CancerTherapeutics Online (http://mct.aacrjournals.org/).

Current address for E.J. Morris: Novartis Oncology Translational Research, Cam-bridge, MA; current address for A. Cooper: CooperMedchemConsulting, LLC, NJ;current address for S. Fawell: Oncology iScience, Astrazeneca, Waltham, MA;current address for L. Zawel: Centers for Therapeutic Innovation, Pfizer, Boston,MA; and current address for L. Jayaraman: Cerulean Pharma Inc, Waltham, MA;current address for A.A. Samatar: TheraMet Biosciences, Princeton Junction, NJ.

S. Jha and E.J. Morris are co-first authors and contributed equally to this article.

A.A. Samatar and L. Jayaraman are co-senior authors and contributed equally tothis article.

Corresponding Authors: Ahmed A. Samatar, TheraMet Biosciences, 6 JacobDrive, Princeton, NJ 08550. Phone: 609-902-6387; Fax: 609-716-6358; E-mail:[email protected]; and Lata Jayaraman, Cerulean Pharma Inc,35 Gatehouse Drive, Waltham, MA 02451. Phone: 781-209-6384;Fax: 844-894-2407; E-mail: [email protected]

doi: 10.1158/1535-7163.MCT-15-0172

�2016 American Association for Cancer Research.

MolecularCancerTherapeutics

Mol Cancer Ther; 15(4) April 2016548

on January 18, 2021. © 2016 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst February 1, 2016; DOI: 10.1158/1535-7163.MCT-15-0172

Page 2: Dissecting Therapeutic Resistance to ERK InhibitionJaren Arbanas1, Daniel McMasters1, Clifford R. Restaino1, Priya Dayananth2, Stuart Black2, Nathaniel L. Elsen 2 , Anthony Mannarino

We recently reported the discovery and characterization of anovel and specific ERK 1/2 inhibitor SCH772984 with a uniquemechanism of action (23). SCH772984 not only inhibits ERKenzymatic activity effectively but also its phosphorylation byMEK. More recently, SCH772984 was shown to be a potentinhibitor of a majority of melanoma cell lines irrespective of theirBRAF mutational status and their sensitivity to BRAF inhibitors(24). Although the clinical efficacy of ERK inhibitors has yet to bedemonstrated, experience with small-molecule kinase inhibitorssuch as BRAF or MEK inhibitors suggests that even when highresponse rates are elicited, tumor cells still retain a strong pro-pensity to develop resistance. Anticipating the development ofresistance to ERK inhibitors in the clinic, we undertook a pro-spective study to investigate the mechanism(s) of resistance toSCH772984. We have also reported previously that SCH772984demonstrated a more limited efficacy profile in RAS-mutant celllines as compared with BRAF-mutant cell lines (23). The resultspresented in this article seek to extend those observations and tounderstand ERK inhibitor resistance specifically in the context of aKRAS-mutant cell line, not a BRAF-mutant cell line. This impetusto focus on a KRAS-mutant line was due to the lower sensitivity ofERK inhibitors in this context (although we acknowledge that theresistance profile generated in a BRAF-mutant cell line couldutilize different mechanisms).

Materials and MethodsTotal RNA extraction and cDNA synthesis

Parental or resistant HCT-116 cells untreated or treated withSCH772984were collected and extraction of RNAwas done usingthe RNeasy Mini Kit (Qiagen), as per the manufacturer's instruc-tions. RNA samples were eluted in 50 mL of nuclease-free water.cDNA was generated from 2 mg of total RNA by using a murineMLV Reverse Transcriptase First-Strand cDNA Synthesis Kit(Applied Biosystems).

Genomic DNA sequencingPrimer extension sequencing was performed by GENEWIZ, Inc

on an Applied Biosystems BigDye version 3.1. The reactions weresubsequently run on Applied Biosystem's 3730xl DNA Analyzerfor sequencing of all coding exons in KRAS, HRAS, NRAS,MAP2K1 (MEK1), MAP2K2 (MEK2), MAPK1 (ERK2), MAPK3(ERK1), or BRAF (coding exon 6 and 11–15).

Real-time PCR for MAPK gene expressionPCRs were performed andmonitored using an ABI Prism 7900

Sequence Detection system (PerkinElmer, Applied Biosystems).The PCR master mix with AmpliTaq Gold DNA polymerase wasused and cDNA samples were analyzed in duplicate. Primers andprobes were used at concentrations of 100 and 125 nmol/L perreaction, respectively. After an initial denaturation stepof 95�Cfor10minutes, the cDNAproductswere amplifiedwith40PCR cycles(denaturation: 95�C for 15 seconds; extension: 60�C for 1 min-ute). For each sample, the Ct value was determined as the cyclenumber at which the fluorescence intensity reached 0.05; thisvalue was chosen after confirming that all curves were in theexponential phase of amplification. Relative expression was cal-culated using the DCt method using the following equations: DCt

(sample) ¼ Ct (target)�Ct (reference); relative quantity ¼ 2�DCt.Differential expression of genes was identified using significanceanalysis. For each cDNA sample, the Ct value of each target

sequence was normalized to the reference genes (GAPDH, tubu-lin, and actin).

Protein overexpression in stable cell linesCells were infected with lentivirus produced from lentiORF

constructs (pLOC vector) expressing either RFP, ERKG186D,ERK1WT, or ERK2 WT. Cells were selected in blasticidin (20mg/mL) for 10 days before Western blotting or testing inproliferation assays.

Western blot analysisCell lysates for Western blotting were made using MPER lysis

buffer (Thermo Scientific) with protease inhibitor cocktail(Roche). Protein samples (20 mg per well) were separated on4% to 20% Tris-HCl gels and then transferred to a nitrocellulosemembrane. Western blot analyses were probed with antibodiesselective for the following proteins: p-ERK1/2 (T202/Y204; CellSignaling Technologies, cat no. 4695), ERK1/2 (Cell SignalingTechnology, cat no. 9107), pRSK (Millipore, cat no. 04-419), RSK(BD Biosciences, cat no. 610226), and actin (Cell SignalingTechnology, cat no. 4967).

Cell Lines and treatmentsParental HCT-116 cell line was obtained from ATCC in June

2013 and authenticated using DNA fingerprinting short tandemrepeat assays. The cell line was then tested for pathogen contam-ination. Parental HCT-116 cells were used within four passagesbefore generating HCT-116–resistant cells. Both parental HCT-116and HCT-116 SCH772984-resistant cells were again tested inFebruary 2014 using fingerprinting short tandem repeat assays andsequenced. Cells weremaintained inDMEMwith 10% FBSwith orwithout inhibitors. Cell proliferation experiments were performedin a 96-well plate format (six replicates), at a plating density of4,000 cells per well. At 24 hours after cell seeding, cells were treatedwith serially diluted inhibitor (10 mmol/L–0.001 mmol/L). Cellswere then quantified on day 5 using ViaLight Luminescence kit(Promega) following the manufacturer's instructions. Data werethen analyzed using the SoftMax pro plate reader/software. Forresistant cell line generation, HCT-116 cells were grown in DMEMwith 10% heat-inactivated FBSmedia in the presence of increasingconcentrations of SCH772984 inhibitor (0.1–2 mmol/L) overapproximately 8 months until resistant cells acquired growthproperties similar to untreated treated parental HCT-116 cells.

Protein purification for crystallizationBL21 (DE3) cellswere transformedwithplasmids coding for rat

wild-type or G167D-mutant ERK2 protein. UnphosphorylatedERK2proteinswere expressed andpurifiedon ice (4�C)as follows:cell pellets were resuspended in 5mL/g of lysis buffer (0.05mol/Lsodium phosphate, pH 8.0, 0.3 mol/L NaCl, 10 mmol/L b-mer-captoethanol, 10,000 U/L benzonase, and 5mL/L of CalbiochemProtease Inhibitor Cocktail Set III), homogenized using a DounceHomogenizer, and then lysed with three passes through a Micro-fluidizer. The lysate was centrifuged at an rmax of 200,000� g for 1hour to remove cell debris. The supernatant was applied to aQiagen Ni-NTA superflow agarose column (40 mL) pre-equili-brated with Buffer A (lysis buffer, pH 7.5, without benzonase andprotease inhibitors). The column was then washed as follows: 1column volume (CV) of Buffer A, 3 CV of Buffer B (Buffer A with25mmol/L imidazole), 2 CVof Buffer B containing 1mol/LNaCl,and 2 CV of Buffer A with 45mmol/L imidazole. The columnwas

ERK Inhibitor Resistance

www.aacrjournals.org Mol Cancer Ther; 15(4) April 2016 549

on January 18, 2021. © 2016 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst February 1, 2016; DOI: 10.1158/1535-7163.MCT-15-0172

Page 3: Dissecting Therapeutic Resistance to ERK InhibitionJaren Arbanas1, Daniel McMasters1, Clifford R. Restaino1, Priya Dayananth2, Stuart Black2, Nathaniel L. Elsen 2 , Anthony Mannarino

then eluted with Buffer A containing 250mmol/L imidazole. Theeluted pool was dialyzed (three exchanges) against 4 L of MonoQBuffer (25 mmol/L Tris-Cl, pH 7.8, 0.05 mol/L NaCl, 10% (v/v)glycerol, 1 mmol/L EDTA, and 5 mmol/L DTT). The dialyzed Ni-NTA pool was applied to an HR 16/10 MonoQ column at 1 mL/minute. After loading, the columnwas washedwith 1 CV (20mL)of MonoQ Buffer and then eluted with a linear sodium chloridegradient (600mL) ofMonoQBuffer (0–0.5mol/LNaCl). Thefirstof twomajor protein peaks representing unphosphorylated ERK2(eluted at �0.15 mol/L NaCl) were pooled and concentrated to15.8 mg/mL using spin concentrators (10,000 MWCO, Amicon).

Protein crystallization, data collection, and structuredetermination

Crystals of rat ERK2 [(MAH6) 1–358 or G167D mutant] weregenerated using the hanging dropmethod. Each drop consisted of2 mL of approximately 15 to 20 mg /mL ERK2 protein (precom-plexed with 1 mmol/L olomoucine) and 2 mL crystallizationsolution (50mmol/LMES, pH6.4, 2.1mol/L ammonium sulfate,5% PEG 400, 1% glycerol, 0.5% DMSO, and 1 mmol/L TCEP).The 4-mL drop was seeded on a glass slide and inverted over areservoir containing 0.7 mL of the same crystallization solutionand incubated at 4�C. In 1 to 2 weeks, crystals grew to a sufficientsize for use in compound-soaking experiments. A crystal wastransferred into 300 mL of a soaking buffer (50 mmol/L MES,pH 6.4, 2.1 mol/L ammonium sulfate, 5% PEG 400, 1% glycerol,0.5% DMSO, 1 mmol/L TCEP) containing 500 mmol/L of SCH772984. After 10 days, the crystal was transferred into a cryogenicbuffer (100 mmol/L MES, pH 6.4, 1.8 mol/L ammonium sulfate,5% PEG 400, and 27.5% glycerol) for 15 seconds, and thenscooped into a loop and flash frozen in liquid nitrogen. Diffrac-tion data were obtained at APS IMCA beam line using a MARM165 CCD for wild-type ERK2 and a Dectris 6M for ERK2G167D.Wild-type ERK2 diffraction data were processed and refined usingHKL2000 and CNX/AutoBuster, while AutoProc/XDS and Auto-buster were used for the mutant diffraction data. The model forSCH772984 was constructed using Grade/Rhofit and positionedin the difference density for the compound. The ERK2 modelswere visualized and refined tofit the experimental electrondensitymap in Coot (25). Data collection and refinement statistics aresummarized in Table 1.

Purification of inactive ERK enzymesDNA encoding human wild-type ERK1 enzyme (Genbank

nucleotide reference sequence NM_002746) with eight addition-al N-terminal Histidine residues (His8) was synthesized (Gene-wiz, Inc.) and cloned into bacterial expression vector pGEX6p-3using the SalI and NotI sites. This construct, when translated,produced an intact recombinant ERK1 enzyme with two N0-terminal affinity tags, GST and His8. ERK1

G186D was cloned fromthe wild-type template using aQuikChange kit (Stratagene). Bothconstructs were transformed into E. coli and induced with 0.5mmol/L IPTG at OD600 nm ¼ �0.8, and overexpression wasallowed to proceed for approximately 16 hours. The inductiontemperature for WT ERK1 enzyme and ERK1G186D mutant was25�Cand16�C, respectively. The inactive formsofWTandmutantERK1 enzyme were each purified as follows: cells from a 10-Lculture were resuspended with 400 mL of lysis buffer (300mmol/L NaCl, 50 mmol/L Tris, pH 8, 10% glycerol, 0.1% TritonX100, 2 mmol/L DTT, 0.5 mmol/L EDTA, and Roche cocktailinhibitor tablets), and lysed using either a microfluidizer or

sonicator. Cell lysates were clarified by ultracentrifugation andthe cleared supernatant was batch bound (30 minutes at 4�C) to15 mL of glutathione beads (GE, cat no. 17-5132-02) precali-brated with wash buffer (1 mol/L NaCl, 5% glycerol, 25 mmol/LTris, pH 7.5, 0.25mmol/L EDTA, and 1mmol/L DTT). The lysate-beads mixture was poured into a gravity column and unboundproteins were washed with approximately 500 mL of wash bufferor until no protein could be detected from the flow through withBradford reagent. The target proteinwas elutedwith elutionbuffercontaining 10 mmol/L reduced glutathione (50 mmol/L HEPES,pH 7.3, 10 mmol/L reduced glutathione, 5% glycerol, 1 mmol/LDTT). Target protein was confirmed on SDS-PAGE prior to pool-ing. Pooled fractions were concentrated to approximately 5 mL(�1.5% of SEC bed volume) and fractionated over a HiLoad 26/600 Superdex 200 Size Exclusion Column (SEC buffer: 25 mmol/L Tris, pH7.2, 0.26mol/LNaCl, 10%glycerol, 1mmol/LDTT, and0.1 mmol/L EDTA). Fractions containing target protein werepooled and concentrated for storage. Final protein yield with>98% purity from a 10-L culture was approximately 20 mg. ERK2wild-type and mutant were similarly generated.

Preparation of activated ERK enzymesThe active form of ERK enzyme used in this manuscript was

generated through in vitro phosphorylation using a constitutivelyactiveMEK1 enzyme. Specifically, purified inactive ERK and activeMEK1were incubated at approximately 1:1molar ratio at approx-imately 20 mmol/L complex concentration in the presence of 50mmol/L HEPES, pH 7.3, 5% glycerol, 1 mmol/L DTT, 20mmol/LMgCl2, and 4 mmol/L ATP for 2 hours at room temperature,followed by overnight incubation at 4�C. MEK1 was successfullyremoved from GST-ERK dimers using a Superdex 200 SECcolumn.

Activated ERK inhibition assays for IC50 determinationThe IC50 of SCH772984 for the activated forms of ERK1 and

ERK2, and their respective DFG single amino acid substitution

Table 1. Statistics for crystal data collection and refinement

Data collection and refinement statistics

CrystalWT rat-ERK2SCH772984

G167D rat-ERK2SCH772984

PDB code 5HD4 5HD7

Data collectionWavelength 1.00001 1.54178Space group P21212 P21212Unit cell parameters (Ǻ, �) a ¼ 71.5, b ¼ 91.4,

c ¼ 63.2a ¼ 71.5, b ¼ 91.4,c ¼ 63.2

a,b,g ¼ 90.0� a,b,g ¼ 90.0�

Resolution range (Ǻ) 23.07 – 1.45 28.03 – 1.69(1.50 – 1.45) (1.70 – 1.69)

Multiplicity 4.4 (3.9) 9.7 (10.0)Unique reflections 72,235 44,558Rmerge (%) 4.9 (52.3) 5.1 (60.4)Completeness (%) 97.8 (94.7) 95.7 (91.4)I/s(I) 24.0 (2.5) 27.6 (3.7)

RefinementRwork/Rfree (%) 17.6/19.9 18.1/20.7No. of protein atoms 5,673 5,812No. of water 276 274No. of heteroatoms 390 377R.m.s.d bonds (Ǻ) 0.009 0.010R.m.s.d angles (�) 1.0 1.0

Jha et al.

Mol Cancer Ther; 15(4) April 2016 Molecular Cancer Therapeutics550

on January 18, 2021. © 2016 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst February 1, 2016; DOI: 10.1158/1535-7163.MCT-15-0172

Page 4: Dissecting Therapeutic Resistance to ERK InhibitionJaren Arbanas1, Daniel McMasters1, Clifford R. Restaino1, Priya Dayananth2, Stuart Black2, Nathaniel L. Elsen 2 , Anthony Mannarino

mutants (ERK1-G186D and ERK2-G169D), were determined using aLANCE Ultra ULight assay (PerkinElmer). IC50 values were deter-mined from a 15-point 3-fold dilution titration range (wild-typeERK, 5 mmol/L–1.05 pmol/L; mutant ERK, 100 mmol/L–21pmol/L) using the following procedure: compound was dis-pensed into each well of a black nonbinding surface 384-wellmicroplate (Corning, cat no. 3820), followed by the addition of7.5 mL of 1X kinase buffer (50mmol/L HEPES, pH 7.3, 0.01%Brij35, 10 mmol/L MgCl2, and 1 mmol/L DTT) containing ATP andenzyme. After preincubation (30 minutes), each reaction wasinitiated by the addition of 2.5 mL 1X kinase buffer containingMBP-ULight labeled peptide (PerkinElmer, cat no. TRF0109-M).The final reaction in each well (10 mL) contained 125 pmol/L ERKenzyme, 50 nmol/LMBP-ULight peptide, and 45 mmol/L ATP and<1% DMSO. Phosphorylation reactions were incubated for 90minutes (ERK1 and ERK1G186D) or 60 minutes (ERK2 andERK2G169D). Reactions were immediately quenched by the addi-tion of 10 mL of 1� quench/detection buffer (PerkinElmer, cat no.CR97-100) containing 4 nmol/L of anti-pY-MBP antibody (Per-kinElmer, cat no. TRF0201-M). Following incubation with detec-tion reagents (60 minutes), reaction plates were read on a Perki-nElmer EnVision plate reader using a standard TR-FRET protocol.IC50 values were determined by a 4-parameter robust fit of log10inhibitor concentrations versus TR-FRET ratio values using Prism(GraphPad Software, version 6.04).

Surface plasmon resonanceSurface plasmon resonance (SPR) experiments were conducted

using a Biacore T200 system (GE Healthcare). Recombinant GST-tagged ERK proteins were diluted to 0.5 mmol/L in 1� Biacorerunning bufferHBS-EPþ (GE, cat no. BR-1006-69) containing 3%DMSO and captured on the surface of a preactivated Series S CM5Biacore chip (GE, cat no. BR-1005-30) using a GST Capture Kit(GE, cat no. BR-1002-23), according to themanufacturer's instruc-tions. Briefly, anti-GST antibody was immobilized onto the chipby amine coupling, followed by ethanolamine injection to blockany remaining active groups. To condition the chip, cycles of GSTinjection, followed by a regeneration step, were utilized to blockany high-affinity GST-binding sites. Final anti-GST antibodyimmobilization levels were typically approximately 7,500 RU.For kinetics experiments conducted in high-performance mode,surface regeneration was followed by a protein capture stepbetween each inhibitor injection. Surface regeneration solution(10mmol/L Glycine-HCl, pH 1.8) was injected twice over all fourCM5 chip flow cells for 120 seconds at 100 mL/minute. Regen-eration was followed by injection of protein (0.5 mmol/L) for 120seconds at 5 mL/minute across the designated flow cell (2, 3, or 4),with injection of GST (same conditions) over flow cell 1 as areference. Final ERK protein capture levels were approximately1,250� 150 RU. To determine binding kinetics of the inhibitor toERK proteins, SCH772984was then passed over all four flow cellsat 50 mL/minute for 180 seconds, followed by a dissociationperiod of 300 to 1,000 seconds. After the dissociation period,the chipwas regenerated back to baseline as described above. Thiswas repeated for each inhibitor injection over nine injections of a2/3-fold dilution titration range starting between 25 nmol/L and5,000 nmol/L that included a single replicate. Binding data weresolvent corrected and double reference subtractedwith a referenceflow cell and blank injections to account for any bulk refractiveindex changes or any small amount of injection noise, baselinedrift, and nonspecific binding. Quality control and evaluation

were performed using the Biacore T200 Evaluation Software. Theentire dataset was fit to either a 1:1 interaction model to give kon,koff, and KD values or a steady-state binding affinity modelyielding KD. For inactive wild-type ERK1 (capture level �1,250� 150 RU) and inactive wild-type ERK2 (performed in previousexperiments with capture level�3,500 RU), binding experimentswere conducted in single-cycle kinetics mode. For inactive wild-type ERK1, inhibitor injections were performed at 50 mL/minutewith a contact time of 420 seconds and a dissociation time of2,400 seconds. Following five sequential injections of a 3-folddilution titration range of SCH772984 starting at 50 nmol/L,surface regeneration was performed as described above. Forinactive wild-type ERK2, inhibitor injections (5 sequential injec-tions of a 3-fold dilution titration range of SCH772984 starting at27 nmol/L) were performed at 30 mL/minute to accommodate alonger contact time of 600 seconds and a dissociation time of3,600 seconds. Both datasets were fit to a 1:1 interaction modelyielding kon, koff, and KD values. All assays were run at 25�C in 1�Biacore running buffer HBS-EPþ (GE, cat no. BR-1006-69) con-taining 3% DMSO.

ResultsAcquired resistance to small-molecule kinase inhibitors often

occurs after prolonged treatment or exposure to drug (26). Wehave previously published on the discovery of a novel ERK1/2inhibitor SCH772984 (23). Tomodel the acquisitionof resistanceto this ERK inhibitor, we used the KRAS-mutant HCT-116 colo-rectal cell line (termed HCT-116P), which is sensitive toSCH772984 (23). To generate drug resistance, cells were seriallypassaged in the presence of increasing concentrations ofSCH772984 over a period of 4 months. Although the vast major-ity of cells were rapidly killed in the presence of this inhibitor, cellline pools eventually emerged (termedHCT-116R pool) that wereresistant to a high dose of drug as evidenced by robust prolifer-ation in the presence of SCH772984 (2 mmol/L). Proliferationdose–response curves showed that the resistant cell pool wasunaffected by up to 10 mmol/L SCH772984 (Fig. 1A). Severalsingle-cell clones were isolated from the resistant pool and celllines established and cultured in the presence of 2 mmol/L ofSCH772984. One resistant clonal cell line (termed HCT-116R)was selected for the series of studies described below to charac-terize the mechanism of resistance. Proliferation assays showedthat HCT-116R was not sensitive to high concentrations ofSCH772984 (Fig. 1B). IncuCyte analysis confirmed that prolifer-ation rates of HCT116R cells were comparable with parental cells(HCT116P; Fig. 1C andD). Proliferation ofHCT116R cells was notsignificantly affected by the presence of high concentrations ofSCH772984. When grown in the absence of SCH772984 for aperiod of 8 weeks, HCT-116R cells maintained their SCH772984-resistant phenotype, indicating a stable mechanism of resistance(data not shown).

Next, we investigated whether acquisition of drug resistancewas accompanied by MAPK pathway reactivation. SCH772984not only inhibits the intrinsic kinase activity of ERK but alsoblocks MEK-mediated ERK phosphorylation (23). Phospho-ERK,which is abolished upon compound treatment of HCT-116P cells,is reestablished in resistant cells to levels similar to untreatedparental cells. Phospho-RSK, a downstream substrate of ERK, wasreduced in parental cells upon SCH772984 treatment. Baselinelevels of phospho-RSK can be clearly seen in the resistant cells

ERK Inhibitor Resistance

www.aacrjournals.org Mol Cancer Ther; 15(4) April 2016 551

on January 18, 2021. © 2016 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst February 1, 2016; DOI: 10.1158/1535-7163.MCT-15-0172

Page 5: Dissecting Therapeutic Resistance to ERK InhibitionJaren Arbanas1, Daniel McMasters1, Clifford R. Restaino1, Priya Dayananth2, Stuart Black2, Nathaniel L. Elsen 2 , Anthony Mannarino

although the levels are lower than that seen with parental cells;importantly, these levels are unresponsive to addition ofSCH772984 (Fig. 2A). To compare downstream pathway activitybetween parental and resistant cells in the presence ofSCH772984, we assessed the transcriptional output from a selectpanel of genes that have been well documented as MAPK-respon-sive genes in BRAF-mutant melanoma (Fig. 2B). Although theoverall output was lower in the resistant cells, we observed cleardifferences between the parental and resistant cell lines inresponse to SCH772984 treatment. In each case, although thetranscriptional readout in the parental cell line decreased stronglyupon drug treatment, the same readout in the resistant lineshowed little to no change. These data suggest that MAPK down-stream transcriptional activity in the resistant line is no longersensitive to ERK inhibition. To distinguish between compound-specific and pathway-specific resistance, we tested the effect of apreviously described ATP-competitive ERK inhibitor and toolcompound, VTX-11e. As shown in Fig. 2C, sensitivity of bothparental and resistant HCT-116 cells to VTX-11e is essentiallysimilar (IC50 of 0.45 mmol/L and 2 mmol/L, respectively). We alsoevaluated whether the resistant cells still retain sensitivity toinhibition of MEK, the direct upstream kinase of ERK. To thisend,we compared the effect of theMEK inhibitorGSK1120212onparental and resistant cells. As shown in Fig. 2D, GSK1120212is equipotent at inhibiting pERK in both cell lines (IC50 of 5.8nmol/L in parental cells vs. 1.4 nmol/L in resistant cells). Takentogether, these data are consistent with the reactivation of theMAPK pathway in SCH772984-resistant cells, albeit at a loweroverall level compared with parental cells; this reactivation, whileinsensitive to SCH772984, can be abrogated by a MEK inhibitor.

In an effort to identify themechanism(s) of acquired resistanceto SCH772984, the coding regions of RAS, RAF, MEK1, MEK2,ERK1, and ERK2 from 15 single-cell–derived clones includingHCT-116R were sequenced. Sequence analysis showed that inevery clone sequenced, SCH772984 resistance is associated withan acquiredmutation inERK1 in theDFGactivation segmentwiththe substitution of glycine186 to aspartic acid. Sequencing traces(with the mutant allelic frequencies high-lighted) for the clonechosen for our studies (HCT-116R) are shown in the Supplemen-tary Fig. S1. Sequencing of parental HCT-116 cells did not detectcells carrying ERK1G186Dmutation preexisting in the population.Of note, we also observed ERK2 G169D mutation in cell lines thatwere made resistant to the ERK inhibitor MK8353 (data notshown). MK8353, which is in the same structural class asSCH772984, is in phase I clinical trials.

To confirm that the G186D mutation did indeed confer resis-tance, we overexpressed ERK1G186Dmutant, or wild-type ERK1 orERK2proteins, inHCT116P cells. Treatment with SCH772984 didnot significantly inhibit proliferation (Fig. 3A) or phospho-ERK /phospho-RSK levels (Fig. 3B) in ERK1G186D-expressing cells,when compared with cells overexpressing wild-type ERK1 orERK2. These data demonstrate that the presence of mutantERK1G186D is sufficient to confer resistance to SCH772984. Giventhe similarity in the activities of ERK1 and ERK2, and the con-served nature of the DFGmotif residues (Fig. 3C), we predict thatthe analogous mutation in ERK2 might function similarly.

To gain a better understanding of themechanistic consequenceof themutation, the halfmaximal inhibitory concentration (IC50)of SCH772984 for the active forms of both wild-type ERK1 andERK1G186D proteins were measured using an in vitro TR-FRET

BA

DC

Hours

SCH772984 log (mol/L)

Viab

ility

(% c

ontro

l)

Viab

ility

(% c

ontro

l)

–10 –8 –6 –40

50

100

80

Con

fluen

ce v

1.5

(%)

Con

fluen

ce v

1.5

(%)

60

40

2010

100

80

60

40

2010

200 40 60 80 100 120 140Hours

200 40 60 80 100 120 140

100

150

SCH772984 log (mol/L)–10 –8 –6 –4

0

50

100

150

HCT-116R clone HCT-116P

HCT-116R pool HCT-116P

HCT-116 Parental HCT-116 Resistant

10 µmol/L0.001 µmol/L DMSO

3.3 µmol/L 1.1 µmol/L 0.37 µmol/L 0.12 µmol/L 0.04 µmol/L 0.013 µmol/L 0.004 µmol/L 10 µmol/L0.001 µmol/L DMSO

3.3 µmol/L 1.1 µmol/L 0.37 µmol/L 0.12 µmol/L 0.04 µmol/L 0.013 µmol/L 0.004 µmol/L

Figure 1.Activity of SCH772984 in SCH772984-resistant HCT-116 cancer cell line pool (A) aswell as single cell clone (B) demonstrates that both are resistant to SCH772984 athigh concentrations. Viability was determined in 4-day CellTiterGlo assays. The y-axis plots percent viability versus SCH772984 concentration on the x-axis.IncuCyte analysis of real-time growth confluence (imaged every 2 hours) of HCT-116P (C) and HCT-116R cells (D).

Jha et al.

Mol Cancer Ther; 15(4) April 2016 Molecular Cancer Therapeutics552

on January 18, 2021. © 2016 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst February 1, 2016; DOI: 10.1158/1535-7163.MCT-15-0172

Page 6: Dissecting Therapeutic Resistance to ERK InhibitionJaren Arbanas1, Daniel McMasters1, Clifford R. Restaino1, Priya Dayananth2, Stuart Black2, Nathaniel L. Elsen 2 , Anthony Mannarino

A

C

D

pRSK

pERK

ERK

RSK

HCT-116R

SCH772984 2 µmol/L –– + +

HCT-116P

HCT-116RHCT-116P

Actin

B

VTX-11e log (mol/L)

Via

bilit

y (%

con

trol)

–10 –8 –6 –40

50

100

150

tRsk

VTX-11e (2 µmol/L)

MEKi

pRsk+ +– –

HCT116RHCT116P

+ +– –

HCT-116 R HCT-116 P

HCT-116 R HCT-116 P

tERK

pERK

GSK 1120212 log (mol/L)

Via

bilit

y (%

con

trol)

–9 –8 –7 –6 –5 –40

50

100

150

0

0.2

0.4

0.6

0.8

1

1.2

1.4

Nor

mal

ized

mR

NA

exp

ress

ion

DUSP4

4 h

16 h

24 h

00.20.40.60.8

11.21.41.61.8 DUSP6

0

0.2

0.4

0.6

0.8

1

1.2

1.4ETV5

00.20.40.60.8

11.21.41.6

ETV1

00.20.40.60.8

11.21.41.61.8 IL8

0

0.2

0.4

0.6

0.8

1

1.2MYC

SCH772984 1 µmol/L 2 µmol/L 1 µmol/L 2 µmol/L— —

Parental cells Resistant cells

Figure 2.A, SCH772984 potently inhibits in na€�ve HCT-116 cells but failed to inhibit phospho-ERK and phospho-RSK in resistant cells. B, SCH772984 effectively inhibits MAPK-dependent pathway gene expression in na€�ve but not in SCH772984-resistant HCT-116 cells. Na€�ve (HCT-116P) or HCT-116R cells were either untreated or treatedwith vehicle (1% DMSO), SCH772984 (1 or 2mmol/L) for 4, 16, or 24 hours. Expressionof theMAPK-regulated geneswas assessed by quantitative TaqManPCR. Compound-treated samples from parental and resistant cells were normalized to their respective DMSO controls. C, comparison of VTX-11e on HCT-116P and HCT-116R

cells (top, proliferation; bottom, target engagement). D, comparison of GSK1120212 on HCT-116P and HCT-116R cells (top, proliferation; bottom, target engagement).

ERK Inhibitor Resistance

www.aacrjournals.org Mol Cancer Ther; 15(4) April 2016 553

on January 18, 2021. © 2016 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst February 1, 2016; DOI: 10.1158/1535-7163.MCT-15-0172

Page 7: Dissecting Therapeutic Resistance to ERK InhibitionJaren Arbanas1, Daniel McMasters1, Clifford R. Restaino1, Priya Dayananth2, Stuart Black2, Nathaniel L. Elsen 2 , Anthony Mannarino

kinase assay. As both enzymes exhibited similarKM (andmaximalvelocity, Vmax) values for ATP (see Supplementary Fig. S2), theATP concentration for wild-type and mutant enzyme was kept atKM (45 mmol/L). Three independent experiments were conductedin parallel for wild-type ERK1 and ERK1G186D proteins. IC50

values of 0.04 and 37 nmol/L respectively, were observed(Fig. 4A; Table 2, shaded region) indicating a 1,000-fold reductionin the ability of SCH772984 to inhibit mutant ERK1.

To further investigate the difference in inhibition demonstratedformutant enzymeby SCH772984,we conductedbinding studiesusing surface plasmon resonance (Biacore T200) in high-perfor-mance mode. In all cases, where saturation was attained, themaximal binding signal observed was >90% of the theoreticalmaximum level within error (see Supplementary Text S1). Thisindicated that all of the protein was properly folded. Inhibitorbinding to the active forms of both wild-type ERK1 (Fig. 4B) andmutant ERK1G186D (Supplementary Fig. S3) was tested across 9-point inhibitor titration ranges of 25 to 0.2 nmol/L (2-folddilution) and 5,000 to 0.76 nmol/L (3-fold dilution), respective-ly. A single replicate was incorporated to demonstrate signalreproducibility. Thedata for activewild-typeERK1fit satisfactorily(c2 ¼ 0.03, U value ¼ 1) to a 1:1 binding model (Fig. 4B, blacklines). For active mutant ERK1G186D, however, inhibitor bindingand release were too rapid to resolve values for the individual onand off rate constants (koff > 2 s�1, based on the instrumentdetection limit). Therefore, these data were fitted (c2¼ 0.08) to asteady-state binding model (solid line, Supplementary Fig. S3),yielding a KD value. Resultant fitted values for all relevant bindingparameters for active wild-type ERK1 and mutant ERK1G186D areshown in Table 2 (left region). The equilibrium binding constants(KD) determined for both active proteins were approximately 10-foldweaker than the corresponding IC50 values determinedbyTR-FRET. However, the large reduction in binding affinity (�1,000fold) previously observed with themutant was similarly observedin this analysis, a shift primarily attributable to differences in theoff-rates.

Binding of SCH772984 to both inactive wild-type ERK1 andinactive mutant ERK1G186D was also analyzed using surfaceplasmon resonance (Biacore T200) in single cycle kinetics and

high-performance modes, respectively (Fig 4C and D). For inac-tive wild-type ERK1, results from the sequential injections at 5different inhibitor concentrations (50–0.62 nmol/L, 3-fold dilu-tion) were fitted satisfactorily (c2 ¼ 0.02, U value ¼ 1) to a 1:1binding model. For inactive mutant ERK1G186D, results from a 9-point inhibitor titration, incorporating a single replicate, from1,666 to 0.76 nmol/L (3-fold dilution) were also fit satisfactorily(c2 ¼ 0.02, U value ¼ 1) to a 1:1 binding model. Values for thefitted parameters are shown in Table 2 (right region). Significant-ly, the KD values for SCH772984 with both the inactive wild-typeERK1 and inactive mutant ERK1G186D differ by approximately200-fold, primarily expressed in a difference in the off-rates, withweaker binding to the mutant protein as observed for the activeprotein.

Given the significant sequence homology between ERK1 andERK2 (Fig. 3C), we sought to extend these observations to ERK2.IC50 of SCH772984 for the active forms of both wild-type ERK2and ERK2G169D proteins was determined using an in vitro TR-FRETkinase assay as described above for ERK1. The results are shown inSupplementary Fig. S4A and summarized in Supplementary TableS1 (shaded region). An identical set of surface plasmon resonanceexperiments was also repeated for both active and inactive wild-type ERK2 and mutant ERK2G169D (see Supplementary Informa-tion, Supplementary Table S1, and Supplementary Figs. S4 andS5). In general, the results were very similar to those observed forERK1 (Table 2), with some minor differences (2- to 4-fold).Somewhat tighter binding of SCH772984 was observed withactive mutant ERK2G169D (7.5-fold difference) than with activemutant ERK1G186D, but the overall trends remained predomi-nantly the same. It may be worth noting that for both ERK1 andERK2, KD values determined by Biacore analysis for the inactivewild-type andmutant proteins (Table 2 and Supplementary TableS1, right region)were similar to the IC50 values determined for theactivewild-type andmutant proteins by TR-FRET analysis (Table 2and Supplementary Table S1, shaded region). Taken together,these data also suggest that ERK2G169D would be expected tofunction similarly to ERK1G186D in the context of resistance andthat the significant increase in SCH772984 off-rates (upon muta-tion) is primarily responsible. Finally, recent studies onERK2with

BA

C

SCH772984 log (mol/L)

Via

bilit

y (%

con

trol)

–10 –8 –6 –40

50

100ERK1G186D

ERK1 WT

ERK2 WT

HC

T-11

6/E

RK

1G18

6D

HC

T-11

6/W

T ER

K1

HC

T-11

6/ W

T ER

K2

HC

T-11

6R

SCH772984 (2 µmol/L)

HC

T-11

6P

Hs MAPK1/ERK2 : 157 - LNTTCDLKICDFGLARVADPDHD - 179

Rn MAPK1/ERK2 : 155 - LNTTCDLKICDFGLARVADPDHD - 177

Hs MAPK3/ERK1 : 174 - INTTCDLKICDFGLARIADPEHD - 196

Rn MAPK3/ERK1 : 175 - INTTCDLKICDFGLARIADPEHD - 197

169

167

186

187

pRSK

pErk

Actin

tRSK

tErk

+ –– + – + – + – +

Figure 3.A,overexpressionofERK1G186DmutantmakesHCT-116 cells resistant toSCH772984.B, SCH772984treatmentofcells overexpressingERKG186Ddoesnot affect ERKorRSKphosphorylation levels. C, sequence alignment of human and rat ERK1 and 2 proteins showing conservation of the DFG motif in ERK1 and 2 and across species.

Jha et al.

Mol Cancer Ther; 15(4) April 2016 Molecular Cancer Therapeutics554

on January 18, 2021. © 2016 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst February 1, 2016; DOI: 10.1158/1535-7163.MCT-15-0172

Page 8: Dissecting Therapeutic Resistance to ERK InhibitionJaren Arbanas1, Daniel McMasters1, Clifford R. Restaino1, Priya Dayananth2, Stuart Black2, Nathaniel L. Elsen 2 , Anthony Mannarino

SCH772984, using a binding kinetics approach, also exhibitedslow off-kinetics for this compound (27).

To visualize the disruption of ERK binding to SCH772984, weattempted to generate X-ray crystal structures of SCH772984 (Fig.5A) bound to either wild-type or mutant rat ERK1 or ERK2proteins. ERK2 bound to SCH772984 yielded crystals of goodquality and high resolution. The structures of rat wild-type andERK2G167D-mutant proteins (corresponding to humanERK1G186D and ERK2G169D; see Fig. 3C) bound to SCH772984were determined at a resolution of 1.45 Å and 1.70 Å, respectively.Given the high degree of homology between ERK1 and ERK2 (andthe observation that the binding mode of SCH772984 is con-served in human ERK1 and ERK2; ref. 28), as well as the resultsobtained from the surface plasmon resonance studies, we believethe observations from these studies are applicable to ERK1 aswell.Rat ERK2 residue numbering is used in the description of the data.

Figure 5B is a surface representation overlay of the structures ofapo ERK2 and ERK2 bound to SCH772984. The apo ERK2structure shows the aromatic rings of Tyr34 and Tyr62 (both inmagenta) stacked against each other to form a p–p interaction.Binding of SCH772984 induces a profound shift in the confor-mation of the glycine-rich loop particularly at Tyr34. The piper-azine phenyl pyrimidine portion of SCH772984 induces a move-ment of Tyr34 (peach) moving it into the ATP site to occupy a

pocket created by an approximately 5 Åmovement of the glycine-rich loop. This causes the aromatic ring of Tyr34 to pack against thepyrrolidine ring of SCH772984. A small shift in position is alsoseen with Tyr62 (green ring). These changes would likely beincompatible with ATP binding.

The structures of apo ERK2 andERK2 in complexwith ATPhavepreviously been published (29, 30). The ATP-competitive natureof SCH772984 is evident by its complete occupancy of the ATP-binding site in ERK2 (Fig. 5C). The indazole moiety ofSCH772984 hydrogen bonds to the backbone carbonyl andamide nitrogen of hinge residues Asp104 and Met106, mimickingthe adenine ring of ATP. The secondary amide linker and pyrro-lidine form an extensive hydrogen bond network including thecatalytic Lys52 and occupy the area where the triphosphatemoietyof ATP binds. The ribose-binding region appears unoccupied bySCH772984.We had previouslymade the interesting observationthat binding of SCH772984 to ERK prevents its phosphorylationby MEK, perhaps due to changes in conformation induced by thecompound (23). The structure of ERK2 bound to SCH772984now reveals the nature of this mechanism of action. Binding ofSCH772984 produces a >1-Å movement at the beginning of theC-helix, which is transmitted to the C-terminal tail (Fig. 5D).These movements in the N-terminal domain likely prevent MEKfrom recognizing ERK2 as a substrate for phosphorylation. This is

C

D

WT t1/2 = 100 min

t1/2 = 0.6 min

A Inactive

InactiveG186D

WT

WT

G186D

Time (sec)

Time (sec)

10

6

2

8

4

0

0

2

4

6

8

0 1,000 2,000 3,000 4,000

200150100500

RU

RU

RU

5,000Log [SCH772984] (mol/L)

% In

hibi

tion

–12 –10 –8 –6 –40

25

50

75

100

Time (sec)

0

2

4

6

8

200 400 600 800 1,0000

B Active

Active

Figure 4.ERK1 TR-FRET assay (IC50) and Biacore data (black line is a fit to a 1:1 binding model, c2 < 0.03, U value ¼ 1) with SCH772984. A, TR-FRET assay data foractive wild-type (filled circles, orange) and active G186D (open circles, blue) ERK1 with SCH772984 and ATP (at KM, 45 mmol/L). Data were fit to 4-parameterdose–response curve (solid line) by nonlinear regression (R2 ¼ 0.99). Experiment was repeated three times on three separate occasions, in singlet.Representative singlet result is shown. B, Biacore datawith activewild-type ERK1 and SCH772984 [25 (light green)�0.2 (red) nmol/L, 2-fold dilution]. As expected,observed binding level increased with increasing drug concentration, with duplicate shown in same color (6.25 nmol/L, yellow). C, single-cycle kineticsBiacore data with inactive wild-type ERK1 and SCH772984 (50–0.62 nmol/L, 3-fold dilution). D, Biacore data with inactive G186D ERK1 and SCH772984 [1,666 (lightgreen) � 0.76 (red) nmol/L, 3-fold dilution]. As expected, observed binding level increased with increasing drug concentration, with duplicate shown insame color (20 nmol/L, purple).

ERK Inhibitor Resistance

www.aacrjournals.org Mol Cancer Ther; 15(4) April 2016 555

on January 18, 2021. © 2016 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst February 1, 2016; DOI: 10.1158/1535-7163.MCT-15-0172

Page 9: Dissecting Therapeutic Resistance to ERK InhibitionJaren Arbanas1, Daniel McMasters1, Clifford R. Restaino1, Priya Dayananth2, Stuart Black2, Nathaniel L. Elsen 2 , Anthony Mannarino

consistent with multiple interaction surfaces on ERK required foreffective recognition and phosphorylation by MEK (29). In addi-tion, thepresence of SCH772984 in the newly formedbinding sitealong the C-helix could lock ERK2 in a conformation that wouldnot allow the N-terminal domain to generate motions requiredfor phosphorylation upon binding of MEK.

When the bindingmode of SCH772984was compared with anATP-competitive ERK inhibitor from a different structural class,VTX-11e (Fig. 5F), we observed that two thirds of VTX-11e overlapclosely with the pyridine, indazole, and pyrrolidine portions ofSCH772984, whereas the chlorophenyl group of VTX-11e tucksunderneath the glycine-rich loop and does not overlaySCH772984. Importantly, Tyr34 of ERK2 occupies completelydifferent sites when bound to each of the two compounds. Theprofound distortion of the glycine-rich loop and flipping of Tyr34

into the ATP site that are induced by SCH772984, do not occurwith VTX-11e, which does not extend as far toward theaChelix ofERK. Thus, thep–p stacking interactionof Tyr34 andTyr62 of ERK isnot significantly disrupted upon VTX-11e binding. Another con-sequence of the different binding modes is that the distancebetween the Ca of G167 and the nearest nonhydrogen atom ofVTX-11e is nearly twice the distance between the Ca and thepiperazine ring of SCH772984 (7.1 Å vs. 3.2 Å).

The structure of rat ERK2G167D-mutant complex withSCH772984 shows a relatively conserved binding mode com-pared with wild-type (Fig. 5E). The pyridine, indazole, andpyrrolidine portions of SCH772984 overlap almost perfectly withwhat is observed with wild-type ERK2. Significant changes start atthe piperazine ring, whichmoves about 1 Å to avoid a steric clashwith the additional atoms from Asp167 of mutant ERK2. Thisproduces an upward shift of the entire right side of SCH772984 aswell as a 4.5-degree counterclockwise rotation about the pyrro-lidine. As a consequence, there is a 1.5 Å movement of theterminal pyrimidine of SCH772984 altering the p–p stackingwith the Tyr62 aromatic ring. Interestingly, Asp167 ofmutant ERK2forms anew salt bridgewithArg65. It isworth noting that althoughthe shifts in the interaction of SCH772984 with mutant ERK2appear minor, these changes are accompanied by a significantdecrease in binding affinity (attributable to a rapid off-rate)observed in the surface plasmon resonance studies and by reac-tivation of the pathway.

DiscussionAcquisition of resistance upon prolonged drug treatment has

been acknowledged as a major limitation of therapy with kinaseinhibitors (26). Development of therapeutic options that over-come resistance to significantly impact the outcome of thesepatients depends on understanding the mechanistic basis for theresistance. We have previously shown that, SCH772984, a noveland selective ERK inhibitor, effectively inhibited cell proliferationand ERK signaling in BRAF or MEK inhibitor-resistant models

(23) and ERK inhibition could therefore be an effective therapyoption for relapsed patients who had reactivated the ERK path-way. In the current study, we consider the possibility that patientstreated with an ERK inhibitor would eventually develop resis-tance. Given that ERK is the final signaling node in the MAPKpathway, understanding the biologic mechanism of ERK resis-tance should enable the design of the next generation of moreeffective therapies for patients who become refractory to an ERKinhibitor.

Our data suggest that a mutation in the highly conserved DFG(Asp-Phe-Gly) motif, a central structural unit in kinases, may bean effective way for cells to acquire resistance to ERK inhibition.The acquired ERK1G186D mutation was accompanied by pathwayreactivation and proliferation in the presence of SCH772984.pERK levels that are abolished by compound treatment in paren-tal cells are restored in resistant cells to levels observed in untreat-ed parental cells. Overexpression of ERK1G186D in the parentalcells is sufficient to mediate the same effect, indicating thatmutation of a single amino acid residue is sufficient to conferresistance. These data support the hypothesis that ERKG186D is abinding-deficient mutant and not an "activated"mutant. It is alsoimportant to note that this is a heterozygous mutation and thuscomplete resistance is not expected with one wild-type allele stillpresent in the cells. Therefore, it is likely that there is some degreeof pathway restoration with the ERK mutant, though not com-plete. In addition, it is likely that the acquisition of resistancecomes at a cost to the ERK enzyme (resulting in a "reduced fitness"phenotype), which is consistent with our observations that basallevel of cellular ERK catalytic activity (as measured by pRSK andother downstream genes) is reduced in the resistant cells com-pared with parental.

How does amutation of the terminal glycine to aspartate in theDFG motif confer insensitivity to SCH772984? For the sake ofconsistency with previously published work (29, 30), rat ERK2amino acid numbering is used in the discussion below. Theevolutionarily conserved DFG motif is found at the base of theactivation loop and its orientation in kinases ("DFG in" or "DFGout") is associated with an "active" or "inactive" ATP-binding siteconformation. In the "DFG in" conformation of ERK, the aspar-tate side chain of the DFG motif faces into the active site tofacilitate catalysis. The neighboring phenylalanine occupies ahydrophobic pocket adjacent to the ATP-binding site. Finally,the glycine holds the active site open by interaction with the C-helix. In particular, interaction ofG167with specific residues in theC helix such as Arg65 (which points toward the active site) andGln64, further promotes the open conformation (31). Thus, theDFG motif plays a crucial role in maintenance of the conforma-tional accessibility of the active site of ERK. SCH772984 binds theactive formof theATP-binding site of ERK,where all the conservedresidues are in the proper orientation for catalysis. WhenGly167 ismutated, the aspartate side chains face the active site creating asteric clash that might contribute to instability of SCH773984

Table 2. Summary of TR-FRET assay (IC50, shaded region) and Biacore data (unshaded region) for SCH772984 binding to both active/inactive wild-type ERK1 andactive/inactive ERK1G186D mutant, showing weaker binding for ERK1G186D

Active InactiveIC50 (nmol/L) KD (nmol/L) kon (M

�1s�1) 106 koff (s�1) 10�4 t1/2 (min) KD (nM) kon (M

�1s�1) 106 koff (s�1) 10�4 t1/2 (min)

ERK1 WT 0.04 � 0.02 0.49 1.24 � 0.01 6.04 � 0.01 19 0.075 1.55 � 0.01 1.16 � 0.001 100ERK1 G186D 37 � 14 590a Rapid Rapid <0.01b 14 1.36 � 0.01 187 � 1 0.6aSee Supplementary Fig. S3.bUpper limit estimate based on detection limit of Biacore instrument.

Jha et al.

Mol Cancer Ther; 15(4) April 2016 Molecular Cancer Therapeutics556

on January 18, 2021. © 2016 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst February 1, 2016; DOI: 10.1158/1535-7163.MCT-15-0172

Page 10: Dissecting Therapeutic Resistance to ERK InhibitionJaren Arbanas1, Daniel McMasters1, Clifford R. Restaino1, Priya Dayananth2, Stuart Black2, Nathaniel L. Elsen 2 , Anthony Mannarino

Figure 5.Structure determination of wild-type and ERK2G167D bound to SCH772984. A, chemical structure of SCH772984. B, comparison of the structures of rat ERK2 apo andERK2 bound to SCH772984. In the apo structure Tyr34 of the glycine-rich loop (shown in magenta color) is stacked against Tyr62 (shown as a green aromatic ring).Compound binding causes Tyr34 of the glycine-rich loop (shown in peach) to flip away from Try62 and instead pack against the pyrrolidine moiety of SCH772984.The N- and C-terminal lobes are depicted in green and purple surfaces, respectively. MEK phosphorylation sites are highlighted in black, the DFG motif in red,and the hinge including thegatekeeper residue in yellow. SCH772984 is shownasorange sticks. C, overlay of SCH772984 andATPbound to rat ERK2protein. Structure ofSCH772984 bound to ERK2 confirms the ATP-competitive nature of the compound. The indazole of SCH772984 overlaps the adenine ring of ATP and the amidelinker and pyrrolidone occupy the area bound by the triphosphate moiety of ATP. SCH772984 and ATP are shown as white and yellow sticks, respectively.ERK2 protein is depicted as a yellow ribbon. D, ribbon diagram of wild-type ERK2 bound to SCH772984 (peach color) overlaid with apo ERK2 (magenta) showingmovement produced in the N-terminal domain could prevent MEK recognition/binding of ERK2 leading to loss of phosphorylation. ERK2 bound to SCH772984is shownaspeach ribbons andApoERK2 is shownasmagenta ribbons. E, overlay of the structuresofwild-type rat ERK2 (asgreen andpeach colored ribbon) andmutantrat ERK2G167D (shown as yellow ribbon) proteins bound to SCH772984 showing a steric clash of the piperazine ring of SCH772984 with the additional atomsof Asp167 of mutant ERK2. SCH772984 bound to wild-type ERK2 and ERK2G167D are shown as orange and yellow sticks respectively. F, overlay of the structures ofSCH772984 (orange) and VTX-11e (green) bound to ERK2 showing that the profound distortion of the glycine-rich loop and flipping of Tyr34 into the ATP siteare uniquely induced by SCH772984.

ERK Inhibitor Resistance

www.aacrjournals.org Mol Cancer Ther; 15(4) April 2016 557

on January 18, 2021. © 2016 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst February 1, 2016; DOI: 10.1158/1535-7163.MCT-15-0172

Page 11: Dissecting Therapeutic Resistance to ERK InhibitionJaren Arbanas1, Daniel McMasters1, Clifford R. Restaino1, Priya Dayananth2, Stuart Black2, Nathaniel L. Elsen 2 , Anthony Mannarino

binding to the active site. In contrast, in the case of VTX-11e, wespeculate that although the aspartate side chain of mutant ERKforces a shift of SCH772984 leading to destabilization of itsbinding, the greater distance observed with VTX-11e (Fig. 5F) maypreclude any impact of the aspartic acid substitution on its bindingto ERK. Thismay be the basis for our experimental observation (Fig2C) where proliferation of both parental and resistant cells wassimilarly sensitive to VTX-11e. We therefore speculate that ERKinhibitors structurally and mechanistically distinct fromSCH772984 may be efficacious in cells resistance to it.

Our previous results (23) implied that SCH772984 bindsand inhibits ERK by a unique mechanism. Not only doesSCH772984 inhibit ERK catalytic activity, but also preventsMEK phosphorylation of ERK. Our structural and biophysicaldata now demonstrate that the binding site of SCH772984completely overlaps with the binding site for ATP (Fig. 5C);moreover, SCH772984 binds ERK2 with higher affinity thanATP (0.5 nmol/L and 45 mmol/L, respectively). Thus, oncebound, SCH772984 can effectively prevent ATP access to thecatalytic site, and by extension, substrate phosphorylation.With regards to preventing phosphorylation by MEK, we hadpreviously hypothesized that binding by SCH772984 inducedor stabilized a conformational state in ERK that was notcompatible with MEK phosphorylation. The structure of ratERK2 bound to SCH772984 shows that although the com-pound shifts the p–p stacking interaction of Tyr34 and Tyr62 andforces a shift in the position of Tyr34, these changes do notappear to substantially shift the position of the residues Tyr183

and Tyr185 which are phosphorylated by MEK. However, com-pound binding does produce significant movements in the N-terminal domain via shifting of the positions of the C-helix andC-terminal tail (Fig. 5D). The C-helix is primarily involved inbinding ATP and as SCH772984 predominantly occupies thesame space as ATP, it pushes against the C-helix. This shift is inturn transmitted to the C-terminal tail. Wilsbacher and collea-gues have previously demonstrated that the integrity of a fairlylarge surface area of the N-terminal structural domain needs tobe maintained for a productive interaction between MEK andERK (31). In particular, the sites of contact between MEK andERK include residues from the C-helix, C-terminal tail, and thephosphorylation lip. Loss of conformational integrity of the N-terminal domain could destabilize MEK binding and/or itsrecognition of ERK.

Mutation of the glycine residue in the DFG loop in the resis-tance setting ensures that SCH772984 does not occupy the ATPsite for any significant length of time. The KD values forSCH772984 with active human wild-type ERK1 and mutantERK1G186D differ by approximately 1,000-fold. In addition, asignificant difference in the off-rates was measured with weakerbinding (rapid release) to themutant protein (Fig. 4; Table 2). Theacquired mutation thereby alleviates the perturbations to the N-terminal domain induced by SCH772984 binding. Therefore,MEK in SCH772984-resistant cells is able to bind and effectivelyphosphorylate ERK,which thenphosphorylates RSK (Fig. 2A, 3A),resulting in the restoration of downstream transcriptional output(Fig. 2B). As a consequence, SCH772984-resistant cells retainsensitivity to a MEK inhibitor (Fig. 2D) raising the interestingpossibility that treatment with a MEK inhibitor could be atherapeutic option for patients who develop resistance to thisclass of ERK inhibitors.

Collectively, these data support the hypothesis thatERKG186D is a binding-deficient mutant and not an "activated"mutant.

Interestingly, we did not identify any mutations in the "gate-keeper" residue in our SCH772984-resistant lines. A number ofdrug-resistant kinase mutants with altered gatekeeper residueshave been identified inmultiple kinases that have been effectivelytargeted clinically:BCR-Abl (T315I), EGFR (T790M),KIT (T670I), andALK (L1196M). Mutation of the gatekeeper residue in rat ERK2(ERK2Q103) to glycine or alanine has been shown to increaseautoactivation by enhancing autophosphorylation of the Thr183

and Tyr185 in the activation loop (32). Exploring this further, wefound that mutation of human ERK2 glutamine105 to glutamicacid (Q105E) was sufficient tomediate resistance in the presence ofan ERK inhibitor through mechanisms that are not fully defined(data not shown). Thus, although no gatekeeper mutations werefound in the current study, it is possible that such mutations mayarise in patients upon sustained ERK inhibition. It is also worthnoting that residue Tyr34 plays a key role in accommodatingSCH772984 in the active site of ERK. In order for the longpiperazine–phenyl–pyrimidine tricyclic ring of SCH772984 tobind ERK, the side-chain of Tyr34 flips under the glycine-rich looptoward the adenine-binding site, thus opening up a new sidepocket. In this new conformation, Tyr34 is positioned directly overthe pyrrolidine ring of SCH772984, making a favorable hydro-phobic interaction. One could speculate that a mutation of Tyr34

to a residue less amenable to this drastic movement, may con-tribute to SCH772984 resistance in patients. Indeed, as Chaikuadand colleagues (28) recently showed, mutation of Tyr34 to glu-tamine or alanine reduced the binding affinity of ERK toSCH772984.

ERK inhibitors are currently undergoing clinical trials inpatients with activated or reactivated MAPK signaling. Acquiredresistance to these agents may emerge in the clinic. This study isthe first to provide preclinical evidence that a single amino acidmutation in the ERK DFG motif can lead to ERK inhibitorresistance. If this mechanism of resistance proves to be clinicallyrelevant, our data suggest that a second-generation ERK inhibitorwith a structurally distinct scaffold may be efficacious in thiscontext (Fig. 2C). In addition, given the compelling clinical proof-of-concept of BRAF/MEK inhibitor combination efficacy, combi-nation studies with MEK/ERK or BRAF/ERK inhibitors may pro-vide an alternative paradigm of treatment. Indeed, Wong andcolleagues (24) recently provided preliminary evidence thatchronic exposure of tumor cells in culture to a combination ofSCH772984 and the BRAF inhibitor vemurafenib increased thetime required to develop acquire resistance. It will therefore beimportant to further understand how resistance to targeting singlenodes in the MAPK pathway compares with these combinations.In particular, selection pressures for mutations inMEK or ERK inthe context of concurrent BRAF inhibition will be important tomodel, to enable administration of therapies most likely toprovide maximal patient benefit.

Disclosure of Potential Conflicts of InterestD. McMasters has ownership interest in Merck & Co., Inc. No potential

conflicts of interest were disclosed by the other authors.

Authors' ContributionsConception and design: S. Jha, E.J. Morris, C.R. Restaino, A. Cooper, S. Fawell,L. Jayaraman, A.A. Samatar

Mol Cancer Ther; 15(4) April 2016 Molecular Cancer Therapeutics558

Jha et al.

on January 18, 2021. © 2016 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst February 1, 2016; DOI: 10.1158/1535-7163.MCT-15-0172

Page 12: Dissecting Therapeutic Resistance to ERK InhibitionJaren Arbanas1, Daniel McMasters1, Clifford R. Restaino1, Priya Dayananth2, Stuart Black2, Nathaniel L. Elsen 2 , Anthony Mannarino

Developmentofmethodology: S. Jha, A.Hruza,M.S.Mansueto, J. Arbanas, C.R.Restaino, P. Dayananth, A. Mannarino, S. Fawell, L. Jayaraman, A.A. SamatarAcquisition of data (provided animals, acquired and managed patients,provided facilities, etc.): S. Jha, A. Hruza, M.S. Mansueto, G.K. Schroeder,J. Arbanas, C.R. Restaino, S. Black, N.L. Elsen, A.A. SamatarAnalysis and interpretationofdata (e.g., statistical analysis,biostatistics, compu-tational analysis): S. Jha, E.J. Morris, A. Hruza, M.S. Mansueto, G.K. Schroeder,J. Arbanas, D. McMasters, C.R. Restaino, A. Cooper, L. Jayaraman, A.A. SamatarWriting, review, and/or revisionof themanuscript: S. Jha, E.J.Morris, A.Hruza,M.S. Mansueto, G.K. Schroeder, J. Arbanas, D. McMasters, N.L. Elsen, S. Fawell,L. Zawel, L. Jayaraman, A.A. SamatarAdministrative, technical, or material support (i.e., reporting or organizingdata, constructing databases): S. Jha, E.J. Morris, M.S. Mansueto, S. Black,A. Mannarino, L. Jayaraman, A.A. SamatarStudy supervision: S. Jha, E.J. Morris, L. Jayaraman, A.A. Samatar

AcknowledgmentsThe authors thank Bo-Sheng Pan, Corey Strickland, Ulrike Philippar, David

Witter, and Bart Lutterbach for discussions.

Grant SupportThis work was financially supported by Merck Research Laboratories.The costs of publication of this article were defrayed in part by the

payment of page charges. This article must therefore be hereby markedadvertisement in accordance with 18 U.S.C. Section 1734 solely to indicatethis fact.

Received March 27, 2015; revised December 7, 2015; accepted December 23,2015; published OnlineFirst February 1, 2016.

References1. Adjei AA. Blockingoncogenic Ras signaling for cancer therapy. JNatl Cancer

Inst 2001;93:1062–74.2. Fernandez-Medarde A, Santos E. Ras in cancer and developmental diseases.

Genes Cancer 2011;2:344–58.3. Davies H, Bignell GR, Cox C, Stephens P, Edkins S, Clegg S, et al. Mutations

of the BRAF gene in human cancer. Nature 2002;417:949–54.4. Krauthammer M, Kong Y, Ha BH, Evans P, Bacchiocchi A, McCusker JP,

et al. Exome sequencing identifies recurrent somatic RAC1 mutations inmelanoma. Nat Genet 2012;44:1006–14.

5. Chapman PB, Hauschild A, Robert C, Haanen JB, Ascierto P, Larkin J, et al.Improved survival with vemurafenib in melanoma with BRAF V600Emutation. N Engl J Med 2011;364:2507–16.

6. Flaherty KT, Infante JR, Daud A, Gonzalez R, Kefford RF, Sosman J, et al.Combined BRAF and MEK inhibition in melanoma with BRAF V600mutations. N Engl J Med 2012;367:1694–703.

7. Flaherty KT, Robert C, Hersey P, Nathan P, Garbe C, Milhem M, et al.Improved survival with MEK inhibition in BRAF-mutated melanoma. NEngl J Med 2012;367:107–14.

8. Flaherty KT, Yasothan U, Kirkpatrick P. Vemurafenib. Nat Rev Drug Discov2011;10:811–2.

9. Hauschild A, Grob JJ, Demidov LV, Jouary T, Gutzmer R, Millward M,et al. Dabrafenib in BRAF-mutated metastatic melanoma: a multicentre,open-label, phase 3 randomised controlled trial. Lancet 2012;380:358–65.

10. Corcoran RB, Dias-Santagata D, Bergethon K, Iafrate AJ, Settleman J,Engelman JA. BRAF gene amplification can promote acquired resistanceto MEK inhibitors in cancer cells harboring the BRAF V600E mutation. SciSignal 2010;3:ra84.

11. Corcoran RB, Ebi H, Turke AB, Coffee EM, Nishino M, Cogdill AP, et al.EGFR-mediated re-activation of MAPK signaling contributes to insensitiv-ity of BRAFmutant colorectal cancers to RAF inhibition with vemurafenib.Cancer Discov 2012;2:227–35.

12. Girotti MR, Pedersen M, Sanchez-Laorden B, Viros A, Turajlic S, Niculescu-Duvaz D, et al. Inhibiting EGF receptor or SRC family kinase signalingovercomes BRAF inhibitor resistance inmelanoma. Cancer Discov 2013;3:158–67.

13. Poulikakos PI, Persaud Y, JanakiramanM, Kong X, Ng C,Moriceau G, et al.RAF inhibitor resistance is mediated by dimerization of aberrantly splicedBRAF(V600E). Nature 2011;480:387–90.

14. Shi H, Hugo W, Kong X, Hong A, Koya RC, Moriceau G, et al. Acquiredresistance and clonal evolution in melanoma during BRAF inhibitortherapy. Cancer Discov 2014;4:80–93.

15. Shi H, Kong X, Ribas A, Lo RS. Combinatorial treatments that overcomePDGFRbeta-driven resistance of melanoma cells to V600EB-RAF inhibi-tion. Cancer Res 2011;71:5067–74.

16. Shi H, Moriceau G, Kong X, Lee MK, Lee H, Koya RC, et al. Melanomawhole-exome sequencing identifies (V600E)B-RAF amplification-mediat-ed acquired B-RAF inhibitor resistance. Nat Commun 2012;3:724.

17. Van Allen EM, Wagle N, Sucker A, Treacy DJ, Johannessen CM, Goetz EM,et al. The genetic landscape of clinical resistance to RAF inhibition inmetastatic melanoma. Cancer Discov 2013;4:94–109.

18. Wagle N, Emery C, Berger MF, Davis MJ, Sawyer A, Pochanard P, et al.Dissecting therapeutic resistance to RAF inhibition inmelanoma by tumorgenomic profiling. J Clin Oncol 2011;29:3085–96.

19. Emery CM, Vijayendran KG, Zipser MC, Sawyer AM, Niu L, Kim JJ, et al.MEK1mutations confer resistance toMEK and B-RAF inhibition. Proc NatlAcad Sci U S A 2009;106:20411–6.

20. WangH,Daouti S, LiWH,Wen Y, RizzoC,Higgins B, et al. Identification ofthe MEK1(F129L) activating mutation as a potential mechanism ofacquired resistance to MEK inhibition in human cancers carrying the B-RafV600E mutation. Cancer Res 2011;71:5535–45.

21. Greger JG, Eastman SD, Zhang V, BleamMR, Hughes AM, Smitheman KN,et al. Combinations of BRAF, MEK, and PI3K/mTOR inhibitors overcomeacquired resistance to the BRAF inhibitor GSK2118436 dabrafenib, medi-ated by NRAS or MEK mutations. Mol Cancer Ther 2012;11:909–20.

22. Rizos H,Menzies AM, Pupo GM, CarlinoMS, Fung C, Hyman J, et al. BRAFinhibitor resistance mechanisms in metastatic melanoma: spectrum andclinical impact. Clin Cancer Res 2014;20:1965–77.

23. Morris EJ, Jha S, Restaino CR, Dayananth P, Zhu H, Cooper A, et al.Discovery of a novel ERK inhibitor with activity in models of acquiredresistance to BRAF and MEK inhibitors. Cancer Discov 2013;3:742–50.

24. Wong DJ, Robert L, Atefi MS, Lassen A, Avarappatt G, Cerniglia M, et al.Antitumor activity of the ERK inhibitor SCH722984 against BRAF mutant,NRAS mutant and wild-type melanoma. Mol Cancer 2014;13:194.

25. Emsley P, Lohkamp B, Scott WG, Cowtan K. Features and development ofCoot. Acta Crystallogr D Biol Crystallogr 2010;66:486–501.

26. Lovly CM, Shaw AT. Molecular pathways: resistance to kinase inhibitorsand implications for therapeutic strategies. Clin Cancer Res 2014;20:2249–56.

27. Rudolph J, Xiao Y, Pardi A, Ahn NG. Slow inhibition and conformationselective properties of extracellular signal-regulated kinase 1 and 2 inhi-bitors. Biochemistry 2015;54:22–31.

28. Chaikuad A, Tacconi EM, Zimmer J, Liang Y, Gray NS, TarsounasM, et al. Aunique inhibitor binding site in ERK1/2 is associated with slow bindingkinetics. Nat Chem Biol 2014;10:853–60.

29. Zhang F, Strand A, Robbins D, Cobb MH, Goldsmith EJ. Atomic structureof the MAP kinase ERK2 at 2.3 A resolution. Nature 1994;367:704–11.

30. Zhang J, Shapiro P, Pozharski E. Structure of extracellular signal-regulatedkinase 2 in complex with ATP and ADP. Acta Crystallogr Sect F Struct BiolCryst Commun 2012;68:1434–9.

31. Wilsbacher JL, Goldsmith EJ, Cobb MH. Phosphorylation of MAP kinasesby MAP/ERK involves multiple regions of MAP kinases. J Biol Chem 1999;274:16988–94.

32. Emrick MA, Lee T, Starkey PJ, Mumby MC, Resing KA, Ahn NG. Thegatekeeper residue controls autoactivation of ERK2 via a pathway ofintramolecular connectivity. Proc Natl Acad Sci U S A 2006;103:18101–6.

www.aacrjournals.org Mol Cancer Ther; 15(4) April 2016 559

ERK Inhibitor Resistance

on January 18, 2021. © 2016 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst February 1, 2016; DOI: 10.1158/1535-7163.MCT-15-0172

Page 13: Dissecting Therapeutic Resistance to ERK InhibitionJaren Arbanas1, Daniel McMasters1, Clifford R. Restaino1, Priya Dayananth2, Stuart Black2, Nathaniel L. Elsen 2 , Anthony Mannarino

2016;15:548-559. Published OnlineFirst February 1, 2016.Mol Cancer Ther   Sharda Jha, Erick J. Morris, Alan Hruza, et al.   Dissecting Therapeutic Resistance to ERK Inhibition

  Updated version

  10.1158/1535-7163.MCT-15-0172doi:

Access the most recent version of this article at:

  Material

Supplementary

  http://mct.aacrjournals.org/content/suppl/2016/01/30/1535-7163.MCT-15-0172.DC1

Access the most recent supplemental material at:

   

   

  Cited articles

  http://mct.aacrjournals.org/content/15/4/548.full#ref-list-1

This article cites 32 articles, 14 of which you can access for free at:

  Citing articles

  http://mct.aacrjournals.org/content/15/4/548.full#related-urls

This article has been cited by 8 HighWire-hosted articles. Access the articles at:

   

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

  Subscriptions

Reprints and

  [email protected]

To order reprints of this article or to subscribe to the journal, contact the AACR Publications Department at

  Permissions

  Rightslink site. Click on "Request Permissions" which will take you to the Copyright Clearance Center's (CCC)

.http://mct.aacrjournals.org/content/15/4/548To request permission to re-use all or part of this article, use this link

on January 18, 2021. © 2016 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst February 1, 2016; DOI: 10.1158/1535-7163.MCT-15-0172