mutualinfluenceofros,ph,andclic1membrane protein in the ...marta peretti1, federica maddalena...

12
Cancer Biology and Translational Studies Mutual Inuence of ROS, pH, and CLIC1 Membrane Protein in the Regulation of G 1 S Phase Progression in Human Glioblastoma Stem Cells Marta Peretti 1 , Federica Maddalena Raciti 1 , Valentina Carlini 1 , Ivan Verduci 1 , Sarah Sertic 1 , Sara Barozzi 2 , Massimiliano Garr e 2 , Alessandra Pattarozzi 3 , Antonio Daga 4 , Federica Barbieri 3 , Alex Costa 1 , Tullio Florio 3,4 , and Michele Mazzanti 1 Abstract Glioblastoma (GB) is the most lethal, aggressive, and dif- fuse brain tumor. The main challenge for successful treatment is targeting the cancer stem cell (CSC) subpopulation respon- sible for tumor origin, progression, and recurrence. Chloride Intracellular Channel 1 (CLIC1), highly expressed in CSCs, is constitutively present in the plasma membrane where it is associated with chloride ion permeability. In vitro, CLIC1 inhibition leads to a signicant arrest of GB CSCs in G 1 phase of the cell cycle. Furthermore, CLIC1 knockdown impairs tumor growth in vivo. Here, we demonstrate that CLIC1 mem- brane localization and function is specic for GB CSCs. Mes- enchymal stem cells (MSC) do not show CLIC1-associated chloride permeability, and inhibition of CLIC1 protein func- tion has no inuence on MSC cell-cycle progression. Investi- gation of the basic functions of GB CSCs reveals a constitutive state of oxidative stress and cytoplasmic alkalinization com- pared with MSCs. Both intracellular oxidation and cyto- plasmic pH changes have been reported to affect CLIC1 membrane functional expression. We now report that in CSCs these three elements are temporally linked during CSC G 1 S transition. Impeding CLIC1-mediated chloride current pre- vents both intracellular ROS accumulation and pH changes. CLIC1 membrane functional impairment results in GB CSCs resetting from an allostatic tumorigenic condition to a homeo- static steady state. In contrast, inhibiting NADPH oxidase and NHE1 proton pump results in cell death of both GB CSCs and MSCs. Our results show that CLIC1 membrane protein is crucial and specic for GB CSC proliferation, and is a prom- ising pharmacologic target for successful brain tumor thera- pies. Mol Cancer Ther; 17(11); 245161. Ó2018 AACR. Introduction Gliomas are the major group of primary malignant brain tumors. Among them, glioblastoma (GB, WHO grade IV) is the most aggressive and lethal (1, 2). To date, no effective therapies for GB have been developed. Surgical removal of GB is often ineffective due to its fast spread and high rate of inltration in brain parenchyma. Chemotherapy is also problematic due to the presence of the bloodbrain barrier and the high expression of antiapoptotic proteins and drug efux transporters on cancer cells. Current pharmacologic treatments for GB target highly prolifer- ating but nontumorigenic cancer bulk cells but spare the tumor- igenic cancer stem cells (CSC), which, being refractory to therapy, determine the poor prognosis of GB patients. Although usually quiescent, CSCs are the reservoir for a potential tumor recurrence (3, 4). Isolated CSCs, which contain all the necessary genetic information to faithfully reproduce the original tumor, can then be used as a model to look for new diagnostic markers and screen potential drugs. In addition, therapies directed specically against CSCs could offer better clinical outcomes (4). The greatest chal- lenge of GB treatment is likely connected to the need to prefer- entially target CSCs. We have previously demonstrated the pivotal role of CLIC1 protein in the tumorigenic potential of CSCs isolated from human GBs, and showed that CLIC1 functional expression in cell mem- branes is essential for the proliferation and self-renewal of GB CSCs (5, 6). CLIC1 was cloned from a human monocytic cell line (7). First identied on nuclear membranes, it was then also found on plasma membranes (7, 8). This protein is highly conserved among vertebrates, ubiquitously expressed in different tissues, and overexpressed in tumors (9, 10). CLIC1 has been character- ized as a metamorphic protein (11), coexisting as both a soluble cytoplasmic form and a membrane-associated element. The tran- sition between its hydrophilic and hydrophobic localizations is promoted by several factors. In vitro studies on an articial lipid bilayer and isolated cells have demonstrated that changes in both 1 Dipartimento di Bioscienze, Universit a degli Studi di Milano, Milan, Italy. 2 Fondazione Istituto FIRC di Oncologia Molecolare, Milan, Italy and Cogentech S.c.a.r.l., IFOM Via Adamello, Milan, Italy. 3 Sezione di Farmacologia, Dipartimento di Medicina Interna & Centro di Eccellenza per la Ricerca Biomedica (CEBR), Universit a di Genova, Genova, Italy. 4 IRCCS Policlinico San Martino and Dipartimento delle Terapie Oncologiche Integrate, Ospedale San Martino, Genova, Italy. Note: Supplementary data for this article are available at Molecular Cancer Therapeutics Online (http://mct.aacrjournals.org/). M. Peretti and F.M. Raciti are the corst authors of this article. T. Florio and M. Mazzanti contributed equally to this article. Current address for M. Peretti: Laboratory of Cell and Molecular Physiology, EA4667, University of Picardie Jules Verne, Amiens, France. Current address for F.M. Raciti: Department of Physiology and Biophysics, University of Miami Miller School of Medicine, 1600 NW 10th Ave, Miami, FL 33136. Corresponding Author: Michele Mazzanti, University of Milan, Via Celoria 26, 20133 Milano, Italy. Phone: 39-0250314958; E-mail: [email protected] doi: 10.1158/1535-7163.MCT-17-1223 Ó2018 American Association for Cancer Research. Molecular Cancer Therapeutics www.aacrjournals.org 2451 on August 11, 2021. © 2018 American Association for Cancer Research. mct.aacrjournals.org Downloaded from Published OnlineFirst August 22, 2018; DOI: 10.1158/1535-7163.MCT-17-1223

Upload: others

Post on 12-Mar-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: MutualInfluenceofROS,pH,andCLIC1Membrane Protein in the ...Marta Peretti1, Federica Maddalena Raciti1,Valentina Carlini1, Ivan Verduci1, Sarah Sertic1, Sara Barozzi2, Massimiliano

Cancer Biology and Translational Studies

Mutual Influence of ROS, pH, and CLIC1 MembraneProtein in the Regulation of G1–S PhaseProgression in Human Glioblastoma Stem CellsMarta Peretti1, Federica Maddalena Raciti1, Valentina Carlini1, Ivan Verduci1,Sarah Sertic1, Sara Barozzi2, Massimiliano Garr�e2, Alessandra Pattarozzi3,Antonio Daga4, Federica Barbieri3, Alex Costa1, Tullio Florio3,4, and Michele Mazzanti1

Abstract

Glioblastoma (GB) is the most lethal, aggressive, and dif-fuse brain tumor. The main challenge for successful treatmentis targeting the cancer stem cell (CSC) subpopulation respon-sible for tumor origin, progression, and recurrence. ChlorideIntracellular Channel 1 (CLIC1), highly expressed in CSCs, isconstitutively present in the plasma membrane where it isassociated with chloride ion permeability. In vitro, CLIC1inhibition leads to a significant arrest of GB CSCs in G1 phaseof the cell cycle. Furthermore, CLIC1 knockdown impairstumor growth in vivo. Here, we demonstrate that CLIC1 mem-brane localization and function is specific for GB CSCs. Mes-enchymal stem cells (MSC) do not show CLIC1-associatedchloride permeability, and inhibition of CLIC1 protein func-tion has no influence on MSC cell-cycle progression. Investi-gation of the basic functions of GB CSCs reveals a constitutive

state of oxidative stress and cytoplasmic alkalinization com-pared with MSCs. Both intracellular oxidation and cyto-plasmic pH changes have been reported to affect CLIC1membrane functional expression. We now report that in CSCsthese three elements are temporally linked during CSC G1–Stransition. Impeding CLIC1-mediated chloride current pre-vents both intracellular ROS accumulation and pH changes.CLIC1 membrane functional impairment results in GB CSCsresetting froman allostatic tumorigenic condition to a homeo-static steady state. In contrast, inhibiting NADPH oxidase andNHE1 proton pump results in cell death of both GB CSCs andMSCs. Our results show that CLIC1 membrane protein iscrucial and specific for GB CSC proliferation, and is a prom-ising pharmacologic target for successful brain tumor thera-pies. Mol Cancer Ther; 17(11); 2451–61. �2018 AACR.

IntroductionGliomas are the major group of primary malignant brain

tumors. Among them, glioblastoma (GB, WHO grade IV) is themost aggressive and lethal (1, 2). To date, no effective therapiesfor GB have been developed. Surgical removal of GB is oftenineffective due to its fast spread and high rate of infiltration in

brain parenchyma. Chemotherapy is also problematic due to thepresence of the blood–brain barrier and the high expression ofantiapoptotic proteins anddrug efflux transporters on cancer cells.Current pharmacologic treatments for GB target highly prolifer-ating but nontumorigenic cancer bulk cells but spare the tumor-igenic cancer stem cells (CSC), which, being refractory to therapy,determine the poor prognosis of GB patients. Although usuallyquiescent, CSCs are the reservoir for a potential tumor recurrence(3, 4). Isolated CSCs, which contain all the necessary geneticinformation to faithfully reproduce the original tumor, can thenbe used as a model to look for new diagnostic markers and screenpotential drugs. In addition, therapies directed specifically againstCSCs could offer better clinical outcomes (4). The greatest chal-lenge of GB treatment is likely connected to the need to prefer-entially target CSCs.

We have previously demonstrated the pivotal role of CLIC1protein in the tumorigenic potential ofCSCs isolated fromhumanGBs, and showed that CLIC1 functional expression in cell mem-branes is essential for the proliferation and self-renewal of GBCSCs (5, 6). CLIC1 was cloned from a humanmonocytic cell line(7). First identified on nuclearmembranes, it was then also foundon plasma membranes (7, 8). This protein is highly conservedamong vertebrates, ubiquitously expressed in different tissues,and overexpressed in tumors (9, 10). CLIC1 has been character-ized as a metamorphic protein (11), coexisting as both a solublecytoplasmic form and a membrane-associated element. The tran-sition between its hydrophilic and hydrophobic localizations ispromoted by several factors. In vitro studies on an artificial lipidbilayer and isolated cells have demonstrated that changes in both

1Dipartimento di Bioscienze, Universit�a degli Studi di Milano, Milan, Italy.2Fondazione Istituto FIRC di Oncologia Molecolare, Milan, Italy and CogentechS.c.a.r.l., IFOMViaAdamello,Milan, Italy. 3Sezione di Farmacologia, Dipartimentodi Medicina Interna & Centro di Eccellenza per la Ricerca Biomedica (CEBR),Universit�a di Genova, Genova, Italy. 4IRCCS Policlinico San Martino andDipartimento delle Terapie Oncologiche Integrate, Ospedale San Martino,Genova, Italy.

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

M. Peretti and F.M. Raciti are the cofirst authors of this article.

T. Florio and M. Mazzanti contributed equally to this article.

Current address for M. Peretti: Laboratory of Cell and Molecular Physiology,EA4667, University of Picardie Jules Verne, Amiens, France.

Current address for F.M. Raciti: Department of Physiology and Biophysics,University of Miami Miller School of Medicine, 1600 NW 10th Ave, Miami, FL33136.

Corresponding Author: Michele Mazzanti, University of Milan, Via Celoria 26,20133 Milano, Italy. Phone: 39-0250314958; E-mail: [email protected]

doi: 10.1158/1535-7163.MCT-17-1223

�2018 American Association for Cancer Research.

MolecularCancerTherapeutics

www.aacrjournals.org 2451

on August 11, 2021. © 2018 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst August 22, 2018; DOI: 10.1158/1535-7163.MCT-17-1223

Page 2: MutualInfluenceofROS,pH,andCLIC1Membrane Protein in the ...Marta Peretti1, Federica Maddalena Raciti1,Valentina Carlini1, Ivan Verduci1, Sarah Sertic1, Sara Barozzi2, Massimiliano

intracellular pH (pHi) and oxidative state are responsible forconveying CLIC1 protein to the membrane (9, 12–14). Once inthe membrane, the protein is associated with increased mem-brane permeability,mainly to chloride ions. It is still controversialwhether CLIC1 is the protein that actually forms the ionic path-ways or whether it modulates an existing population of ionchannels. Supporting the first hypothesis, previous experimentson different cell types using ionic channel blockers (15), siRNA(5, 6), and point mutation (16) have demonstrated a clearlink between CLIC1 presence in the lipid bilayer and anincreased transmembrane chloride current. However, in bothcases, the activated protein functions to increase ionic membraneflux (17). CLIC1 protein downregulation and direct blockadeof CLIC1-activated membrane current (5, 6) result in CSC accu-mulation in the G1 phase and an elongation of CSC cell cyclingtime (5, 6).

Over the last two decades, researchers have defined an impor-tant role for ion channels in tumor development and growth, andthey are now novel targets for cancer therapy. Channel dysfunc-tions may have a strong effect on cell physiology and signaling,and it has become clear that abnormal triggering of ion channelactivity could support the high proliferation rate, mobility, andinvasiveness of tumor cells. Ion channels expression is oftenaltered in tumor cells. During the change from a physiologicaltoward a neoplastic state, like several other proteins, most ionicpermeabilities result overexpressed or altered in their function(9, 18, 19). The main impediment to using ionic channels astherapeutic targets is their diffuse presence in the cell mem-brane, both in physiologic and pathologic conditions. This isnot the case for CLIC1-associated ionic current (10). Thechloride membrane current in GB CSCs, as well as in severaltumor cell lines, can be downregulated by CLIC1 siRNA andblocked by IAA94, a specific inhibitor of CLIC1-induced chlo-ride current (5, 6). Conversely, there is no measurable currentin differentiated GB CSCs or umbilical cord mesenchymal stemcells (MSC), although CLIC1 protein is abundant in MSCcytoplasm (5). Furthermore, the abundance of CLIC1 proteinand its associated membrane current density is directly pro-portional to the tumor's aggressiveness (6). CLIC1 protein'scytoplasmic role, if any, is still unclear and requires moreinvestigation. As previously reported, CLIC1 translocation fromthe cytosolic aqueous phase to the membrane lipid bilayer ismodulated by pH changes and fluctuating oxidative levelswithin the internal face compartment (9, 14). In general,cellular metabolic activity rises as internal pH increases. Oscil-lations in pHi contribute to the control of cell-cycle progressionand cell division in both prokaryotic and eukaryotic cells. Arapid alkalinization of pHi may be important for G1–S phaseprogression (18). High pHi, associated with a more acidicextracellular space, is involved in the processes of neoplastictransformation and tumor development (19, 20). Oxidative-level oscillations in the intracellular compartment also con-tribute to the regulation of cell-cycle progression (21, 22).Menon and colleagues have shown that a transient increasein pro-oxidant levels early in G1 is required for the cells toenter S phase (23). This periodicity of intracellular redox staterequires a delicate balance between ROS production and itssubsequent removal by antioxidants (24, 25). Therefore, ROSoverproduction and/or the inhibition of antioxidant efficiencycould perturb the redox cycle, which in turn could lead toaberrant proliferation.

Alterations in cells' basal levels of oxidation typically contributeto many tumorigenic processes (22, 26, 27). Thus, it is notsurprising that CLIC1 channel activity, induced by oxidation, ishigher in hyperproliferating tumor cells. As we previously dem-onstrated in activatedmicroglia, cancer cells could take advantageof a feed-forwardmechanismbetweenCLIC1 channel activity andROS production by NADPH oxidase (12).

In this study, we highlight the crucial role of CLIC1 membranefunction in GBM CSCs during cell-cycle progression. Our resultsdemonstrate that CLIC1 protein in the plasma membrane of GBCSCs is instrumental in stabilizing, in a time-dependent manner,the hyperactivated metabolic state defined as "allostasis" (28).Therefore, inhibiting CLIC1 function should help reestablishthe homeostatic condition typical of quiescent cells. The unique-ness ofCLIC1 functional expression inCSCmembranesmakes it apromising potential pharmacologic target for GB treatment.

Materials and MethodsReagents and antibodies

Indanyloxyacetic acid 94 (IAA94), diphenyleneiodoniumchloride (DPI), 3-amino-6-chloro-N-(diaminomethylene)-5-(ethyl(isopropyl)amino)pyrazine-2-carboxamide (EIPA), andN-Formyl-L-methionyl-L-leucyl-L-phenylalanine (FMLP) werepurchased from Sigma Aldrich.

Antibodies used were mouse monoclonal anti-human CLIC1(Santa Cruz Biotechnology, sc 81873), rabbit monoclonal anti-cyclin D1 (Cell Signaling Technology, 92G2, 1:1,000), rabbitmonoclonal anti-p27 Kip1 (Cell Signaling Technology,D37H1, 1:1,000), anti-mouse Alexa Fluor 488 (Life Technolo-gies), and secondary antibody-HRP conjugated (Sigma-Aldrich).In some experiments, we used a custom-made antibody againstthe last 30 amino acids of the amino terminus (AB-NH2) of CLIC1(6, 8, 15).

Cell culturesWe obtained a CSC culture from a 48-year-old GB patient

(coded as GB3). We also used two other CSC preparations fromdifferent patients (GB23 and GB19) to reproduce key experi-ments. CSCs, isolated as described in previous work (29, 30),were grown in stem cell–permissive medium enriched with10 ng/mL human bFGF and 20 ng/mL human EGF (31, 32). Wevalidated the CSC properties as described (32), in particular,tumorigenicity was confirmed by orthotropic xenograft inNOD-SCID mice after Policlinico San Martino-IST InstitutionalAnimal Care and Use Committee approval. CSCs, grown as amonolayer onMatrigel, were synchronized in G1 phase of the cellcycle through 60 hours of starvation in growth factor–free medi-um. Cells were released from the G1 blockade by changing tocomplete stem cell–permissive medium. Human umbilical cordMSCs were obtained and characterized following the minimalcriteria proposed by the International Society for Cell Therapy, aspreviously described (6). ForCSCandMSC isolation,we collectedsamples after the Institutional Ethical Committee approved thepatients' written-informed consent.

Cell count and viabilityGBCSCs orMSCs were plated in a 24-well plate (2� 104/well)

and treated for 72 to 96 hours with different inhibitors. At the endof the incubation time, we collected and counted the cells usingthe Trypan Blue Dye exclusion test (Thermo Fisher Scientific)

Peretti et al.

Mol Cancer Ther; 17(11) November 2018 Molecular Cancer Therapeutics2452

on August 11, 2021. © 2018 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst August 22, 2018; DOI: 10.1158/1535-7163.MCT-17-1223

Page 3: MutualInfluenceofROS,pH,andCLIC1Membrane Protein in the ...Marta Peretti1, Federica Maddalena Raciti1,Valentina Carlini1, Ivan Verduci1, Sarah Sertic1, Sara Barozzi2, Massimiliano

and the Countess II FL Automated Cell Counter (Thermo FisherScientific).

Gene silencingshRNA sequence targeting human CLIC1 (target sequence: 50-

GATGATGAGGAGATCGAGCTC-30) was cloned using EcoRI andAgeI sites in pLKO.1-TCR cloning vector (Addgene). As controlSCRAMBLE, shRNA cloned in pLKO.1 vectors was used(Addgene). Lentiviral vectors were produced by cotransfectionof pLKO and packaging plasmids psPAX2 (Addgene) and pMD2.G (Addgene) into HEK293T. Virus was harvested at 60 hours aftertransfection, and infections were carried out to stably expressconstructs in GB CSCs.

ElectrophysiologyPatch-clamp electrophysiology was performed in a perforated-

patch whole-cell configuration as previously reported (5). Thevoltage protocol consisted of 800 ms pulses from �40 mV toþ60 mV (20 mV voltage steps). The holding potential was setaccording to the resting potential of the single cell (between40 and�80 mV); we applied a 15 ms prepulse of�40 mV beforestarting the voltage steps.

CLIC1-mediated chloride currents were isolated from the cells'other ionic currents by perfusing the specific inhibitor IAA94(100 mmol/L). Patch-clamp solutions (22) included the follow-ing: bath (mmol/L): 125 NaCl, 5.5 KCl, 24 HEPES, 1 MgCl2,0.5 CaCl2, 5D-glucose, 10NaOH, pH7.4; pipette (mmol/L): 135KCl, 10 HEPES, 10 NaCl, 1 MgCl2, pH 7.2. In the experimentsshown in Fig. 7, the bath solution pH was changed at the valuereported in the figure legend.

Protein extraction and Western blotCells (2 � 105 cells each) were collected by directly scraping

with a lysis buffer (LB; in water: 0.25 mol/L Tris-HCl pH 6.8, 4%SDS, 20% glycerol). Complete lysis of all cells was obtained by 10minutes of incubation in LB at 95�C. Concentration of proteinlysates was assessed by BCA assay (Pierce). We added LDS SampleBuffer 4X (Thermo Fisher Scientific) to samples consisting ofequal amounts of protein and loaded each sample onto a 12%SDS-PAGE under reducing conditions. Resolved proteins wereelectroblotted onto a nitrocellulose membrane (Bio-Rad Labora-tories) and probed with specific antibodies. Western blot datawere normalized on housekeeping protein (tubulin) and plotted.

ImmunofluorescenceCSC monolayers of 5 � 103 cells, seeded on 12-mm diameter

cover glass, were fixed with 2% paraformaldehyde and incubatedwith specific antibodies. To visualize the nuclei, we treated thecells with DAPI and mounted them on microscope slides using aglycerol-based mounting agent. Samples were observed under aconfocal microscope (Leica TCS SP2) with a Leica HCX PL APO63X/1.4NA oil immersion objective for Fig. 2. Images wereanalyzed using ImageJ software.

pHi measurementWe measured pHi using the probe carboxy SNARF-1 AM

(carboxy SemiNaphthoRhodaFluor acetoxymethyl ester, LifeTechnologies), calibrated with high-potassium buffers and niger-icin as described in Chow and Hedley's protocol (33). Briefly,CSCs were harvested, centrifuged for 10 minutes at 600 rpm, andresuspended in 1X PBS (1� 106 cells/mL PBS). Carboxy SNARF-1

AM was added at the final concentration of 5 mmol/L, and cellswere incubated for 30 minutes at 37�C in the dark. Samples werecentrifuged at room temperature, resuspended in 500 mL of thedesired solution, incubated for 20 minutes at room temperaturein the dark, and then analyzed by flow cytometry with a FACSAria (BD Biosciences) to determine the 640/580 fluorescenceratio. The solutions used included: samples (in mmol/L) 125NaCl, 5.5 KCl, 24 HEPES, 1 MgCl2, 0.5 CaCl2, 5 glucose, 10NaOH, pH 7.4 � desired treatment; standard high-potassiumbuffers (pH 6.8–7.0–7.2–7.4–7.6–7.8)þ nigericin (final concen-tration 2 mg/mL).

Figure 1.

Modulation of cell-cycle time course in human GB cancer stem cells.A, Time-lapse analysis of progression toward first cell division aftersynchronization release. The first time interval (late S) begins immediatelyafter cell release from G1 synchronization and runs until the morphologicchange to a round shape (n ¼ 19). The second time interval (G2) endswith complete cell division (M; n ¼ 20). Average time spent by the cell inS was 7.03 � 0.51 hours for control cells, 10.42 � 1.09 hours for IAA94-treatedcells, and 9.38 � 1.18 hours for AB-NH2–treated cells. Differences weresignificant in both treatments compared with the control (���� , P < 0.0001,two-sample t test).B,Cell-cycle phase dynamic histograms from0 time (releasefrom starvation) to 28 hours for GB stem cells in control conditions andincubated with IAA94. C, The plot on the left shows 96-hour GB stem cellsgrowth curves in control condition (squares), in the presence of IAA94(triangles) and incubated with the AB-NH2 (inverted triangles). The plot onthe right depicts growth curves of control CSCs and cells infected withCLIC1 shRNA (circles).

CLIC1 Activity Controls Glioblastoma Stem Cell Proliferation

www.aacrjournals.org Mol Cancer Ther; 17(11) November 2018 2453

on August 11, 2021. © 2018 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst August 22, 2018; DOI: 10.1158/1535-7163.MCT-17-1223

Page 4: MutualInfluenceofROS,pH,andCLIC1Membrane Protein in the ...Marta Peretti1, Federica Maddalena Raciti1,Valentina Carlini1, Ivan Verduci1, Sarah Sertic1, Sara Barozzi2, Massimiliano

ROS measurementsTo measure the redox conditions of the cells, we performed

experiments using 20,70- dichlorofluorescin diacetate (DCFH-DA,Sigma). CSCs were plated as a monolayer and incubated withDCFHDA, which passively diffuses into cells as a nonfluorescentprobe. DCFH-DA is de-esterified by intracellular esterase andbecomes a highly fluorescent adduct that is trapped inside thecell upon oxidation, which allows for rapid quantification ofoxygen-reactive species in response to oxidative metabolism.

After 24 hours of treatment, cells were incubated with Hank'sBuffered Salt Solution (HBSS) with added DCFH-DA (final con-centration 20mmol/L) for 30minutes at 37� in the dark. For short-term treatments, the compounds were added right before mea-surements in fresh HBSS (0.137 mol/L NaCl, 5.4 mmol/L KCl,0.25 mmol/L Na2HPO4, 0.1 g/L glucose, 0.44 mmol/L KH2PO4,1.3 mmol/L CaCl2, 1.0 mmol/L MgSO4, and 4.2 mmol/LNaHCO3).

Cells were imaged in vitro using a Nikon Ti-E inverted fluores-cence microscope (Nikon) with a CFI Plan Apo VC 20X objective.Excitation light was produced by a Prior Lumen 200 PROfluorescent lamp (Prior Scientific) at 470/40 nm. Images werecollected with a Hamamatsu Dual CCD Camera ORCA-D2(Hamamatsu Photonics), and emissions were gathered with a505–530 nm bandpass filter. Exposure time was 100 ms with nobinning for the excitation wavelength, and images were acquiredevery minute for 31 minutes. Filters and a dichroic mirror werepurchased from Chroma (Chroma Technology). The NISElement(Nikon) was used as platform to control the microscope, illumi-nator, camera, and postacquisition analyses.

It is important to note that the data reported in this article arenot an absolute value of the cells cytoplasmic oxidation state.They are rather a measurement of the ROS production/ROSelimination slope detected at any given time and consequentlythey give an idea of the general oxidative level of the cells.

Cell-cycle analysis by flow cytometryCells were washed in PBS, fixed with ice-cold 95% ethanol

at 4�C for 1 hour, and resuspended in staining solution (PBS,

20 mg/mL RNAse A, 50 mg/mL propidium iodide, 0.5% TritonX-100, all from Sigma-Aldrich). DNA content was quantified byFACScalibur (BD Bioscience) as reported (34). Cell-cycle profilewas determined by analysis of Listmode data using ModFit LTsoftware (Verity Software House). At least 10,000 events werecollected gating single nuclei and excluding aggregates.

Time-lapse microscopyFor proliferation experiments, we plated CSCs in a 6-well plate

(3 � 105 cells/well) as monolayer. Time lapses were performedusing a ScanR system (Olympus) based on either an IX81 invertedmicroscope equipped with a Hamamatsu ORCA-Flash4 cameraand driven by CellSens software (Olympus), or an EclipseTE200-E microscope (Nikon) equipped with a Cascade II-512camera (Photometrics) driven by Metamorph software.

During the entire observation process, cells were kept in amicroscope stage incubator (Okolab) for environmental control.We analyzed the cells using a 20x LUCPlanFLN (NA0.45) or a 10XPlanFluor (NA 0.30) objective. Images were acquired every 15minutes for 36 hours and were analyzed using Fiji software (35).

TIRF measurementsWe used TIRF microscopy in different experiments to visualize

the spatial-temporal dynamics of CLIC1 near the cell surface.Using a Leica AM TIRF microscope and an Andor iXon DU-885camera, we obtained time-lapse video with images taken every 15minutes for 3 or 7 hours with a HCX PL APO 63X/1.47NAobjective. CSCs were transfected with a plasmid carrying theCLIC1-GFP sequence, seeded in adhesion on circular cover glasses(�¼ 25mm), and then synchronized inG1 phase. After 36 hours,cells were released in complete medium or in external solution atpH ¼ 8 (see Bath Solution). Emitted fluorescence was recordedwith a Filtercube GFP-T ET TIRF MC filter for laser line 488(ex-dic-em: 475/40-500-530/50).

Imaging analysisIn ROS and TIRF experiments, we determined fluorescence

intensities for every image over regions of interest corresponding

Figure 2.

CLIC1 inhibition delays cyclin D1 risingtime. A and B, RepresentativeWestern blot analysis of p27Kip1 andcyclin D1 proteins in CSCs at differenttime points after release from G1

synchronization. B, Time course ofdensitometric data (n ¼ 3) showingp27Kip1 reduction (open symbols) andcyclin D1 increase (filled symbols)showing that the inversion occurredbetween 6 and 8 hours after therelease. C, Western blot of cyclin D1between 0 and 12 hours in controlconditions and in the presence ofIAA94 in CSCs after release fromsynchronization. D, Time course ofdensitometric data (n ¼ 2) showingcyclin D1 increment in control GBCSCs (open triangles) and in cancerstem cells incubated with IAA94(open squares).

Peretti et al.

Mol Cancer Ther; 17(11) November 2018 Molecular Cancer Therapeutics2454

on August 11, 2021. © 2018 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst August 22, 2018; DOI: 10.1158/1535-7163.MCT-17-1223

Page 5: MutualInfluenceofROS,pH,andCLIC1Membrane Protein in the ...Marta Peretti1, Federica Maddalena Raciti1,Valentina Carlini1, Ivan Verduci1, Sarah Sertic1, Sara Barozzi2, Massimiliano

to the cells analyzed by subtracting the specific background andnormalizing the values to the first one detected for each cell. Allanalyses were performed using Fiji software (35).

Statistical analysisWe repeated all experiments at least twice; quantitative data

were collected from experiments performed in triplicate or qua-druplicate and expressed as mean � SEM, except for time-lapseexperiments (mean � SD). CLIC1-mediated current density/voltage relationships were analyzed as following: for each con-dition, we plotted every experiment and extrapolated the linearregression for the curve (all significant, with R2 � 0.9). Foranalyses of two conditions,weused a two-sample t test; otherwise,we performed one-way ANOVA analysis on the slopes of everyexperimental group.

In the box chart plots, the solid lines within the boxes representthemedian values. The squarewithin the boxes is the average, andthe boxes show the 25th and 75th percentile range of the mea-sured elements. Themaximum andminimumvalues are depictedas horizontal bars, and the symbols to the right of the box are thenumbers of trials.

ResultsCLIC1 controls G1–S phase transition

After 60 hours of starvation, human GB CSCs were syn-chronized and accumulated (more than 90%) in G1 phase ofthe cell cycle. Synchronized cells released in complete medi-um progressed to the S phase (5), and the small percentage ofcells (�10%) already in late S and G2 phase underwent theirfirst division after 8 to 12 hours. CLIC1 contribution to cell-cycle progression is already visible for these cells. GB CSCsincubated with IAA94, a CLIC1 ionic conductance inhibitor,displayed a longer transition time before the G2–Mphase. Figure 1A shows the results of a time-lapse observationof CSC cell-cycle dynamics confirming the effect of CLIC1-membrane protein inhibitor. The same results were obtainedby incubating CSCs with a custom-made antibody directedagainst the extracellular portion of the protein, which isknown to inhibit CLIC1 function at the membrane level(6, 8, 15).

According with cell-cycle progression measurements, CLIC1inhibition has a most pronounced effect on G1–S transitiontime. Flow cytometry analysis of DNA content distributionshows that CLIC1 inhibition by IAA94 treatment delays thetransition of GB CSCs from G1 phase (Fig. 1B). In fact, cellstreated with IAA94 required about 8 hours more to reachsimilar percentage of cells in S phase as shown in Fig. 1B andSupplementary Fig. S1.

Also analysis of GB CSC growth curve reflected the elongationof the G1–S phase in condition of CLIC1 inhibition. In IAA94 aswell as in CLIC1-silenced cells, CSCs' proliferation is stronglydownregulated (Fig. 1C). In particular, although shRNA transfec-tion does not completely eliminate the protein (SupplementaryFig. S2), cell growth is drastically compromised.

Western blot evaluation of cyclin D1 and p27 expression every2 hours after G1-synchronization release established the startingtime of G1–S transition in control cells (Fig. 2A and B). IAA94treatment of released cells shifted the time of cyclin D1 increasebetween 2 and 4 hours (Fig. 2C and D). The plot in Fig. 2Dsupports the results obtained by FACS analysis.

Membrane localization of CLIC1 in S phaseImmunolocalization of CLIC1 with a commercial antibody

confirms its metamorphic nature. At the end of starvation period,CLIC1 protein appeared diffuse in the cytoplasm. After 8 hours in

Figure 3.

CLIC1 protein localization and function. A, Confocal microscope images ofindirect immunolocalization of CLIC1 in GB CSCs using a whole-proteinantibody at 0 and 8 hours from the release from G1 arrest. After 8 hours,there was a marked localization of the protein at membrane level. B, FACSanalysis of CSCs after staining cell surfaceswith the primary antibody specific forCLIC1 NH2 terminus (AB- NH2) at 0 and 8 hours after release fromsynchronization. Antibody specificity was confirmed by saturating AB-NH2 withthe same peptide used to immunize the mice. Eight hours after the release(151.2 � 3.20, n ¼ 5), there was a higher fluorescence signal compared with0 hours (119.4 � 6.16, n ¼ 6), showing an increase of the protein in themembrane (one-way ANOVA; �� , P < 0.01, Tukey test). C, IAA94-sensitivecurrent density/voltage relationships of CSCs at different times from the releasefrom G1 synchronization (0 h circles, n ¼ 5; 8 h triangles, n ¼ 7; 12 h diamonds,n ¼ 7) and randomly cycling (squares, n ¼ 8). Statistics analysis: one-wayANOVA, followed by Tukey test (��� , P < 0.001 vs. randomly cycling cells, 0 and12 hours). D, CSCs transiently transfected with CLIC1-GFP plasmid wereprocessed with TIRF time-lapse analysis from 0 to 7 hours (n ¼ 5) and from 7to 12 hours (n¼ 5) after release in completemedium fromG1 arrest. A significantincrease of the fluorescence occurred between 0 and 4 hours, and wasmaintained up to 8 hours. After this time, fluorescence signal declined,reaching basal level after 12 hours. Examples of fluorescence images at 0,4, and for a different cell, at 8 and 12 hours from G1 release are shown inthe panels on the top of the plots.

CLIC1 Activity Controls Glioblastoma Stem Cell Proliferation

www.aacrjournals.org Mol Cancer Ther; 17(11) November 2018 2455

on August 11, 2021. © 2018 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst August 22, 2018; DOI: 10.1158/1535-7163.MCT-17-1223

Page 6: MutualInfluenceofROS,pH,andCLIC1Membrane Protein in the ...Marta Peretti1, Federica Maddalena Raciti1,Valentina Carlini1, Ivan Verduci1, Sarah Sertic1, Sara Barozzi2, Massimiliano

complete medium, when the cell cycle had progressed to S phase,the protein localized close to the plasma membrane and concen-trated around the nucleus in several cells (Fig. 3A). We quantifiedCLIC1 presence in the membrane in three ways: (i) cytofluori-metric analysis, (ii) perforated patch-clamp configuration experi-ments, and (iii) TIRF measurements. For the first method, wecollected CSCs at 0 and 8 hours after synchronization release andincubated live cells with a custom-made antibody (AB-NH2)directed against the external portion of CLIC1 transmembraneprotein (8). The higher fluorescence signal detected at 8 hours wasproportional to the translocation of the protein from thecytoplasm to the membrane. We confirmed the specificity of thedetected signal by competition with the peptide used for immu-nization (Fig. 3B). Furthermore, whole-cell electrophysiologicalmeasurement of IAA94-sensitive membrane current not onlysupported the idea of a protein-rich membrane, but alsoconfirmed the increase of CLIC1-associated chloride current dur-ing G1 phase progression (Fig. 3C). The presence of other IAA94-sensitive member of the CLIC family (namely CLIC5A, ref. 36) inGB CSCs was excluded by qRT-PCR and thus confirms the spec-ificity of above results (Supplementary Fig. S3). The slope of thecurrent/voltage relationships shown in Fig. 3C represents mem-brane conductance. At 8 hours after synchronization release(138 pS, triangles), it is almost 5 times higher compared withcells measured just after release (32 pS, circles) or 12 hours later(35.3 pS, diamonds), and to randomly cycling CSCs (23.5 pS,squares).

Finally, we monitored CLIC1 membrane-localization dynam-ics using the TIRF procedure. We observed CSCs transientlytransfected with CLIC1-GFP plasmid, looking for surface fluores-cence in two time windows: 0–7 hours and 7–12 hours aftersynchronization release (Fig. 3D). Membrane fluorescence sud-denly increased and reached themaximumvalue 4 hours after cellrelease. The fluorescence level held steady up to 8 hours, when itbegan a slow decline, returning to basal signal after 12 hours.

Surface CLIC1 inhibition reduces ROS production and pHiIt has been reported that metabolic conditions are chronically

altered in GB cancer cells compared with normal stem cells. Thecytoplasmic redox ratio and pHi levels are not exceptions. Inparticular, because a transient increase in reactive oxygen species(ROS) is essential for cell-cycle progression from G1 to S phase(21), we would expect GB CSCs to have a greater ability togenerate ROS. Similar considerations can be applied to pH levels:previousmeasurements showed higher pHi values in GB cell linescompared with nontumor cells (20). ROS ratio between produc-tion and antioxidant action monitored for 30 minutes in ran-domly cycling CSCs is completely abolished in the presence ofIAA94 (Fig. 4A). Similar results have been obtained with AB-NH2.Both treatments reduced ROS content in GB CSCs to the level ofMSCs. pHi is also perturbed by CLIC1 inhibition. Randomlycycling GB CSCs display an average pHi of 7.46. Impairing

Figure 4.

Effect of reciprocal inhibition of ROS production, pH changes, and CLIC1current. A, ROS dynamic production measurements in GB CSCs and MSCsover 30 minutes of continuous recording. Chronic treatments with 100 mmol/LIAA94 (n ¼ 26, circles) and with AB-NH2 (n ¼ 31, diamonds) wereperformed and compared with the measurements obtained in randomlycycling GB CSCs (n ¼ 30, squares) and MSCs (n ¼ 9, triangles). Slope valuesbetween CSCs and each of treated cells and MSCs were analyzed (alwaysP < 0.0001, t test). B, Box chart plot of pHi measurements of randomlycycling untreated CSCs, and CSCs treated with 100 mmol/L IAA94, AB-NH2

antibody, and randomly cycling untreated MSCs. The mean of 11 measurementswas 7.46 � 0.02 for CSCs. A 24-hour treatment with IAA94 and AB-NH2

significantly acidified cell cytoplasm to pHi 7.24 � 0.01 (n ¼ 3, t test ��� , P <0.001) and 7.26 � 0.03 (n ¼ 3, t test ��� , P < 0.001), respectively. Mean pHi

measured in MSCs was 7.25 � 0.07 (n ¼ 5, t test ��� , P < 0.001). C, Boxchart plot of CSC CLIC1 current density at þ60 mV step released from G1 arrestin the presence of proton-pump and membrane-oxidase inhibitors. Wecompared CLIC1-mediated chloride current density between randomlycycling CSCs (2.35 � 0.51 pA/pF; n ¼ 9), at time 0 after release fromsynchronization (3.6� 0.4 pA/pF; n¼ 3) and after 8 hours (14.32� 0.87 pA/pF;n ¼ 9). (Continued on the following column.)

(Continued.) The presence of EIPA 25 mmol/L (2.53� 0.25 pA/pF; n¼ 5) or DPI10 mmol/L (3.99 � 0.57 pA/pF; n ¼ 5) drastically reduced the membrane ionicflow. The same effect was observed in the presence of AB-NH2 antibodyin the external solution (2.35 � 0.33 pA/pF; n ¼ 7). The average current ofrandomly cycling MSCs was 0.95� 0.15 pA/pF (n ¼ 5; one-way ANOVA,���� , P < 0.0001, Dunnett multiple comparison test). Current/voltagerelationships at 4 and 8 hours after synchronization release in the presenceof either DPI or EIPA are shown in Supplementary Figs. S4 and S5, respectively.

Peretti et al.

Mol Cancer Ther; 17(11) November 2018 Molecular Cancer Therapeutics2456

on August 11, 2021. © 2018 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst August 22, 2018; DOI: 10.1158/1535-7163.MCT-17-1223

Page 7: MutualInfluenceofROS,pH,andCLIC1Membrane Protein in the ...Marta Peretti1, Federica Maddalena Raciti1,Valentina Carlini1, Ivan Verduci1, Sarah Sertic1, Sara Barozzi2, Massimiliano

CLIC1-associated membrane current by treatment with IAA94or AB-NH2 for 24 hours caused pronounced cytoplasmic acidi-fication, reaching values proximal to MSC pHi (Fig. 4B). The linkbetween CLIC1-associated membrane current and NADPHoxidase (12) appears to be shared also by the NHE1 protonpump, one of the membrane transporters responsible for thecontrol of pHi (20). Both elements are upregulated in GB cellmembranes (20, 37, 38) and are specifically inhibited byDPI (38)or EIPA (39), respectively. CSC membrane current recordingshows that CLIC1 current density is inhibited at 8 hours aftersynchronization release by AB-NH2, DPI, and EIPA (Fig. 4C),confirming the functional link between the three membraneproteins (Supplementary Figs. S4 and S5).

CLIC1 membrane protein is specific for GB CSCsThe results showing a tight interconnection between NADPH

oxidase, NHE1 proton pump, and CLIC1-associated transmem-brane chloride current suggest that CSC proliferation could bereduced by downregulating one of these three membrane pro-teins. Incubating CSCs for 72 hours with IAA94, AB-NH2, DPI, orEIPA resulted in a drastic reduction of proliferation (Fig. 5A).However, inMSCs the reduction of cell viability induced by IAA94and AB-NH2 was negligible compared with the marked effect ofDPI and EIPA (Fig. 5B). The inserts in Fig. 5A and B depict singleexamples of complete time course evaluating cell growth in GBCSCs andMSCs evidencing that only in CSCs AB-NH2was able toaffect proliferation. The reduction in the number of both CSCs

and MSCs treated with membrane oxidase (DPI) and protonpump (EIPA) inhibitors was mainly due to cell death, whereasIAA andAB-NH2 reduce cell number by reducing the proliferationrate of GB CSC (Fig. 5C and D).

Cross-modulation of pHi and ROS by NADPH oxidase andNHE1 inhibitors

As we previously demonstrated, modifying CLIC1 membranefunction affects ROS and pHi dynamics. NADPH oxidase andNHE1 are also mutually dependent. Figure 6 depicts the changesthat occur in ROS ratio (i) and pH modification (ii) when one ofthe two events is compromised. As a comparison, the plots alsoshow the effects of acute and prolonged CLIC1 current inhibitionon ROS content and pH. All the average data reported wereexpected, except for the early effect on pH caused by IAA94.Short-term CLIC1 inhibition not only reduces ROS unbalancebut also exacerbates cytoplasmic alkalinization. This result sug-gests that the link between CLIC1 and NHE1 is not direct but islikely mediated by NADPH oxidase function. Timing disjunctionis evident in a parallel study of the kinetic parameters after CSCrelease from starvation. Figure 7 shows the ROS dynamics (i),CLIC1 current density (ii), and cytoplasmic pH evaluation (iii)from cell release through the G1–S transition (see Fig. 1). ROSaccumulation and membrane ionic mechanism were previouslycharacterized in microglia cells (12). However, although NADPHoxidase activity shows a peak around 3 to 4 hours, CLIC1 currentreaches a maximum after 8 hours, synchronous with the

Figure 5.

CLIC1 membrane protein function isspecific for CSC proliferation. A andB, GB CSCs or MSCs were treated for72 hours with IAA94 (100 mmol/L),AB-NH2 antibody (2.5 mg/mL), DPI(10 mmol/L), and EIPA (25 mmol/L).CLIC1 inhibitors reduced cell numbersin CSCs but not in MSCs. NADPHoxidase and NHE1 blockers drasticallydecreased the number of both GB andmesenchymal cells. The insets showsingle examples of complete GB CSCand MSC growth curves from 0 to96 hours in the control condition(squares) and in the presence ofAB-NH2 antibody (triangles). C andD, Dead/live cell ratio calculated fromthe data depicted in A. After 72 hours,the majority of the cells treated withCLIC1 inhibitors were alive. On thecontrary, ROS production or protonpump impairment resulted in morethan 80% trypan blue positive inboth cell types.

CLIC1 Activity Controls Glioblastoma Stem Cell Proliferation

www.aacrjournals.org Mol Cancer Ther; 17(11) November 2018 2457

on August 11, 2021. © 2018 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst August 22, 2018; DOI: 10.1158/1535-7163.MCT-17-1223

Page 8: MutualInfluenceofROS,pH,andCLIC1Membrane Protein in the ...Marta Peretti1, Federica Maddalena Raciti1,Valentina Carlini1, Ivan Verduci1, Sarah Sertic1, Sara Barozzi2, Massimiliano

maximum level of cytoplasmic alkalinization. As previouslyreported (21), a temporary oxidation increase can be interpretedas a permissive signal to cell-cycle progression. Instead, thetransient alkalinization of the cytoplasm occurs toward the endof theG1–S transition, which suggests several potential functionalinterpretations. One hypothesis is that pH change works as an offsignal, downregulating the G1–S transition machinery. CLIC1-associated current drops synchronously with the maximum cyto-plasmic pH value. The ability to mediate the interaction between

CLIC1 and the biological membrane by pH change is well-documented in previous publications (14, 37, 40, 41). In perfo-rated whole-cell experiments, external pH was systematicallydropped in the external solution from 7.4 to 7.0 and to 6.5 inrandomly cycling CSCs. Electrophysiologic experiments measur-ing CLIC1 current demonstrated that sudden acidification ofextracellular pH inhibits CLIC1 current (Fig. 8A). Figure 8B showsthe average current inhibition relative to IAA94 action for externalpH 7 and 6.5. Delivering the specific inorganic blocker before pHacidification prevents any further decrease in membrane current,which confirms that pH changes and IAA94 act on the same

Figure 6.

Mutual modulation of ROS, pH, and CLIC1 ionic current. A, Box chart plot ofROS production comparing CSCs treated with CLIC1-associated current blockerand proton pump inhibitor with randomly cycling CSCs and MSC (dotted lines).The data reported are linear fittings of slope curves obtained by time courseexperiments as shown in Fig. 3A. Short (15 minutes) and long (24 hours)treatments were performed with both compounds. Acute stimulation withEIPA 10 mmol/L (n ¼ 19, down triangles, not significant) had no effect onROS production. Conversely, 15 minutes of incubation with IAA94 100 mmol/L(n¼ 24, circles, ���� ,P<0.0001)was sufficient todownregulateROSproduction.Chronic exposure to EIPA 10 mmol/L (n¼ 33, circles, ���� , P < 0.0001) or IAA94100 mmol/L (n¼ 25, up triangles, ���� , P < 0.0001) decreased ROS production toMSC level. All statistical analysis was done with one-way ANOVA, Dunnettmultiple comparison test. B, Box chart of pHi measurement in CSCs treated withNADPH oxidase and CLIC1 current inhibitors compared with randomly cyclingCSCs and MSCs (dotted lines, values from Fig. 3B). DPI (10 mmol/L) led tosignificant acidification both in acute (15minutes; n¼ 4, 7.28�0.04, �� , P <0.01)and chronic (24 hours; n ¼ 4, 7.33 � 0.06, �, P < 0.05) treatments. IAA94(100 mmol/L) had divergent effects: short exposure caused a strongalkalinization (n ¼ 3, 7.69 � 0.01, ��� , P < 0.001), whereas chronic treatmentresulted in decreased pHi (n ¼ 3, 7.24 � 0.01, ��� , P < 0.001) similar to thatrecorded in MSCs. All statistical analysis was done with t test vs. CSCcontrol conditions.

Figure 7.

Time course of ROS production, CLIC1 ionic current, and pHi during G1–Stransition.A, The top traces of the graph showROSproductionmeasurements inGB CSCs during cell-cycle progression. After synchronization, the cells werereleased in complete medium with (circles) or without IAA94 100 mmol/L(squares). Three hours after release, controls show a peak (n ¼ 32) notdetectable in treated cells (n ¼ 10, t test � , P < 0.05). Each value representsthe average slope of the fitted curves (for each time point) obtained bytime-course experiments as shown in Fig. 3A. Thus, the peak signifies that therate of ROS production at 3 hours after release in the control condition issignificantly higher than in treated cells, and corresponds to the increase inCLIC1-mediated current density from 4 to 8 hours after release from G1synchronization (B). The peak observed in CLIC1-mediated current starts todecrease simultaneously with the peak of alkalinization measured between8 and 10 hours after release (C). As in ROS production, cells were synchronizedand released, and measurements were performed for each time point.

Peretti et al.

Mol Cancer Ther; 17(11) November 2018 Molecular Cancer Therapeutics2458

on August 11, 2021. © 2018 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst August 22, 2018; DOI: 10.1158/1535-7163.MCT-17-1223

Page 9: MutualInfluenceofROS,pH,andCLIC1Membrane Protein in the ...Marta Peretti1, Federica Maddalena Raciti1,Valentina Carlini1, Ivan Verduci1, Sarah Sertic1, Sara Barozzi2, Massimiliano

membrane target. Measurements of current inhibition lag timeafter perfusing the test solution (Fig. 8C) suggest that acidificationdoes not work as an ion channel blocker. Removing CLIC1protein from themembrane could be an alternative way to reducemembrane current. As already shown in Fig. 3 and confirmedin Fig. 8D (empty circles), TIRF experiments demonstrate that dueto the presence of CLIC1-GFP at the membrane level of GB CSCs,the fluorescent signal progressively fades 8 hours after releasefrom G1 accumulation. This occurs synchronously with the intra-cellular alkalinization peak (Fig. 7C) and with CLIC1-associatedcurrent decrease (Fig. 7B). TIRF measurements obtained whilemaintaining the external solution at pH 8 (open square) show nofluorescent signal dimming (Fig. 8D).

DiscussionHuman GB CSC-enriched cultures show a steady-state condi-

tion in which both ROS balance and pHi are chronically altered.Randomly cycling GB CSCs have a 5-fold increase in their abilityto produce ROS and a pHi almost 0.3 units more alkaline thanhuman MSCs (Fig. 4A and B). Many different solid tumorsdemonstrate similar, consistent overexpression and upregulationof several metabolic cell components (42–45). Physiologically,

this cellular hyperactivation is a transient condition. The processof achieving a new state of stability throughphysiologic changes isknown as "allostasis" (28). This phenomenon is essential tomaintain cell viability during fluctuating basal conditions. How-ever, when stress is persistent and cells are unable to restore a basalhomeostatic state, allostasis can become chronic. Most of the cellsare not able to cope with these new conditions prolonged in timeand they die.Maintaining allostasis is energetically very costly andrequires cells with altered basal metabolism. As an example, achronic condition of oxidative stress results in an overproductionof antioxidant component to balance the system. This is the casefor GB CSCs, in which the most evident allostatic outcome is ahigh rate of cell division (5, 6). Several proteins present in theallostatic GB cell plasma membrane are deregulated in eitherexpression or function (20, 46–48). Altering any of these proteinscan disrupt the allostatic equilibrium, causing drastic functionalchanges and, in some cases, even cell death. Theoretically,manyofthese proteins can be considered as valuable pharmacologictargets. However, most of them also play important roles inphysiologic homeostatic conditions. Consequently, modifyingtheir functions would affect not only cancer cells but also healthycell populations. On the contrary, CLIC1 protein only becomescontinuously active as a membrane charge carrier during periodsof chronic stress, whereas during homeostatic conditions, it isessentially irrelevant. This is the most important result of ourinvestigation. Figure 5 shows that the inhibition of NADPHoxidase or the NHE1 proton pump, both overexpressed in CSCs,causes cell death in CSCs and healthy MSCs. On the contrary,impairing CLIC1 membrane protein, using either the chloridechannel blocker IAA94 or the specific antibody against the exter-nal portion of the protein, prolongs CSC cell-cycle duration andleaves MSC functions unaltered. Confirming this result, GB CSCgrowth curves in the presence of IAA94 show that the earlyapoptosis rate is less than 10% (Supplementary Fig. S6) eventhough the proliferation rate drops drastically (Fig. 4A). Thepresent work suggests that CLIC1 protein and its associatedchloride permeability are principal elements involved in thestabilization of GB CSC allostasis. Consistent with previousresults, CLIC1 membrane current and NADPH oxidase are partof a feed-forward mechanism that sustains ROS production, evenin GB CSCs (12). This result suggests that GB CSCs are in aconstant state of oxidative stress. ROS accumulation is consideredto be a signal to promote cell-cycle progression fromG1 to S phase(23). Our working hypothesis is that CLIC1 acts as cell-cycleaccelerator in GB CSCs.

As a metamorphic protein, CLIC1 membrane localization isstimulated by increased ROS generation. Simultaneously, bypromoting transmembrane chloride flux, CLIC1 compensates fornegative charges released by NADPH oxidase, creating a self-sustainingmetabolic pathway (12). Our experiments suggest thatto trigger G1–S transition, ROS levels must reach a definedthreshold. We can hypothesize that this increased ROS level canbe due either to an increase in the production or a decrease in thescavenging or even both the phenomena at the same time. Thefeed-forward mechanism generated by the cooperative activity ofthe membrane oxidase and CLIC1 sets the limiting rate of cell-cycle progression kinetics based on the active number of the twoelements: the more proteins involved, the faster the transition.Cytoplasm alkalinization is another characteristic of GB CSCs.Our experimental setting highlights themutualmodulation of thethree elements: ROS, pHi and chloride flux. In randomly cycling

Figure 8.

pH regulation of CLIC1 membrane activity. A, Membrane current time coursein CSCs at different external pH. Every 15 seconds, 100 mV voltage steps(from �40 to þ60 mV) for 800 ms were delivered to a whole-cell membrane.Whole-cell perforated-patch current was continuously recorded in thecontrol condition (pH 7.4) after cells were perfused in bath solution with pH 7.0and 6.5 for 5minutes each. Then, total CLIC1-associated current was inhibited byperfusing IAA94 100 mmol/L. B, Box chart plot of the current amplitudedistribution at external pH 7.0 and 6.5 as a rate of IAA94-sensitive current.C, Comparison of current inhibition lag time measured from solution flow startand the beginning of current decrease for external pH 7.0 and IAA94 100mmol/Lperfusion. D, TIRF time-lapse analysis of the presence of CLIC1 on the celladhesion surface between 6 and 9 hours after release from G1 arrest. In thecontrol condition (external pH 7.4), F/F0 values were 0.70 � 0.15 and0.49� 0.08 (n¼ 5) at 8 and 9 hours, respectively. At external pH¼ 8, the F/F0values were 1.197 � 0.07 and 1.31 � 0.13 (n ¼ 4) at 8 and 9 hours,respectively (t test � , P < 0.05 at 8 hours and �� , P < 0.01 at 9 hours aftersynchronization release).

CLIC1 Activity Controls Glioblastoma Stem Cell Proliferation

www.aacrjournals.org Mol Cancer Ther; 17(11) November 2018 2459

on August 11, 2021. © 2018 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst August 22, 2018; DOI: 10.1158/1535-7163.MCT-17-1223

Page 10: MutualInfluenceofROS,pH,andCLIC1Membrane Protein in the ...Marta Peretti1, Federica Maddalena Raciti1,Valentina Carlini1, Ivan Verduci1, Sarah Sertic1, Sara Barozzi2, Massimiliano

CSCs, pHi is chronically elevated by almost 0.3 units and isfunctionally linked with both NADPH oxidase and CLIC1 activ-ities (Fig. 5). Blocking NADPH oxidase or inhibiting CLIC1membrane function for prolonged periods results in cytoplasmicacidification (Fig. 5B). However, within minutes of suppressingCLIC1 current, we observed an unexpected transient intracellularalkalinization (Fig. 5B). This suggests the proton pump still runsafter chloride inflow stops, lowering internal [Hþ]. Thus, no directlink appears to exist between internal pH and CLIC1-associatedchloride current. However, after 24 hours of CLIC1 inhibition,pHi acidifies to MSC levels. The link is represented by NADPHoxidase: CLIC1 inhibition impairs ROS production, causingdownregulation of NHE1 function.

The role of the oxidative stress peak is well-established (21).Our results support ROS accumulation as a trigger for G1–Sprogression. However, it is not clear whether cytoplasmic pHmodulation has a role during the cell cycle. Time-lapse experi-ments on synchronized cells show ROS production growing earlyafter G1 release in parallel with increasing membrane current(Fig. 6A and B). pHi depicts a cytosolic alkalinization peakdelayed in the G1–S time progression (Fig. 6C). All our experi-ments, from immunolocalization, FACS analysis, and electro-physiology (Fig. 2) to TIRF (Figs. 2 and 7), show a gradualreduction of both CLIC1 membrane localization and its associ-ated current synchronous with transient intracellular alkaliniza-tion (Fig. 6B andC).Wehavepreviously demonstrated thatCLIC1chloride current sustains ROS production by a feed-forwardmechanism (12). Based on our electrophysiology and dynamiclocalization data in the presence of pH changes (Fig. 7), wepropose that transient alkalinization can reduce CLIC1 activityand, as a consequence, ROS production, by determining de factothe beginning of a new cell-cycle phase.

In conclusion, GB CSCs survive and proliferate in a "stable"condition in which all cell systems governing cell-cycle progres-sion (including NADPH oxidase, NHE1 proton pump, CLIC1,and others; refs. 20, 46, 48, 49) are upregulated. CLIC1 ability tobalance CSC allostatic state means that CLIC1 membrane per-meability can be considered an allostatic state stabilizer. Itsfunction is negligible in homeostasis but fundamental duringchronic stress. Among the three elements we considered in these

experiments, CLIC1 is the only target that can be proposed forpotential pharmacologic therapy with minimal side effects. Inhi-biting CLIC1-associated membrane current would not only com-promise proliferation, but would also return CSCs from chronicallostasis to steady homeostasis. Furthermore, in light of drugrepositioning approaches for cancer research (50), CLIC1 directinhibition is one of the proposedmechanisms for the antitumoraleffects of the antidiabetic drug metformin (5), which paves theway for the development of a novel class of inhibitors.

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

Authors' ContributionsConception and design: T. Florio, M. MazzantiDevelopment of methodology: M. Peretti, F.M. Raciti, V. Carlini, S. Barozzi,A. PattarozziAcquisition of data (provided animals, acquired and managed patients,provided facilities, etc.): M. Peretti, F.M. Raciti, V. Carlini, S. Sertic, S. Barozzi,M. Garr�e, A. Pattarozzi, A. Daga, A. CostaAnalysis and interpretation of data (e.g., statistical analysis, biostatistics,computational analysis): M. Peretti, F.M. Raciti, V. Carlini, I Verduci,A. Pattarozzi, F. Barbieri, T. Florio, M. MazzantiWriting, review, and/or revision of the manuscript: M. Peretti, F.M. Raciti,V. Carlini, I Verduci, A. Daga, A. Costa, T. Florio, M. MazzantiAdministrative, technical, or material support (i.e., reporting or organizingdata, constructing databases): I VerduciStudy supervision: T. Florio, M. Mazzanti

AcknowledgmentsWe are thankful to Francesca Cianci and Xueting Bai for assistance in several

molecular biology experiments, Francesca Casagrande and Ilaria Costa forconfocal imaging and time-lapsemicroscopy, andErika Pizzi for critical reading.This work was supported by grant no.16713, IG 2015 to M. Mazzanti from theItalian Association for Cancer Research (AIRC), Italy.

The costs of publication of this article were defrayed in part by thepayment of page charges. This article must therefore be hereby markedadvertisement in accordance with 18 U.S.C. Section 1734 solely to indicatethis fact.

Received December 8, 2017; revised June 6, 2018; accepted August 17, 2018;published first August 22, 2018.

References1. Schwartzbaum JA, Fisher JL, Aldape KD, Wrensch M. Epidemiology and

molecular pathology of glioma. Nat Clin Pract Neurol 2006;2:494–503;quiz 491 p following 516.

2. StuppR, BradaM, vandenBentMJ, Tonn JC, PentheroudakisG.High-gradeglioma: ESMO Clinical Practice Guidelines for diagnosis, treatment andfollow-up. Ann Oncol 2014;25:iii93–101.

3. Lathia JD, Mack SC, Mulkearns-Hubert EE, Valentim CL, Rich JN. Cancerstem cells in glioblastoma. Genes Dev 2015;29:1203–17.

4. Vescovi AL, Galli R, Reynolds BA. Brain tumour stem cells. Nat Rev Cancer2006;6:425–36.

5. GrittiM,Wurth R, AngeliniM, Barbieri F, PerettiM, Pizzi E, et al.Metforminrepositioning as antitumoral agent: selective antiproliferative effects inhuman glioblastoma stem cells, via inhibition of CLIC1-mediated ioncurrent. Oncotarget 2014;5:11252–68.

6. SettiM, Savalli N,Osti D, Richichi C, AngeliniM, Brescia P, et al. Functionalrole of CLIC1 ion channel in glioblastoma-derived stem/progenitor cells.J Natl Cancer Inst 2013;105:1644–55.

7. Valenzuela SM, Martin DK, Por SB, Robbins JM, Warton K, Bootcov MR,et al. Molecular cloning and expression of a chloride ion channel of cellnuclei. J Biol Chem 1997;272:12575–82.

8. Tonini R, Ferroni A, Valenzuela SM,Warton K, Campbell TJ, Breit SN, et al.Functional characterization of the NCC27 nuclear protein in stable trans-fected CHO-K1 cells. FASEB J 2000;14:1171–8.

9. Averaimo S, Milton RH, Duchen MR, Mazzanti M. Chloride intracellularchannel 1 (CLIC1): sensor and effector during oxidative stress. FEBS Lett2010;584:2076–84.

10. Peretti M, Angelini M, Savalli N, Florio T, Yuspa SH, Mazzanti M.Chloride channels in cancer: focus on chloride intracellular channel 1and 4 (CLIC1 AND CLIC4) proteins in tumor development andas novel therapeutic targets. Biochim Biophys Acta 2015;1848:2523–31.

11. Murzin AG. Biochemistry. Metamorphic proteins. Science 2008;320:1725–6.

12. Milton RH, Abeti R, Averaimo S, DeBiasi S, Vitellaro L, Jiang L, et al. CLIC1function is required for beta-amyloid-induced generation of reactiveoxygen species by microglia. J Neurosci 2008;28:11488–99.

13. NovarinoG, Fabrizi C, Tonini R,DentiMA,Malchiodi-Albedi F, LauroGM,et al. Involvement of the intracellular ion channel CLIC1 in microglia-mediated beta-amyloid-induced neurotoxicity. J Neurosci 2004;24:5322–30.

Peretti et al.

Mol Cancer Ther; 17(11) November 2018 Molecular Cancer Therapeutics2460

on August 11, 2021. © 2018 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst August 22, 2018; DOI: 10.1158/1535-7163.MCT-17-1223

Page 11: MutualInfluenceofROS,pH,andCLIC1Membrane Protein in the ...Marta Peretti1, Federica Maddalena Raciti1,Valentina Carlini1, Ivan Verduci1, Sarah Sertic1, Sara Barozzi2, Massimiliano

14. Tulk BM, Kapadia S, Edwards JC. CLIC1 inserts from the aqueous phaseinto phospholipid membranes, where it functions as an anion channel.Am J Physiol Cell Physiol 2002;282:C1103–12.

15. Valenzuela SM, Mazzanti M, Tonini R, Qiu MR, Warton K, Musgrove EA,et al. The nuclear chloride ion channel NCC27 is involved in regulation ofthe cell cycle. J Physiol 2000;529:541–52.

16. Averaimo S, Abeti R, Savalli N, Brown LJ, Curmi PM, Breit SN, et al. Pointmutations in the transmembrane region of the clic1 ion channel selectivelymodify its biophysical properties. PLoS One 2013;8:e74523.

17. Littler DR, Harrop SJ, Goodchild SC, Phang JM, Mynott AV, Jiang L, et al.The enigma of the CLIC proteins: Ion channels, redox proteins, enzymes,scaffolding proteins? FEBS Lett 2010;584:2093–101.

18. Madshus IH. Regulation of intracellular pH in eukaryotic cells. Biochem J1988;250:1–8.

19. Harguindey S,Orive G, Luis Pedraz J, Paradiso A, Reshkin SJ. The role of pHdynamics and the Naþ/Hþ antiporter in the etiopathogenesis and treat-ment of cancer. Two faces of the same coin–one single nature. BiochimBiophys Acta 2005;1756:1–24.

20. McLean LA, Roscoe J, JorgensenNK,Gorin FA, Cala PM.Malignant gliomasdisplay altered pH regulation by NHE1 compared with nontransformedastrocytes. Am J Physiol Cell Physiol 2000;278:C676–88.

21. Menon SG, Goswami PC. A redox cycle within the cell cycle: ring in the oldwith the new. Oncogene 2007;26:1101–9.

22. Sarsour EH, Kumar MG, Chaudhuri L, Kalen AL, Goswami PC. Redoxcontrol of the cell cycle in health and disease. Antioxid Redox Signal2009;11:2985–3011.

23. Menon SG, Coleman MC, Walsh SA, Spitz DR, Goswami PC. Differentialsusceptibility of nonmalignant human breast epithelial cells and breastcancer cells to thiol antioxidant-induced G(1)-delay. Antioxid RedoxSignal 2005;7:711–8.

24. BokochGM,KnausUG.NADPHoxidases: not just for leukocytes anymore!Trends Biochem Sci 2003;28:502–8.

25. Di Mascio P, Murphy ME, Sies H. Antioxidant defense systems: the role ofcarotenoids, tocopherols, and thiols. Am J Clin Nutr 1991;53:194S–200S.

26. Oberley LW, Oberley TD, Buettner GR. Cell division in normal andtransformed cells: the possible role of superoxide and hydrogen peroxide.Med Hypotheses 1981;7:21–42.

27. Burhans WC, Heintz NH. The cell cycle is a redox cycle: linking phase-specific targets to cell fate. Free Radic Biol Med 2009;47:1282–93.

28. McEwen BS, Wingfield JC. The concept of allostasis in biology andbiomedicine. Horm Behav 2003;43:2–15.

29. Carra E, Barbieri F, Marubbi D, Pattarozzi A, Favoni RE, Florio T, et al.Sorafenib selectively depletes human glioblastoma tumor-initiating cellsfrom primary cultures. Cell Cycle 2013;12:491–500.

30. Griffero F, Daga A, Marubbi D, Capra MC, Melotti A, Pattarozzi A, et al.Different response of human glioma tumor-initiating cells to epidermalgrowth factor receptor kinase inhibitors. J Biol Chem 2009;284:7138–48.

31. BajettoA, Porcile C, Pattarozzi A, Scotti L, AcetoA,DagaA, et al.Differentialrole of EGF and BFGF in human GBM-TIC proliferation: relationship toEGFR-tyrosine kinase inhibitor sensibility. J Biol Regul Homeost Agents2013;27:143–54.

32. Wurth R, Pattarozzi A, Gatti M, Bajetto A, Corsaro A, Parodi A, et al.Metformin selectively affects human glioblastoma tumor-initiating cellviability: A role for metformin-induced inhibition of Akt. Cell Cycle 2013;12:145–56.

33. Chow S, Hedley D. Flow cytometric measurement of intracellular pH. CurrProtoc Cytom 2001;Chapter 9:Unit 9 3.

34. Pattarozzi A,Carra E, Favoni RE,Wurth R,MarubbiD, Filiberti RA, et al. Theinhibition of FGF receptor 1 activity mediates sorafenib antiproliferativeeffects in human malignant pleural mesothelioma tumor-initiating cells.Stem Cell Res Ther 2017;8:119.

35. Schindelin J, Arganda-Carreras I, Frise E, Kaynig V, Longair M, Pietzsch T,et al. Fiji: an open-source platform for biological-image analysis. NatMethods 2012;9:676–82.

36. Neveu B, Spinella JF, Richer C, Lagace K, Cassart P, Lajoie M, et al. CLIC5: anovel ETV6 target gene in childhood acute lymphoblastic leukemia.Haematologica 2016;101:1534–43.

37. Fanucchi S, AdamsonRJ,DirrHW. Formationof anunfolding intermediatestate of soluble chloride intracellular channel protein CLIC1 at acidic pH.Biochemistry 2008;47:11674–81.

38. Hsieh CH, Shyu WC, Chiang CY, Kuo JW, Shen WC, Liu RS. NADPHoxidase subunit 4-mediated reactive oxygen species contribute to cyclinghypoxia-promoted tumor progression in glioblastoma multiforme. PLoSOne 2011;6:e23945.

39. Steffan JJ, Williams BC, Welbourne T, Cardelli JA. HGF-induced invasionby prostate tumor cells requires anterograde lysosome trafficking andactivity of Naþ-Hþ exchangers. J Cell Sci 2010;123:1151–9.

40. Stoychev SH, Nathaniel C, Fanucchi S, Brock M, Li S, Asmus K, et al.Structural dynamics of soluble chloride intracellular channel proteinCLIC1 examined by amide hydrogen-deuterium exchange mass spectrom-etry. Biochemistry 2009;48:8413–21.

41. Warton K, Tonini R, Fairlie WD, Matthews JM, Valenzuela SM, Qiu MR,et al. Recombinant CLIC1 (NCC27) assembles in lipid bilayers via a pH-dependent two-state process to form chloride ion channels with identicalcharacteristics to those observed in Chinese hamster ovary cells expressingCLIC1. J Biol Chem 2002;277:26003–11.

42. Blum R, Kloog Y. Metabolism addiction in pancreatic cancer. Cell DeathDis 2014;5:e1065.

43. Brown DG, Rao S, Weir TL, O'Malia J, Bazan M, Brown RJ, et al. Metabo-lomics and metabolic pathway networks from human colorectal cancers,adjacent mucosa, and stool. Cancer Metab 2016;4:11.

44. Masui K, Cavenee WK, Mischel PS. Cancer metabolism as a central drivingforce of glioma pathogenesis. Brain Tumor Pathol 2016;33:161–8.

45. Romero-Garcia S, Lopez-Gonzalez JS, Baez-Viveros JL, Aguilar-Cazares D,Prado-GarciaH. Tumor cell metabolism: an integral view. Cancer Biol Ther2011;12:939–48.

46. Pollak J, Rai KG, FunkCC,Arora S, Lee E, Zhu J, et al. Ion channel expressionpatterns in glioblastoma stem cells with functional and therapeutic impli-cations for malignancy. PLoS One 2017;12:e0172884.

47. Schlenska-Lange A, Knupfer H, Lange TJ, Kiess W, Knupfer M. Cell prolif-eration andmigration in glioblastomamultiforme cell lines are influencedby insulin-like growth factor I in vitro. Anticancer Res 2008;28:1055–60.

48. Shono T, Yokoyama N, Uesaka T, Kuroda J, Takeya R, Yamasaki T, et al.Enhanced expression of NADPH oxidase Nox4 in human gliomas and itsroles in cell proliferation and survival. Int J Cancer 2008;123:787–92.

49. Di Cristofori A, Ferrero S, Bertolini I, Gaudioso G, RussoMV, Berno V, et al.The vacuolar Hþ ATPase is a novel therapeutic target for glioblastoma.Oncotarget 2015;6:17514–31.

50. Wurth R, Thellung S, Bajetto A, Mazzanti M, Florio T, Barbieri F. Drug-repositioning opportunities for cancer therapy: novel molecular targets forknown compounds. Drug Discov Today 2016;21:190–9.

www.aacrjournals.org Mol Cancer Ther; 17(11) November 2018 2461

CLIC1 Activity Controls Glioblastoma Stem Cell Proliferation

on August 11, 2021. © 2018 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst August 22, 2018; DOI: 10.1158/1535-7163.MCT-17-1223

Page 12: MutualInfluenceofROS,pH,andCLIC1Membrane Protein in the ...Marta Peretti1, Federica Maddalena Raciti1,Valentina Carlini1, Ivan Verduci1, Sarah Sertic1, Sara Barozzi2, Massimiliano

2018;17:2451-2461. Published OnlineFirst August 22, 2018.Mol Cancer Ther   Marta Peretti, Federica Maddalena Raciti, Valentina Carlini, et al.   Stem Cells

S Phase Progression in Human Glioblastoma−1Regulation of GMutual Influence of ROS, pH, and CLIC1 Membrane Protein in the

  Updated version

  10.1158/1535-7163.MCT-17-1223doi:

Access the most recent version of this article at:

  Material

Supplementary

  http://mct.aacrjournals.org/content/suppl/2018/08/22/1535-7163.MCT-17-1223.DC1

Access the most recent supplemental material at:

   

   

  Cited articles

  http://mct.aacrjournals.org/content/17/11/2451.full#ref-list-1

This article cites 50 articles, 12 of which you can access for free at:

  Citing articles

  http://mct.aacrjournals.org/content/17/11/2451.full#related-urls

This article has been cited by 1 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/17/11/2451To request permission to re-use all or part of this article, use this link

on August 11, 2021. © 2018 American Association for Cancer Research. mct.aacrjournals.org Downloaded from

Published OnlineFirst August 22, 2018; DOI: 10.1158/1535-7163.MCT-17-1223