tmem16a induces mapk and contributes directly …...whitney rank-sum test was used to test the...

13
Molecular and Cellular Pathobiology TMEM16A Induces MAPK and Contributes Directly to Tumorigenesis and Cancer Progression Umamaheswar Duvvuri 1,7 , Daniel J. Shiwarski 1 , Dong Xiao 1 , Carol Bertrand 2 , Xin Huang 6 , Robert S. Edinger 5 , Jason R. Rock 10 , Brian D. Harfe 11 , Brian J. Henson 8,9 , Karl Kunzelmann 12 , Rainer Schreiber 12 , Raja S. Seethala 3 , Ann Marie Egloff 1 , Xing Chen 1 , Vivian W. Lui 1 , Jennifer R. Grandis 1,4 , and Susanne M. Gollin 1,3,8,9 Abstract Frequent gene amplication of the receptor-activated calcium-dependent chloride channel TMEM16A (TAOS2 or ANO1) has been reported in several malignancies. However, its involvement in human tumorigenesis has not been previously studied. Here, we show a functional role for TMEM16A in tumor growth. We found TMEM16A overexpression in 80% of head and neck squamous cell carcinoma (SCCHN), which correlated with decreased overall survival in patients with SCCHN. TMEM16A overexpression signicantly promoted anchorage- independent growth in vitro, and loss of TMEM16A resulted in inhibition of tumor growth both in vitro and in vivo. Mechanistically, TMEM16A-induced cancer cell proliferation and tumor growth were accompanied by an increase in extracellular signalregulated kinase (ERK)1/2 activation and cyclin D1 induction. Pharmacologic inhibition of MEK/ERK and genetic inactivation of ERK1/2 (using siRNA and dominant-negative constructs) abrogated the growth effect of TMEM16A, indicating a role for mitogen-activated protein kinase (MAPK) activation in TMEM16A-mediated proliferation. In addition, a developmental small-molecule inhibitor of TMEM16A, T16A-inh01 (A01), abrogated tumor cell proliferation in vitro. Together, our ndings provide a mechanistic analysis of the tumorigenic properties of TMEM16A, which represents a potentially novel therapeutic target. The development of small-molecule inhibitors against TMEM16A may be clinically relevant for treatment of human cancers, including SCCHN. Cancer Res; 72(13); 327081. Ó2012 AACR. Introduction Amplication of the chromosomal band 11q13 is frequently seen in malignancies arising from breast, bladder, head and neck (SCCHN), and esophagus (1). Detailed molecular analysis of the 11q13 region led to identication of TMEM16A (also known as TAOS2, DOG1, and ANO1; refs. 24). This gene belongs to the TMEM16 family, characterized by the presence of 8 transmembrane domains and a highly conserved domain of unknown function (DUF590). Recently, TMEM16A was shown to be a calcium-activated chloride channel (CaCC; refs. 57). The biologic importance of Tmem16a is underscored by the lethal phenotype in the knock- out mouse because of abnormal tracheal development (810). TMEM16A contributes to many important physiologic functions, including the control of epithelial uid transport, saliva produc- tion, and gastrointestinal tract motility (1014). The third extra- cellular loop of TMEM16A (between transmembrane regions 5 and 6) has been suggested to be the putative pore-forming domain (6). Using mutagenesis approaches, the lysine residue at amino acid 610 (in the putative pore region) was found to be critical for its channel function (15). Specically, mutation of this residue from lysine to alanine (TMEM16A-K610A) was shown to be hypomorphic, with greatly reduced chloride conductance. CaCCs can be proapoptotic and suppress tumor formation in breast epithelia (16). However, there are exceptions, the putative CaCC, bestrophin (BEST1) is upregulated in colon cancer cells, and can promote proliferation (17). In addition, several reports have shown that TMEM16A is overexpressed in many tumor types including esophageal cancers, gastrointes- tinal stromal tumors, and SCCHN (1820). These seemingly discordant ndings prompted us to investigate the role of TMEM16A in epithelial malignancies. We initially determined the effects of TMEM16A on tumor proliferation (in vitro and in vivo) through gain- and loss-of- function experiments. Our results show that TMEM16A Authors' Afliations: Departments of 1 Otolaryngology, 2 Cell Biology and Physiology, 3 Pathology, 4 Pharmacology and Chemical Biology, and 5 Renal and Electrolyte Division, University of Pittsburgh Medical Center; 6 Depart- ment of Obstetrics, Gynecology, and Reproductive Sciences, University of Pittsburgh School of Medicine and Magee-Women's Research Institute, University of Pittsburgh Medical Center; 7 Veterans Affairs Pittsburgh Health System; 8 University of Pittsburgh Cancer Institute; 9 Human Genet- ics, University of Pittsburgh Graduate School of Public Health, Pittsburgh, Pennsylvania; 10 Department of Cell Biology, Duke University Medical Center, Durham, North Carolina; 11 Department of Molecular Genetics and Microbiology and the Genetics Institute, University of Florida, Gainesville, Florida; and 12 Department of Physiology, University of Regensburg, Regensburg, Germany Note: Supplementary data for this article are available at Cancer Research Online (http://cancerres.aacrjournals.org/). Current address for J.R. Rock: Department of Anatomy, University of California San Francisco. Corresponding Author: Umamaheswar Duvvuri, University of Pittsburgh, 200 Lothrop Street, W948 Biomedical Science Tower, Pittsburgh, PA 15213. Phone: 412-647-0954; Fax: 412-647-2080; E-mail: [email protected] doi: 10.1158/0008-5472.CAN-12-0475-T Ó2012 American Association for Cancer Research. Cancer Research Cancer Res; 72(13) July 1, 2012 3270 on November 1, 2020. © 2012 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from Published OnlineFirst May 7, 2012; DOI: 10.1158/0008-5472.CAN-12-0475-T

Upload: others

Post on 08-Aug-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: TMEM16A Induces MAPK and Contributes Directly …...Whitney rank-sum test was used to test the association of overexpression of TMEM16A with 11q13 amplification and TMEM16A expression

Molecular and Cellular Pathobiology

TMEM16A Induces MAPK and Contributes Directly toTumorigenesis and Cancer Progression

Umamaheswar Duvvuri1,7, Daniel J. Shiwarski1, Dong Xiao1, Carol Bertrand2, Xin Huang6, Robert S. Edinger5,JasonR.Rock10, BrianD.Harfe11, Brian J.Henson8,9, Karl Kunzelmann12, Rainer Schreiber12, RajaS. Seethala3,Ann Marie Egloff1, Xing Chen1, Vivian W. Lui1, Jennifer R. Grandis1,4, and Susanne M. Gollin1,3,8,9

AbstractFrequent gene amplification of the receptor-activated calcium-dependent chloride channel TMEM16A (TAOS2

or ANO1) has been reported in several malignancies. However, its involvement in human tumorigenesis has notbeen previously studied. Here, we show a functional role for TMEM16A in tumor growth. We found TMEM16Aoverexpression in 80% of head and neck squamous cell carcinoma (SCCHN), which correlated with decreasedoverall survival in patients with SCCHN. TMEM16A overexpression significantly promoted anchorage-independent growth in vitro, and loss of TMEM16A resulted in inhibition of tumor growth both in vitro andin vivo. Mechanistically, TMEM16A-induced cancer cell proliferation and tumor growth were accompanied by anincrease in extracellular signal–regulated kinase (ERK)1/2 activation and cyclin D1 induction. Pharmacologicinhibition of MEK/ERK and genetic inactivation of ERK1/2 (using siRNA and dominant-negative constructs)abrogated the growth effect of TMEM16A, indicating a role for mitogen-activated protein kinase (MAPK)activation in TMEM16A-mediated proliferation. In addition, a developmental small-molecule inhibitor ofTMEM16A, T16A-inh01 (A01), abrogated tumor cell proliferation in vitro. Together, our findings provide amechanistic analysis of the tumorigenic properties of TMEM16A, which represents a potentially noveltherapeutic target. The development of small-molecule inhibitors against TMEM16A may be clinically relevantfor treatment of human cancers, including SCCHN. Cancer Res; 72(13); 3270–81. �2012 AACR.

IntroductionAmplification of the chromosomal band 11q13 is frequently

seen in malignancies arising from breast, bladder, head andneck (SCCHN), and esophagus (1). Detailed molecular analysisof the 11q13 region led to identification of TMEM16A (alsoknown as TAOS2, DOG1, and ANO1; refs. 2–4). This genebelongs to the TMEM16 family, characterized by the presence

of 8 transmembrane domains and a highly conserved domainof unknown function (DUF590).

Recently, TMEM16A was shown to be a calcium-activatedchloride channel (CaCC; refs. 5–7). The biologic importance ofTmem16a is underscored by the lethal phenotype in the knock-out mouse because of abnormal tracheal development (8–10).TMEM16Acontributes tomany important physiologic functions,including the control of epithelial fluid transport, saliva produc-tion, and gastrointestinal tract motility (10–14). The third extra-cellular loop of TMEM16A (between transmembrane regions5 and 6) has been suggested to be the putative pore-formingdomain (6). Using mutagenesis approaches, the lysine residue atamino acid 610 (in the putative pore region) was found to becritical for its channel function (15). Specifically, mutation of thisresidue from lysine to alanine (TMEM16A-K610A) was shown tobe hypomorphic, with greatly reduced chloride conductance.

CaCCs can be proapoptotic and suppress tumor formationin breast epithelia (16). However, there are exceptions, theputative CaCC, bestrophin (BEST1) is upregulated in coloncancer cells, and can promote proliferation (17). In addition,several reports have shown that TMEM16A is overexpressed inmany tumor types including esophageal cancers, gastrointes-tinal stromal tumors, and SCCHN (18–20). These seeminglydiscordant findings prompted us to investigate the role ofTMEM16A in epithelial malignancies.

We initially determined the effects of TMEM16A on tumorproliferation (in vitro and in vivo) through gain- and loss-of-function experiments. Our results show that TMEM16A

Authors' Affiliations: Departments of 1Otolaryngology, 2Cell Biology andPhysiology, 3Pathology, 4Pharmacology andChemical Biology, and 5Renaland Electrolyte Division, University of Pittsburgh Medical Center; 6Depart-ment of Obstetrics, Gynecology, and Reproductive Sciences, University ofPittsburgh School of Medicine and Magee-Women's Research Institute,University of Pittsburgh Medical Center; 7Veterans Affairs PittsburghHealth System; 8University of Pittsburgh Cancer Institute; 9Human Genet-ics, University of Pittsburgh Graduate School of Public Health, Pittsburgh,Pennsylvania; 10Department of Cell Biology, Duke University MedicalCenter, Durham, North Carolina; 11Department of Molecular Genetics andMicrobiology and the Genetics Institute, University of Florida, Gainesville,Florida; and 12Department of Physiology, University of Regensburg,Regensburg, Germany

Note: Supplementary data for this article are available at Cancer ResearchOnline (http://cancerres.aacrjournals.org/).

Current address for J.R. Rock: Department of Anatomy, University ofCalifornia San Francisco.

Corresponding Author: Umamaheswar Duvvuri, University of Pittsburgh,200 Lothrop Street, W948 Biomedical Science Tower, Pittsburgh, PA15213. Phone: 412-647-0954; Fax: 412-647-2080; E-mail:[email protected]

doi: 10.1158/0008-5472.CAN-12-0475-T

�2012 American Association for Cancer Research.

CancerResearch

Cancer Res; 72(13) July 1, 20123270

on November 1, 2020. © 2012 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Published OnlineFirst May 7, 2012; DOI: 10.1158/0008-5472.CAN-12-0475-T

Page 2: TMEM16A Induces MAPK and Contributes Directly …...Whitney rank-sum test was used to test the association of overexpression of TMEM16A with 11q13 amplification and TMEM16A expression

induces potent and specific stimulation of the extracellularsignal–regulated kinase (ERK)1/2 (as determined by phospho-ERK1/2), and contributes to the growth of cancer cell lines.These studies provide, to the best of our knowledge, the firstmechanistic description of the role of TMEM16A in humanmalignancies and suggest that this protein may play an impor-tant role in facilitating tumor growth.

Materials and MethodsTMEM16A antibody and immunoblottingA rabbit polyclonal serum was obtained by immunizing

rabbits with an epitope (CARVLEKSLKKESRNKEKR) fromexon 14 of TMEM16A. This epitope was chosen on the basisof a BLAST search that defined a unique 21 amino-acidsequence. For immunoblotting, equal amounts of protein wereseparated on SDS-PAGE and transferred to nitrocellulosemembranes. The membranes were then probed with anti-TMEM16A serum, phospho-ERK1/2, pERK1/2, p-ERK5, ERK5,B-Raf, C-Raf, phospho-AKT Ser 472, AKT (Cell Signaling), andCyclin D1 M-20 (Santa Cruz). b-Tubulin or actin was used as aloading control. Ras activation kit was used according toinstructions (Millipore, EMD). All immunoblots were scannedat 600 dpi and postprocessed using Photoshop software (Ado-be Systems). Anymanipulation was applied to the entire imageto preserve image integrity.

Cell cultureHEK-293T cells were obtained from American Type Culture

Collection (ATCC; Manassas, VA). UM-SCC1 and T24 cellswere obtained from the University of Michigan (a gift fromDr Thomas Carey). EPC1 cells were a gift fromDrH. Nakagawa.All cell lines were genotyped to establish identity within 6months of experimentation. Stable overexpressing clones weremade using DNA transfection or retroviral infection. Cells wereselected after transduction and viable cells were pooled. Indi-vidual clones were identified by the method of limiting dilu-tions. Each clone was kept for 10 passages, after which an earlypassage sample was thawed and used.

Plasmid/siRNA transfections, retrovirus generation,shRNA transductionPlasmid transfections were conducted using Fugene (DNA)

or Lipofectamine 2000 (siRNA) according to the manufac-turer's instructions. TMEM16A and TMEM16A-K610A mutantwere subcloned into pBabe-puro vectors. Retroviruses weregenerated by transfecting PhoenixAmpho cells with theseplasmids. Lentiviral shRNA and retroviral particles were usedto transduce cells. Antibiotic selection was carried out. Alltransduction experiments were repeated at least 3 separatetimes to ensure reproducibility. GFP-tagged dominant nega-tive ERK2 construct was previously described (21). ERK1/2siRNA was obtained from Cell Signaling.

Whole-cell patch clampingWhole-cell patch clamping was conducted as previously

described (22). All pipette solutions were previously described(22). All experiments used a standard protocol that alternated acurrent–voltage (I/V) step measurement. The I/V measure-

ment stepped the holding potential from�100 toþ100 mV in20 mV steps. All patch clamp results were normalized by thecell capacitance recorded at the start of the experiment.

MQAE fluorescence assaysMQAE chloride efflux assays were conducted on cells plated

onto optical Petri dishes (Matek) precoated with poly-L-lysineas previously described (23). MQAE was introduced into thecells using a hypotonic shock followed by recovery for 10minutes before the start of the experiment. The MQAE-loadedcells were then mounted on the stage of an IX-81 Olympusmicroscope and perfused to 37�C. MQAE fluorescence inten-sity was captured every 15 seconds at the 445-nm wavelengthin response to excitation at 340 nm.

The magnitude of fluorescence in each cell in the field wasquantified froma circular region of interest (ROI) drawnwithinthe cell, and the time course offluorescence changewas plottedas the average � SEM of all ROIs in the field (typically 10–50cells, "n") for a single coverslip. The rate of change in fluores-cence upon the switch from high chloride to low chloride wasdetermined for each ROI. All ROIs exhibiting a positive, linearrate of change (R2 � 0.75) from at least 3 separate, identicalexperiments (coverslips, "N") were pooled, and statistical sig-nificance was assessed for all ROIs using Student's t test.

Cell viability assay and drug treatmentsFor proliferation and viability analysis, cells were plated

in 96-well optical plates at 5� 103 cells/well. The plates weretreated the following day, as indicated. One to 3 days aftertreatment, the CellTiter-Glo Assay (Promega) was usedaccording to the manufacturer's directions.

Soft agar assay, anchorage-independent viabilitySoft agar assays were conducted as previously described

(24). Briefly, 5� 104 cells suspended in 0.7% agar solution wereplated in a 35-mm dish on top of 1.4% agar. Colonies werecounted 3weeks after plating. Colonies with a diameter greaterthan 100 mm were counted using crystal violet. Anchorage-independent viability was determined by plating cells (5� 103)in poly-HEMA coated plates as previously described (25). Cellviability was assessed 3 days after plating.

In vivo growthAll animal studies were conducted under approval from the

University of Pittsburgh and were conducted in accordancewith established guidelines. Nude mice were injected on eachflank and subsequent tumor volumes were measured when apalpable tumor was noticed. Measurements of length andwidth were recorded and used to determine the volume ofeach tumor. At the conclusion of the experiment, tumors wereharvested and processed for further evaluation.

Primary tissue samples, tissue arrayPaired normal and tumor tissueswere collected after obtain-

ing informed consent and approval from our InstitutionalReview Board. Normal adjacent mucosa is defined as histo-logically benign appearing mucosa (as judged by an experi-enced pathologist) acquired from the margins of the tumor

Role of TMEM16A in Tumor Progression

www.aacrjournals.org Cancer Res; 72(13) July 1, 2012 3271

on November 1, 2020. © 2012 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Published OnlineFirst May 7, 2012; DOI: 10.1158/0008-5472.CAN-12-0475-T

Page 3: TMEM16A Induces MAPK and Contributes Directly …...Whitney rank-sum test was used to test the association of overexpression of TMEM16A with 11q13 amplification and TMEM16A expression

resection. Tissue arrays containing replicate cores werecreated from patients who underwent curative surgery forSCCHN at our institution. Staining was conducted withanti-TMEM16A antisera and scored using a semiquantitativesystem (H-score) and the results correlated with survival.H-score was defined as the relative intensity, scored on a 0 to 3scale, multiplied by the percentage of positively stained cells.The H-scores for the population were analyzed to determinethemedian score. High and low expressors were categorized ashaving H-scores above or below the median.

Oncomine analysisData were abstracted from the Oncomine database, and

used to assess the relative expression of TMEM16A in tumorsversus normal adjacent mucosa. We specifically evaluated theexpression the RefSeq version of TMEM16A: NM_018043.5. Afold-change of at least 3 was used as a cut-off value.

FISH studiesFISH studies were carried out on the tissuemicroarray using

a probe for the centromere of chromosome 11 (CEP11) labeledwith SpectrumGreen (AbbottMolecular) and a probe preparedfrom a BAC clone (RP11-805J14; CHRI) and labeled by nicktranslation with SpectrumOrange. Slide processing and scor-ing were described previously (26).

Knockout mouse experimentsTmem16a knockout mice were generated as previously

described (8, 9). Tissues were obtained by dissecting the oralcavity mucosa from mice after genotyping. Tissues were snapfrozen and subsequently used for immunochemistry, RT-PCR,and immunoblotting as described earlier.

Quantitative reverse transcription-PCRThe reverse transcriptions were carried out with prede-

signed TaqMan primer and probe pairs as described earlier(2). Reverse transcriptase (RT) controls were carried out foreach RNA input each time. Quantitative PCR (qPCR) wasconducted for TMEM16A and GAPDH was used as an endog-enous control. The primer and probe sequences, conditions,and concentration have been described previously (4).

Statistical analysisAll data are reported as mean � SEM unless stated other-

wise. Cell viability and tumor xenograft measurements wereanalyzed with a 2-tailed Student's t test. A paired t test wasused to test for differences in TMEM16A expression betweenmatched tumor and benign mucosa. A two-sided Mann–Whitney rank-sum test was used to test the association ofoverexpression of TMEM16A with 11q13 amplification andTMEM16A expression on IHC. Kaplan–Meier and log ranktests of equality of survivors were used to evaluate differencesin overall survival or disease-specific survival by high versuslow TMEM16A IHC scores, as defined by the median value.Multivariable Cox proportional hazards models were usedto evaluate TMEM16A tumor protein levels as a predictorof overall survival after adjusting for age, sex, and diseasestage. The assumption of proportional hazards was tested by

evaluation of Schoenfeld residuals. P < 0.05 was considered asstatistically significant.

ResultsTMEM16A overexpression correlates with decreasedsurvival in SCCHN

To examine the protein expression profile, rabbit polyclonalantibody to TMEM16A, was generated and used for immuno-blotting and IHC (Supplementary Fig. S1a and S1b). Asexpected, Tmem16a derived from murine tissues migrates ata higher molecular weight (�150 kDa) than human TMEM16A(�115 kDa). This is consistent with previously published data(6). We further validated this antibody's specificity in IHC byevaluating tissues obtained fromwild-type and knockoutmice.

TMEM16A protein is overexpressed by 5-fold in tumorswhen compared with paired adjacent normal mucosa (Fig.1A and B; P < 0.001). Overexpression was noted in 14/17 (83%)samples. Next, we confirmed that TMEM16A mRNA is over-expressed by 5-fold in a separate cohort of primary tumors andpaired adjacent normal mucosa (Fig. 1C; P < 0.05).

TMEM16A is endogenously expressed in benign secretorytissues such as salivary gland and breast tissue (14). Wetherefore wanted to confirm that breast malignancies alsooverexpress TMEM16A. We used the Oncomine database todetermine the expression of TMEM16A in normal and malig-nant tissues from a variety of tumor types. We found thatalthough TMEM16Amay be expressed at a high level in normalbreast tissue, its expression is even higher in neoplastic breasttissue (Fig. 1D). This suggests that although endogenousTMEM16A expression may be high in some normal tissues,malignant cells derived from these tissues further upregulateTMEM16A. This finding implicates TMEM16A as a potentialtarget gene in malignant transformation.

We then determined the association between TMEM16Aoverexpression and clinical outcome of patients with SCCHN.TMEM16A expression was detected in approximately 85% ofSCCHN tumors. Kaplan–Meier survival analysis showed thatpatients with high-level tumor expression of TMEM16A haddecreased overall survival (P¼ 0.04, log rank test; Fig. 1E). HighTMEM16A tumor levels tended to be associated withdecreased overall survival in a univariable Cox proportionalhazards model (HR ¼ 3.04; 95% CI, 0.97–9.46) and in amultivariable Cox proportional hazards model adjusting forage, sex, and nodal stage (HR ¼ 2.8; 95% CI, 0.86–9.16).

However, Kaplan–Meier analysis showed no significantcorrelation between overall or disease-free survival andTMEM16A gene amplification (Supplementary Fig. S2a andS2b). Although patients whose tumors harbored TMEM16Aamplification had a median survival of 50.7 months, whereasthose not amplified for TMEM16A did not reach a mediansurvival, suggesting that theremay be a trend toward improvedsurvival in patients without gene amplification. Gene ampli-fication was not strongly correlated with protein expression.

TMEM16A promotes tumor growth and proliferationNext, we investigated the effects of TMEM16Amanipulation

(through gain-of-function and loss-of-function) on the

Duvvuri et al.

Cancer Res; 72(13) July 1, 2012 Cancer Research3272

on November 1, 2020. © 2012 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Published OnlineFirst May 7, 2012; DOI: 10.1158/0008-5472.CAN-12-0475-T

Page 4: TMEM16A Induces MAPK and Contributes Directly …...Whitney rank-sum test was used to test the association of overexpression of TMEM16A with 11q13 amplification and TMEM16A expression

proliferation of cancer cells (Supplementary Fig. S3a and S3b).To evaluate the role of TMEM16A independently of 11q13amplification, we chose to use the UM-SCC1 SCCHN cell line(that harbors 11q13 amplification) and the T24 bladder cancercell line (that does not contain the 11q13 amplicon). LentiviralshRNA was used to "knock-down" TMEM16A in both cell lines.We identified 2 independent shRNA sequences that caused asignificant reduction (�80%) in TMEM16A protein expression.TMEM16A knockdown led to a measurable change in whole-cell chloride conductance (�50%), as assayed by MQAE fluo-rescence assays (Supplementary Fig. S3c–S3e; ref. 27). Forcedoverexpression of a mutant version of TMEM16A (TMEM16A-K610A) led to significantly smaller chloride conductance (Sup-plementary Fig. S3f). For further experiments, we used one ofthe sequences (shRNA#5).To determine the consequences of TMEM16A knockdown

on in vivo tumor growth, shRNA-treated UM-SCC1 cells were

inoculated subcutaneously into nude mice. TMEM16A shRNAled to a significant decrease in xenograft growth (Fig. 2A andB).Furthermore, we found that TMEM16A shRNA-treated tumorsexhibited decreased Ki-67 staining (Fig. 2C and D). Severalreports have showed that themitogen-activated protein kinase(MAPK)/ERK signaling pathway enhances the growth of epi-thelial cancer cells, in particular, bladder cancer, and SCCHN(28). We noted that pERK staining was less prominent intumors derived from TMEM16A shRNA-treated cells versuscontrol shRNA tumors (Fig. 2D).

Next, we exogenously overexpressed TMEM16A in severalcell lines that do not harbor endogenous gene amplification. InT24 cells, TMEM16A was trafficked to the cell surface (Sup-plementary Fig. S1b), led to increased chloride fluxes, andresulted in enhanced xenograft growth (Fig. 3A–C). Tumorxenografts confirmed overexpression of TMEM16A andshowed increased pERK1/2 and Ki-67 staining (Fig. 3D).

Figure 1. TMEM16A isoverexpressed in SCCHN andcorrelates with decreased survival.TMEM16A expression is upregulatedin 17 paired tumor (T) versus pairedadjacent normal mucosa (N) asdetected by immunoblotting (A).There is a 5-fold increase ofTMEM16A expression in tumorsrelative to normal adjacent mucosa(B; mean � SEM; n ¼ 17; ���, P <0.001). A 5-fold increase in tumor:normal TMEM16A mRNA wasobserved in a separate cohort(C; n ¼ 15; �, P < 0.05). TMEM16Aexpression data were abstractedfrom Oncomine and analyzed forrelative overexpression (D). HighTMEM16A expression wasassociated with decreased overallpatient survival (E; n ¼ 34, P < 0.05).Representative images of TMEM16Astaining are shown.

Role of TMEM16A in Tumor Progression

www.aacrjournals.org Cancer Res; 72(13) July 1, 2012 3273

on November 1, 2020. © 2012 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Published OnlineFirst May 7, 2012; DOI: 10.1158/0008-5472.CAN-12-0475-T

Page 5: TMEM16A Induces MAPK and Contributes Directly …...Whitney rank-sum test was used to test the association of overexpression of TMEM16A with 11q13 amplification and TMEM16A expression

TMEM16A overexpression led to increased in vitro prolifer-ation of HEK-293T cells, and increased anchorage-indepen-dent viability in immortalized keratinocytes (SupplementaryFig. S4a and S4b). TMEM16A knockdown in 2 other cell lines(Cal33 and PCI-15B) suppressed cell growth in vitro (Supple-mentary Fig. S4c). Furthermore, TMEM16A cooperated withconstitutively active H-RAS to induce focus formation inimmortalizedMEFs (Supplementary Fig. S4d). Similarly, forcedoverexpression of TMEM16A conferred increased colonogenicsurvival in immortalized keratinocytes and tumor cell lines(Supplementary Fig. S4e). Taken together, these data suggestthat TMEM16A overexpression can facilitate oncogenic trans-formation through cooperation with potent oncogenes.

We next determined the effect of TMEM16A knockdown onin vitro growth. Treatment with TMEM16A shRNA led to asignificant retardation in proliferation when compared withcontrol shRNA (Fig. 4A andB).TMEM16A shRNA abrogated theincrease in both anchorage-dependent and -independent cellproliferation noted in overexpressing cells (Fig. 4C and D).TMEM16A knockdown induced an accumulation of cells inG0/G1 and a concomitant decrease in the S/G2 phase, suggest-

ing a block in cell-cycle progression (Supplementary Fig. S5a).Interestingly, TMEM16A overexpression impaired caspase-3/7activation, suggesting that TMEM16A may in fact impairapoptotic cell death, along with promoting cell growth (Sup-plementary Fig. S5b).

TMEM16A induces phosphorylation of ERK1/2Because we observed differential expression of pERK1/2 in

tumor xenografts, we examinedwhether TMEM16A expressioninfluenced ERK1/2 activation. TMEM16A overexpression wasassociated with an increase in phosphorylated ERK1/2 (�2-fold) and cyclin D1 (�5-fold; Fig. 5A). TMEM16A siRNA led to adecrease in phospho-ERK1/2 and cyclin D1 (Fig. 5B). We notedthat AKT and phospho-AKT levels were not significantlyinfluenced by TMEM16A overexpression, suggesting specificityto the ERK1/2 pathway (Fig. 5B). Similarly, there were nochanges in p-ERK5 or ERK5 (Supplementary Fig. S6c).

It is well known that RAS oncogenic signaling can activateERK1/2. HRAS, in particular, has been associated withSCCHN development (29, 30). In fact, no mutations in KRASor NRAS were observed in 2 recent genomic studies of

BA

C

D

Scrambled

shRNA

TMEM16A

shRNA #5

Ki-67p-ERK1/2TMEM16A

Figure 2. TMEM16A knockdownresults in significant growthinhibition of SCCHN tumorxenografts in vivo. TMEM16Aknockdown in vivo led to smallertumors (A and B; �, P < 0.05).Xenografts generated fromTMEM16A shRNA cells showeddecreased proliferation (asdetermined by Ki-67 staining) andreduced pERK1/2 (C and D;�, P < 0.05). Representative imagesshowing TMEM16A, pERK1/2, andKi-67 staining are shown (n ¼ 6).Inset image shows �10magnification. hpf, high-powerfield.

Duvvuri et al.

Cancer Res; 72(13) July 1, 2012 Cancer Research3274

on November 1, 2020. © 2012 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Published OnlineFirst May 7, 2012; DOI: 10.1158/0008-5472.CAN-12-0475-T

Page 6: TMEM16A Induces MAPK and Contributes Directly …...Whitney rank-sum test was used to test the association of overexpression of TMEM16A with 11q13 amplification and TMEM16A expression

SCCHN (30, 31). We therefore evaluated the presence ofHRAS, KRAS, and NRAS mutations in the cell lines used inthis study. T24 cells harbor the activating HRASG12V muta-tion, however, HEK-293T, EPC1, and UM-SCC1 cells did notharbor mutant HRAS, KRAS, or NRAS (data not shown).Taken together, these data suggest that the impact ofTMEM16A on ERK1/2 signaling is independent of activatingHRAS mutations.Next, we explored the association between Tmem16a and

cyclinD1 in oral cavity tissues obtained from wild-type andTmem16a�/� embryos. As expected, Tmem16a levels werereduced in knockout mouse tissues as compared with wild-type and heterozygous mice (Fig. 5C). Interestingly, cyclinD1 was also decreased in tissues derived from knockoutmice, but not in their heterozygous or wild-type littermates.We observed intense nuclear cyclin D1 staining in the

proliferative mesoderm of wild-type mice but not in theTmem16a�/� littermates (Supplementary Fig. S6a). We fur-ther explored this finding by measuring mRNA expression ofTmem16a and Ccnd1 in these tissues. As expected, knockouttissues had significantly lower levels of cyclin D1 (Supple-mentary Fig. S6b). This finding suggested that TMEM16Amay be influencing cell proliferation on a fundamental leveland occurs in vivo.

We postulated that if TMEM16A affects proliferation byactivating MEK/ERK, inhibition of MEK/ERK should abro-gate TMEM16A-induced growth. Indeed, treatment witheither UO126 or dominant-negative ERK1/2 led to a com-plete abrogation of TMEM16A-induced growth (Fig. 5D andE). Similar data were observed with the specific ERK inhib-itor AZD6244 and with siRNA against ERK1/2 (Supplemen-tary Fig. S5c and S5d). Treatment with TMEM16A shRNA

Figure 3. TMEM16Aoverexpression (OE) induces in vivotumor growth. Overexpression ofTMEM16A markedly promotedtumor growth in the T24 xenograftmodel (A and B). TMEM16A-overexpressing cells showedincreased proliferation (asdetermined by Ki-67 staining)and pMAPK staining (C and D;��, P < 0.01). Increased levels ofp-ERK1/2 are also found in tissuederived from tumor explants (D).Inset image shows �10magnification. OE3 and OE7 refer toindividual clones. Ctl, control; hpf,high-power field.

A

C

B

DTMEM16A

Ctl

Ctl

50

40

30

20

10

0Ki-

67–p

osi

tive

nu

clei

/hp

f

TMEM16A

OE #3

TMEM16AOE #3

TMEM16A

OE #7

TMEM16AOE #7

Ki-67p-ERK1/2

Role of TMEM16A in Tumor Progression

www.aacrjournals.org Cancer Res; 72(13) July 1, 2012 3275

on November 1, 2020. © 2012 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Published OnlineFirst May 7, 2012; DOI: 10.1158/0008-5472.CAN-12-0475-T

Page 7: TMEM16A Induces MAPK and Contributes Directly …...Whitney rank-sum test was used to test the association of overexpression of TMEM16A with 11q13 amplification and TMEM16A expression

reversed the activation of MEK and ERK1/2 inducedby TMEM16A overexpression (Fig. 5F). This observationsuggests that TMEM16A overexpression directly impactsERK1/2 activation.

Recent data suggests that ERK activation may impact thechloride conductance of TMEM16A (15). Similarly, muta-tions in the putative pore forming domain can impact thewhole-cell chloride conductance in forced overexpressionexperiments (32). We sought to determine if the putativepore-forming region of TMEM16A was necessary for activa-tion of ERK1/2. Forced overexpression of a mutant versionof TMEM16A (TMEM16A-K610A) that has been describedto display significantly abrogated chloride conductance(15) did not induce ERK1/2 or phospho-ERK1/2 (Fig. 6A).We verified that TMEM16A-K610A was trafficked to thecell membrane using biotinylation experiments (data notshown). Our data raise the intriguing possibility that

TMEM16A affects ERK activation by modulating intracellu-lar chloride levels. Unfortunately, manipulation of intracel-lular chloride levels can itself induce changes in the expres-sion of ion channels. Therefore, further work is necessary todissect the effect of intracellular chloride on ERK activation.

Overexpression of mutant TMEM16A-K610A did not pro-mote anchorage-independent viability compared with vec-tor controls (Fig. 6B–D). To confirm that the effects areindeed directly dependent on TMEM16A expression, werescued TMEM16A shRNA-treated cells with shRNA-resis-tant versions of TMEM16A or TMEM16A-K610A (Supplemen-tary Fig. S7a). The reduction in viability induced byTMEM16A shRNA was rescued by expression of anshRNA-resistant version of TMEM16A, but not resistantTMEM16A-K610A (Fig. 6E). This approach provides strongevidence to show that TMEM16A directly induces the pro-liferative phenotype.

Figure 4. TMEM16A expressionmediates both anchorage-dependentand -independent growth. TMEM16Aknockdown caused a decrease inproliferation (mean � SEM; n ¼ 3;�, P < 0.05; A and B). Representativeimmunoblots are shown. The growthadvantage conferred by TMEM16Aoverexpression was abrogated byshRNA (C; mean � SEM; n ¼ 3,P < 0.05). Immunoblotting confirmsoverexpression (OE) and subsequentknockdown of TMEM16A (C). Softagar colony formation was alsoenhanced by TMEM16Aoverexpression and rescued bysubsequent knockdown (D; mean �SEM; n ¼ 4, P < 0.01).

Duvvuri et al.

Cancer Res; 72(13) July 1, 2012 Cancer Research3276

on November 1, 2020. © 2012 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Published OnlineFirst May 7, 2012; DOI: 10.1158/0008-5472.CAN-12-0475-T

Page 8: TMEM16A Induces MAPK and Contributes Directly …...Whitney rank-sum test was used to test the association of overexpression of TMEM16A with 11q13 amplification and TMEM16A expression

A B

C D

p-MEK

MEK

p-ERK1/2

ERK1/2

β-Tubulin

+ -shRNA

+ + TMEM16A -

-

T24F

E

EGFP DN-Erk20.00

0.25

0.50

0.75

1.00

1.25

1.50

Control

TMEM16A

**

*

Cel

l vi

abil

ity

(no

rmal

ized

)

β-Tubulin

GFP-DN-ERK

SCC1

p-ERK1/2

GFP

Control TMEM16A

EGFP DN-ERK2 EGFP DN-ERK2

Figure 5. Themitogenic phenotype of TMEM16A is associatedwith ERK1/2 activation. TMEM16Aoverexpression inducedphospho-ERK1/2 andcyclin D1 (A).TMEM16A-siRNA led to decreasedphospho-ERK1/2 and cyclin D1. Phospho-Akt levelswere not consistentlymodulated by TMEM16Amanipulation (B). Therelationship between TMEM16A and cyclin D1 in head and neck tissues derived from Tmem16a knockout (�/�) versuswild-type (þ/þ) or heterozygous (þ/�)mice (C). ERK inhibition with UO126 or dominant-negative ERK2 abrogated the growth advantage in overexpressing cells (D and E; mean � SEM; n ¼ 3;��, P < 0.01). Treatment of overexpressing cells with TMEM16A shRNA abrogated the activation MEK and ERK1/2 (F). Ctl, control.

Role of TMEM16A in Tumor Progression

www.aacrjournals.org Cancer Res; 72(13) July 1, 2012 3277

on November 1, 2020. © 2012 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Published OnlineFirst May 7, 2012; DOI: 10.1158/0008-5472.CAN-12-0475-T

Page 9: TMEM16A Induces MAPK and Contributes Directly …...Whitney rank-sum test was used to test the association of overexpression of TMEM16A with 11q13 amplification and TMEM16A expression

A

B

***

293T

293T T24

T24 T24 growth rescue

TMEM16A

Ctl/ DNA

TMEM16A-K610A

2.0

1.5

1.0

0.5

0.0

Ctl DNA

Ctl DNA

Ctl shRNA

TMEM16A DNA TMEM16A-K610A Ctl TMEM16A DNA

TMEM16A shRNA

TMEM16A DNA*

TMEM16A-K610A DNA*

TMEM16A-K610A

Ctl TMEM16A TMEM16A-K610A

An

ch

ora

ge

-in

de

pe

nd

en

t via

bili

ty

(no

rma

lize

d)

2.0

1.5

1.0

0.5

0.0An

ch

ora

ge

-in

de

pe

nd

en

t via

bili

ty

(no

rma

lize

d)

12.5

10.0

7.5

5.0

2.5

0.0

Ce

ll via

bili

ty

(fo

ld c

ha

ng

e)

2.0

1.5

1.0

0.5

0.0An

ch

ora

ge

in

de

pe

nd

en

t via

bili

ty

(no

rma

lize

d)

TMEM16A

p-ERK1/2

ERK1/2

p-Akt

Akt

Cyclin D1

β-Tubulin

SCC1

SCC1

***

***

C

D E

Figure 6. The putative pore-forming region of TMEM16Amediates ERK1/2 activation and subsequent mitogenic effects. Expression of the TMEM16A-K610Amutant fails to induceERK1/2 and cyclin D1 (A). TMEM16A-K610Adid not increase anchorage-independent viability (B–D;mean�SEM; n¼ 3; ���,P < 0.001).TMEM16A shRNA–induced decrease in growthwas subsequently rescuedwith TMEM16A shRNA–resistant DNA (TMEM16ADNA�), but not with TMEM16A-K610A shRNA–resistant DNA (TMEM16A-K610A DNA�; E; mean � SEM; n ¼ 3, P < 0.05). Ctl, control.

Duvvuri et al.

Cancer Res; 72(13) July 1, 2012 Cancer Research3278

on November 1, 2020. © 2012 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Published OnlineFirst May 7, 2012; DOI: 10.1158/0008-5472.CAN-12-0475-T

Page 10: TMEM16A Induces MAPK and Contributes Directly …...Whitney rank-sum test was used to test the association of overexpression of TMEM16A with 11q13 amplification and TMEM16A expression

TMEM16A activates the Ras-Raf-MEK-ERK pathwayThe observation that TMEM16A induces activation of

ERK1/2 led us to interrogate the RAS-RAF-MEK-ERK1/2pathway. Indeed we observed robust activation of RAS-RAF-MEK-ERK pathway upon TMEM16A overexpression(Fig. 7A and B). Densitometric quantitation is shown inSupplementary Fig. S7b. Interestingly, this activation wasindependent of HRAS mutation status, because T24 cellsharbor an activating HRAS mutation, but SCC1 cells do not(data not shown). We next sought to determine if Rasinhibition (using a dominant negative construct) couldabrogate the observed phenotype. Forced overexpression ofa dominant-negative mutant of H-Ras (H-RasN17) abrogatedthe observed activation of MEK, ERK1/2 and also the growthphenotype (Fig. 7C and D).

Pharmacologic inhibition of TMEM16A induces cancercell deathTo validate TMEM16A as a potential target in epithelial

malignancies, we wanted to determine if small-molecule inhi-bition of TMEM16A could inhibit tumor cell proliferation.

Therefore, we treated UM-SCC1 and T24 cells with a novelTMEM16A inhibitor (T16A-inh01; refs. 33, 34). T16A-inh01induced a dose-dependent reduction in cell viability (Supple-mentary Fig. S8a and S8B). To determine whether combinedinhibition of TMEM16A and MAPK had an additive effect oncell viability, we treated cells with the MEK/ERK inhibitorUO126 alone or in combination with TMEM16A shRNA.TMEM16A knockdown resulted in a modest decrease in cellproliferation, however, we observed an additive effect withMAPK inhibition (Supplementary Fig. S8c and S8d). Thechemical structure of T16A-inh01 is shown in SupplementaryFig. S8e. We did not observe additive inhibitory effects on ERKphosphorylation. These data show that TMEM16A and MAPKinhibition may act in an additive fashion to retard tumor cellproliferation, suggesting that ERK1/2may not uniquely controlTMEM16A-induced cell growth.

DiscussionThere is accumulating evidence that chloride channels

influence tumor growth and progression (35–37), howeverthe mechanism(s) by which this occurs remains unclear.Calcium-activated chloride channels (CaCCs) are a uniquesubset of chloride channels that play important roles inmany fundamental physiologic processes (38, 39). Recently,ANO1/TMEM16A was described as a bona fide CaCC (5–7).However, it remains controversial whether TMEM16A isitself a functional CaCC, or forms a subunit within theprotein complex that facilitates CaCC activity (7, 40, 41).Recent reports suggest that CaCCs can both promote andretard tumor cell proliferation (16, 17). These contradictoryreports suggest that improved understanding of the impactof CaCC regulation and activation on cell proliferation mayhelp to define whether these molecules can serve as a futuretherapeutic target.

TMEM16A is frequently overexpressed in several tumorsincluding squamous cell carcinoma of the head and neck,esophageal cancer, and gastrointestinal stromal tumors (GIST;refs. 4, 18, 42, 43). To date, the exact role(s) that TMEM16Aplays in tumor development/progression remains unclear.Ayoub and colleagues have recently reported that TMEM16Aoverexpression facilitates cell motility and may contribute tothe development of metastases (44); however, no mechanismwas proposed to explain this phenotype.

This report provides the first mechanistic link betweenTMEM16A expression and cell proliferation in human cancer.Our data show that TMEM16A expression directly impactscellular proliferation. TMEM16A also cooperates with onco-genic H-RAS to induce focus formation in immortalized MEFs.Taken together, these data strongly suggest that TMEM16Amay function as a proto-oncogene, and that its overexpressiondrive tumor growth.

It has recently been shown that TMEM16A channel activityis linked to ERK1/2 activation and that the ERK inhibitorUO126 can inhibit TMEM16A channel function (15). Our datasuggest that TMEM16A activates the RAS-RAF-MEK-ERK1/2pathway, thereby influencing cellular proliferation. However,this activation does not occur when a hypomorphic mutant

C

PCDNA3.1-p

-EGFP

PCDNA3.1-R

asN17

TMEM16A-p

-EGFP

TMEM16A-R

asN17

0

25

50

75

100

125

150

17524 h

48 h

* *

*

*

* *

Per

cen

t o

f su

rviv

al

D T24

BControl TMEM16A

SCC1

p-MEK

MEK

p-ERK1/2

ERK1/2

β-Tubulin

Cyclin D1

S445p-B-Raf

S338p-C-Raf

Ras

Ras-GTP

p-MEK

MEK

p-ERK1/2

ERK1/2

β-Tubulin

Control TMEM16A

p-EGFP + - + -RasN17 - + - +

Cyclin D1

T24

GFP

AControl TMEM16A

T24

p-MEK

MEK

p-ERK1/2

ERK1/2

β-Tubulin

Cyclin D1

Ras-GTP

S445p-B-Raf

S338p-C-Raf

Ras

Figure 7. TMEM16A expression activates the Ras-Raf-MEK-ERK1/2pathway. Activation of Ras, B-Raf, and C-Raf was observed in T24 andUM-SCC1 cells after TMEM16A overexpression (A and B). Activation ofthe pathway was abrogated by forced expression of dominant-negativeRas (C). Forced expression of dominant-negative Ras abrogated thegrowth advantage conferred by TMEM16A overexpression (D).

Role of TMEM16A in Tumor Progression

www.aacrjournals.org Cancer Res; 72(13) July 1, 2012 3279

on November 1, 2020. © 2012 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Published OnlineFirst May 7, 2012; DOI: 10.1158/0008-5472.CAN-12-0475-T

Page 11: TMEM16A Induces MAPK and Contributes Directly …...Whitney rank-sum test was used to test the association of overexpression of TMEM16A with 11q13 amplification and TMEM16A expression

(TMEM16A-K610A) is expressed. These data raise the intrigu-ing possibility that the chloride conductance of TMEM16Aimpacts ERK1/2 activation. However, the exact mechanism bywhich this occurs remains unclear.

The availability of small-molecule inhibitors againstTMEM16A provides a novel potential method to inhibit tumorcell growth. The currently available small-molecule inhibitorsof TMEM16A (niflumic acid and NPPB) exhibit off-targeteffects and have been shown to block other chloride channels.However, T16A-inh01 is a novel and potentially more specificTMEM16A inhibitor, and provides amethod to inhibit channelactivity (33, 34, 45). This developmental molecule likely has off-target effects, and therefore definitive conclusions cannot bedefined at this time.

The ubiquitous expression of TMEM16A suggests that theendogenous protein has important physiologic roles that maybe adversely impacted by pharmacologic inhibition. However,TMEM16A is known to undergo alternative splicing, andspecific variants have been isolated fromdiseased tissues (suchas diabetic gastroparesis; ref. 46–48). Recently, mutations inTMEM16A have been described in SCCHN. The existence ofspecificmutations inTMEM16A that have been identified fromwhole-exome sequencing of SCCHN, suggests that tumor-specific targeting may be possible in the future (30).

In conclusion, we have shown that TMEM16Aexpression: (1)is associated with decreased patient survival; (2) modulatesproliferation in vitro and in vivo and induces ERK1/2 activation;(3) the pore-forming region of this molecule is necessary tofacilitate these phenomena; and (4) pharmacologic inhibition(using a novel TMEM16A inhibitor) or siRNA-mediated knock-down of TMEM16A leads to a significant decrease in tumor cellviability by downregulating the TMEM16A-ERK1/2 signaling.In closing, these findings may describe a novel therapeutic role

for TMEM16A in cancer treatment, because TMEM16A mayact as a potential pharmacologic target.

Disclosure of Potential Conflicts of InterestNo potential conflicts of interest were disclosed.

Authors' ContributionsConception and design: U. Duvvuri, D.J. Shiwarski, V.W. Lui, S.M. GollinDevelopment of methodology: U. Duvvuri, D.J. Shiwarski, D. Xiao, X. HuangAcquisition of data (provided animals, acquired and managed patients,provided facilities, etc.): U. Duvvuri, D.J. Shiwarski, D. Xiao, C. Bertrand, R.S.Edinger, J.R. Rock, R.D. Harfe, B.J. Henson, R.S. Seethala, A.M. EgloffAnalysis and interpretation of data (e.g., statistical analysis, biostatistics,computational analysis): U. Duvvuri, D.J. Shiwarski, D. Xiao, C. Bertrand, R.S.Edinger, R.S. Seethala, A.M. Egloff, X. Chen, J.R. GrandisWriting, review, and/or revision of the manuscript: U. Duvvuri, D.J.Shiwarski, C. Bertrand, X. Huang, K. Kunzelmann, X. Chen, V.W. Lui, J.R. Grandis,S.M. GollinAdministrative, technical, or material support (i.e., reporting or orga-nizing data, constructing databases): U. Duvvuri, K. Kunzelmann, R. Schrei-ber, R.S. Seethala, X. ChenStudy supervision: U. Duvvuri, S.M. Gollin

AcknowledgmentsThe authors thank Dr. R.W. Sobol and the University of Pittsburgh Cancer

Institute Lentiviral Facility for the generation of lentiviral particles.

Grant SupportThis work was supported by the Head and Neck Cancer SPORE grant (NIH

P50CA097190 J.R. Grandis), through a Veterans Administration CDA; CPPF (U.Duvvuri) and start-up funds from the Department of Otolaryngology (U. Duv-vuri). The molecular cytogenetic studies were carried out in the University ofPittsburgh Cancer Institute Cytogenetics Facility, supported in part byP30CA047904 to N.E. Davidson/R.B. Herberman/S.M. Gollin. KK was supportedby Deutsche Krebshilfe Projekt 109438.

The costs of publication of this article were defrayed in part by the payment ofpage charges. This article must therefore be hereby marked advertisement inaccordance with 18 U.S.C. Section 1734 solely to indicate this fact.

Received February 8, 2012; revised April 26, 2012; accepted April 28, 2012;published OnlineFirst May 7, 2012.

References1. Akervall JA, Jin Y, Wennerberg JP, Zatterstrom UK, Kjellen E, Mertens

F, et al. Chromosomal abnormalities involving 11q13 are associatedwith poor prognosis in patients with squamous cell carcinoma of thehead and neck. Cancer 1995;76:853–9.

2. Huang X, Gollin SM, Raja S, Godfrey TE. High-resolution mapping ofthe 11q13 amplicon and identification of a gene, TAOS1, that isamplified and overexpressed in oral cancer cells. Proc Natl Acad SciUSA 2002;99:11369–74.

3. Katoh M, Katoh M. FLJ10261 gene, located within the CCND1-EMS1locus on human chromosome 11q13, encodes the eight-transmem-brane protein homologous to C12orf3, C11orf25 and FLJ34272 geneproducts. Int J Oncol 2003;22:1375–81.

4. Huang X, Godfrey TE, Gooding WE, McCarty KS Jr, Gollin SM.Comprehensive genome and transcriptome analysis of the 11q13amplicon in human oral cancer and synteny to the 7F5 ampliconin murine oral carcinoma. Genes Chromosomes Cancer 2006;45:1058–69.

5. Caputo A, Caci E, Ferrera L, Pedemonte N, Barsanti C, Sondo E, et al.TMEM16A, a membrane protein associated with calcium-dependentchloride channel activity. Science 2008;322:590–4.

6. Yang YD, Cho H, Koo JY, Tak MH, Cho Y, Shim WS, et al. TMEM16Aconfers receptor-activated calcium-dependent chloride conductance.Nature 2008;455:1210–5.

7. Schroeder BC, Cheng T, Jan YN, Jan LY. Expression cloning ofTMEM16A as a calcium-activated chloride channel subunit. Cell2008;134:1019–29.

8. Rock JR, Futtner CR, Harfe BD. The transmembrane proteinTMEM16A is required for normal development of the murine trachea.Dev Biol 2008;321:141–9.

9. Rock JR, Harfe BD. Expression of TMEM16 paralogs during murineembryogenesis. Dev Dyn 2008;237:2566–74.

10. Rock JR, O'Neal WK, Gabriel SE, Randell SH, Harfe BD, BoucherRC, et al. Transmembrane protein 16A (TMEM16A) is a Ca2þ-reg-ulated Cl� secretory channel in mouse airways. J Biol Chem2009;284:14875–80.

11. Huang F, Rock JR, Harfe BD, Cheng T, Huang X, Jan YN, et al. Studieson expression and function of the TMEM16A calcium-activated chlo-ride channel. Proc Natl Acad Sci U S A 2009;106:21413–8.

12. Hwang SJ, Blair PJ, Britton FC, O'Driscoll KE, Hennig G, BayguinovYR, et al. Expression of anoctamin 1/TMEM16A by interstitial cells ofCajal is fundamental for slowwave activity in gastrointestinal muscles.J Physiol 2009;587:4887–904.

13. Gomez-Pinilla PJ, Gibbons SJ, Bardsley MR, Lorincz A, Pozo MJ,Pasricha PJ, et al. Ano1 is a selective marker of interstitial cells of Cajalin the human and mouse gastrointestinal tract. Am J Physiol Gastro-intest Liver Physiol 2009;296:G1370–81.

14. Romanenko VG, Catalan MA, Brown DA, Putzier I, Hartzell HC, Mar-morstein AD, et al. Tmem16A encodes theCa2þ-activatedCl� channelin mouse submandibular salivary gland acinar cells. J Biol Chem2010;285:12990–3001.

15. Almaca J, Tian Y, Aldehni F, Ousingsawat J, Kongsuphol P, Rock JR,et al. TMEM16 proteins produce volume-regulated chloride currents

Duvvuri et al.

Cancer Res; 72(13) July 1, 2012 Cancer Research3280

on November 1, 2020. © 2012 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Published OnlineFirst May 7, 2012; DOI: 10.1158/0008-5472.CAN-12-0475-T

Page 12: TMEM16A Induces MAPK and Contributes Directly …...Whitney rank-sum test was used to test the association of overexpression of TMEM16A with 11q13 amplification and TMEM16A expression

that are reduced in mice lacking TMEM16A. J Biol Chem 2009;284:28571–8.

16. Elble RC, Pauli BU. Tumor suppression by a proapoptotic calcium-activated chloride channel in mammary epithelium. J Biol Chem2001;276:40510–7.

17. SpitznerM,Martins JR, Soria RB, Ousingsawat J, Scheidt K, SchreiberR, et al. Eag1 and Bestrophin 1 are up-regulated in fast-growingcolonic cancer cells. J Biol Chem 2008;283:7421–8.

18. Carles A, Millon R, Cromer A, Ganguli G, Lemaire F, Young J, et al.Head and neck squamous cell carcinoma transcriptome analysis bycomprehensive validated differential display. Oncogene 2006;25:1821–31.

19. Espinosa I, Lee CH, KimMK, Rouse BT, Subramanian S, MontgomeryK, et al. A novel monoclonal antibody against DOG1 is a sensitive andspecific marker for gastrointestinal stromal tumors. Am J Surg Pathol2008;32:210–8.

20. KashyapMK,MarimuthuA,KishoreCJ, Peri S, Keerthikumar S,PrasadTS, et al. Genomewide mRNA profiling of esophageal squamous cellcarcinoma for identification of cancer biomarkers. Cancer Biol Ther2009;8:36–46.

21. Dagda RK, Zhu J, Kulich SM, Chu CT. Mitochondrially localized ERK2regulates mitophagy and autophagic cell stress: implications for Par-kinson's disease. Autophagy 2008;4:770–82.

22. Sun F, Mi Z, Condliffe SB, Bertrand CA, Gong X, Lu X, et al. Chaperonedisplacement from mutant cystic fibrosis transmembrane conduc-tance regulator restores its function in human airway epithelia. FASEBJ 2008;22:3255–63.

23. Silvis MR, Bertrand CA, Ameen N, Golin-Bisello F, Butterworth MB,Frizzell RA, et al. Rab11b regulates the apical recycling of the cysticfibrosis transmembrane conductance regulator in polarized intestinalepithelial cells. Mol Biol Cell 2009;20:2337–50.

24. Koppikar P, Lui VW, Man D, Xi S, Chai RL, Nelson E, et al. Constitutiveactivation of signal transducer and activator of transcription 5 con-tributes to tumor growth, epithelial-mesenchymal transition, and resis-tance to epidermal growth factor receptor targeting. Clin Cancer Res2008;14:7682–90.

25. Fukazawa H, Mizuno S, Uehara Y. A microplate assay for quantitationof anchorage-independent growth of transformed cells. Anal Biochem1995;228:83–90.

26. ReshmiSC,HuangX,SchoppyDW,BlackRC,SaundersWS, SmithDI,et al. Relationship between FRA11F and 11q13 gene amplification inoral cancer. Genes Chromosomes Cancer 2007;46:143–54.

27. Ousingsawat J, Martins JR, Schreiber R, Rock JR, Harfe BD, Kunzel-mann K. Loss of TMEM16A causes a defect in epithelial Ca2þ-depen-dent chloride transport. J Biol Chem 2009;284:28698–703.

28. MacLaineNJ,WoodMD,Holder JC,ReesRW,Southgate J. Sensitivityof normal, paramalignant, and malignant human urothelial cells toinhibitors of the epidermal growth factor receptor signaling pathway.Mol Cancer Res 2008;6:53–63.

29. Rathcke IO, Gottschlich S, Gorogh T, Lippert BM, Werner JA.[Incidence of point mutations in Ki-ras codon 12 and 13 in squamousepithelial carcinomas of the head-neck region]. Laryngorhinootologie1996;75:465–70.

30. Stransky N, Egloff AM, Tward AD, Kostic AD, Cibulskis K, SivachenkoA, et al. The mutational landscape of head and neck squamous cellcarcinoma. Science 2011;333:1157–60.

31. Agrawal N, Frederick MJ, Pickering CR, Bettegowda C, Chang K, LiRJ, et al. Exome sequencing of head and neck squamous cell

carcinoma reveals inactivating mutations in NOTCH1. Science2011;333:1154–7.

32. Kunzelmann K, Kongsuphol P, Chootip K, Toledo C, Martins JR,Almaca J, et al. Role of the Ca2þ-activated Cl� channels bestrophinand anoctamin in epithelial cells. Biol Chem 2011;392:125–34.

33. Namkung W, Phuan PW, Verkman AS. TMEM16A inhibitors revealTMEM16A as aminor component of CaCC conductance in airway andintestinal epithelial cells. J Biol Chem 2011;286:2365–74.

34. Namkung W, Thiagarajah JR, Phuan PW, Verkman AS. Inhibition ofCa2þ-activated Cl� channels by gallotannins as a possible molecularbasis for health benefits of red wine and green tea. FASEB J2010;24:4178–86.

35. Habela CW, Ernest NJ, Swindall AF, Sontheimer H. Chloride accumu-lation drives volume dynamics underlying cell proliferation and migra-tion. J Neurophysiol 2009;101:750–7.

36. Habela CW, OlsenML, Sontheimer H. ClC3 is a critical regulator of thecell cycle in normal and malignant glial cells. J Neurosci 2008;28:9205–17.

37. Sontheimer H. An unexpected role for ion channels in brain tumormetastasis. Exp Biol Med (Maywood) 2008;233:779–91.

38. Galietta LJ. The TMEM16 protein family: a new class of chloridechannels? Biophys J 2009;97:3047–53.

39. Hartzell HC, Yu K, Xiao Q, Chien LT, Qu Z. Anoctamin/TMEM16 familymembers are Ca2þ-activated Cl� channels. J Physiol 2009;587:2127–39.

40. Sheridan JT, Worthington EN, Yu K, Gabriel SE, Hartzell HC, Tarran R.Characterization of the oligomeric structure of the Ca2þ-activated Cl�

channel Ano1/TMEM16A. J Biol Chem 2011;286:1381–8.41. FallahG,Roemer T, Detro-DassenS,BraamU,Markwardt F, Schmalz-

ing G. TMEM16A(a)/anoctamin-1 shares a homodimeric architecturewith CLC chloride channels. Mol Cell Proteomics 2011;10:M110.004697.

42. Carneiro A, Isinger A, Karlsson A, Johansson J, Jonsson G, BendahlPO, et al. Prognostic impact of array-based genomic profiles inesophageal squamous cell cancer. BMC Cancer 2008;8:98.

43. West RB, CorlessCL, ChenX, Rubin BP, Subramanian S,MontgomeryK, et al. The novel marker, DOG1, is expressed ubiquitously in gas-trointestinal stromal tumors irrespective of KIT or PDGFRA mutationstatus. Am J Pathol 2004;165:107–13.

44. Ayoub C, Wasylyk C, Li Y, Thomas E, Marisa L, Robe A, et al. ANO1amplification and expression in HNSCC with a high propensity forfuture distant metastasis and its functions in HNSCC cell lines. Br JCancer 2010;103:715–26.

45. Tian Y, Kongsuphol P, Hug M, Ousingsawat J, Witzgall R, Schrei-ber R, et al. Calmodulin-dependent activation of the epithelialcalcium-dependent chloride channel TMEM16A. FASEB J 2011;25:1058–68.

46. Ferrera L,CaputoA,Ubby I, Bussani E, Zegarra-MoranO,RavazzoloR,et al. Regulation of TMEM16A chloride channel properties by alterna-tive splicing. J Biol Chem 2009;284:33360–8.

47. Ferrera L, Scudieri P, Sondo E, Caputo A, Caci E, Zegarra-Moran O,et al. A minimal isoform of the TMEM16A protein associated withchloride channel activity. Biochim Biophys Acta 2011;1808:2214–23.

48. MazzoneA,BernardCE, StregePR,BeyderA,Galietta LJ, PasrichaPJ,et al. Altered expression of Ano1 variants in human diabetic gastro-paresis. J Biol Chem 2011;286:13393–403.

Role of TMEM16A in Tumor Progression

www.aacrjournals.org Cancer Res; 72(13) July 1, 2012 3281

on November 1, 2020. © 2012 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Published OnlineFirst May 7, 2012; DOI: 10.1158/0008-5472.CAN-12-0475-T

Page 13: TMEM16A Induces MAPK and Contributes Directly …...Whitney rank-sum test was used to test the association of overexpression of TMEM16A with 11q13 amplification and TMEM16A expression

2012;72:3270-3281. Published OnlineFirst May 7, 2012.Cancer Res   Umamaheswar Duvvuri, Daniel J. Shiwarski, Dong Xiao, et al.   Tumorigenesis and Cancer ProgressionTMEM16A Induces MAPK and Contributes Directly to

  Updated version

  10.1158/0008-5472.CAN-12-0475-Tdoi:

Access the most recent version of this article at:

  Material

Supplementary

  http://cancerres.aacrjournals.org/content/suppl/2012/05/07/0008-5472.CAN-12-0475-T.DC1

Access the most recent supplemental material at:

   

   

  Cited articles

  http://cancerres.aacrjournals.org/content/72/13/3270.full#ref-list-1

This article cites 47 articles, 19 of which you can access for free at:

  Citing articles

  http://cancerres.aacrjournals.org/content/72/13/3270.full#related-urls

This article has been cited by 29 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://cancerres.aacrjournals.org/content/72/13/3270To request permission to re-use all or part of this article, use this link

on November 1, 2020. © 2012 American Association for Cancer Research. cancerres.aacrjournals.org Downloaded from

Published OnlineFirst May 7, 2012; DOI: 10.1158/0008-5472.CAN-12-0475-T