Synergistic Interaction between Histone Deacetylase and ...

2 downloads 100 Views 766KB Size Report
Materials and Methods. Chemicals and antibodies. Trichostatin A, sodium butyrate (NaB), valproic acid, mitoxantrone, and teniposide (VM-26) were purchased.
Cancer Therapy: Preclinical

Synergistic Interaction between Histone Deacetylase and Topoisomerase II Inhibitors Is Mediated through Topoisomerase IIB Douglas C. Marchion, Elona Bicaku, Joel G. Turner, Adil I. Daud, Daniel M. Sullivan, and Pamela N. Munster

Abstract

Background: DNA topoisomerase II inhibitors and poisons are among the most efficacious drugs for the treatment of cancer. Sensitivity of cancer cells to the cytotoxic effects of topoisomerase II targeting agents is thought to depend on the expression of the topoisomerase IIa isoform, and drug resistance is often associated with loss or mutation of topoisomerase IIa. Histone deacetylase inhibitors (HDACi) are a novel class of compounds that potentiate the antitumor effects of topoisomerase II ^ targeting agents. Methods: The interaction between HDACi and topoisomerase II ^ targeting agents in cancer cells was evaluated as a function of topoisomerase IIa and topoisomerase IIh expression. Topoisomerase II isoforms were selectively depleted using small interfering RNA and antisense. Druginduced formation of cleavable complexes involving topoisomerase IIa and topoisomerase IIh was evaluated by trapped-in-agarose DNA immunostaining and band depletion assays in the presence and absence of HDACi. Results: Preexposure to HDACi increased the cytotoxicity of topoisomerase II poisons.This was associated with a down-regulation of topoisomerase IIa expression but had no effects on topoisomerase IIh. In the setting of HDACi-induced chromatin decondensation and topoisomerase IIa depletion, topoisomerase II poison cytotoxicity was mediated through topoisomerase IIh cleavable complex formation. The HDACi-induced sensitization was also observed in cells with targetspecific resistance to topoisomerase II poisons. Conclusions: The recruitment of topoisomerase IIh as a target may overcome primary or emergent drug resistance to topoisomerase II ^ targeting agents and hence may broaden the applicability of this important class of anticancer agents.

Type

II topoisomerases are essential enzymes that regulate the topological state of DNA to facilitate several biological processes. Topoisomerase II exists in two isoforms, topoisomerase IIa and topoisomerase IIh. Although these enzymes have high sequence homology (1) and a similar mechanism of action (2), they are regulated independently and have been associated with different cellular functions. Topoisomerase IIa is necessary for many biological processes involving doublestranded DNA, including replication, mitosis, and chromatin condensation (3, 4). Topoisomerase IIa is mainly located in the nucleus and peaks in expression during G2-M (5, 6). In contrast, topoisomerase IIh expression remains stable through-

Authors’ Affiliation: Experimental Therapeutics Program, Department of Interdisciplinary Oncology, H. Lee Moffitt Cancer Center and Research Institute, Tampa, Florida Received 5/13/05; revised 8/15/05; accepted 8/31/05. Grant support: Susan G. Koman Breast Cancer Foundation and Don Shula Career Development Award. The costs of publication of this article were defrayed in part by the payment of page charges. This article must therefore be hereby marked advertisement in accordance with 18 U.S.C. Section 1734 solely to indicate this fact. Requests for reprints: Pamela N. Munster, H. Lee Moffitt Cancer Center, 12902 Magnolia Drive, MRC 4E,Tampa, FL 33612. Phone: 813-745-8948; Fax: 813-7451984; E-mail: Munstepn@ moffitt.usf.edu. F 2005 American Association for Cancer Research. doi:10.1158/1078-0432.CCR-05-1073

www.aacrjournals.org

out the cell cycle (7) and may have an important role in transcription (8). Although it was reported that topoisomerase IIh may not be essential in mitosis, topoisomerase IIa knockdown experiments have shown that topoisomerase IIh may only partially substitute topoisomerase IIa during chromatin decondensation and cell segregation (9). Due to their essential roles, topoisomerase II enzymes have been successfully targeted for the treatment of cancer. To maintain genetic integrity during the cleavage and religation of DNA, both topoisomerase isoenzymes form transient bonds with the cleaved DNA called the cleavable complex (10). The stabilization and persistence of these cleavable complexes by topoisomerase II poisons have been associated with DNA damage and cell death (11, 12). Multiple in vitro and in vivo studies have suggested that sensitivity of tumor cells to topoisomerase II poisons depends on access to the targets, as well as target location and expression levels. For clinical use, topoisomerase IIa has been considered the more relevant target (13 – 17). A class of drugs that may enhance access to DNA and thereby increase the antitumor activity of topoisomerase II – targeting agents are the histone deacetylase inhibitors (HDACi; refs. 14, 18, 19). We previously reported a sequence-specific and timedependent potentiation of topoisomerase II – targeting agents by the HDACi, suberoylanilide hydroxamic acid (SAHA), and valproic acid (14, 20). Our data indicated that prolonged

8467

Clin Cancer Res 2005;11(23) December 1, 2005

Cancer Therapy: Preclinical

treatment of cancer cells with an HDACi resulted in the depletion of proteins involved in the maintenance of heterochromatin and led to subsequent chromatin decondensation. Chromatin decondensation increased the access of topoisomerase II – targeting agents to the DNA substrate, which potentiated cell death induced by topoisomerase II – targeting agents. Topoisomerase II drug cytotoxicity is largely dependent upon topoisomerase IIa expression. Here, we investigated the relevance of topoisomerase IIa and topoisomerase IIh as targets of topoisomerase II poisons in the presence of HDACi. We show that in the setting of HDACi pretreatment, the formation of cleavable complexes involving topoisomerase IIh rather than topoisomerase IIa leads to topoisomerase II poison – induced cell death. This may broaden the clinical applicability of topoisomerase II poisons and overcome drug resistance due to acquired depletion and mutations of topoisomerase IIa.

Materials and Methods Chemicals and antibodies. Trichostatin A, sodium butyrate (NaB), valproic acid, mitoxantrone, and teniposide (VM-26) were purchased from Sigma Chemical Co. (St. Louis, MO). SAHA was provided by Aton Pharma (Merck, Whitehouse Station, NJ). Epirubicin was purchased from Pfizer, Inc. (New York, NY). XK469 was a kind gift from Dr. Robert Snapka. All other reagents were of analytic grade and purchased from standard suppliers. Antibodies used were as follows: Ki-S1 (monoclonal, Chemicon, Temecula, CA) and 454 (polyclonal, developed by Dr. Dan Sullivan) for topoisomerase IIa, anti-topoisomerase IIh (monoclonal, BD Biosciences, San Jose, CA) and JAB (polyclonal, developed by Dr. Dan Sullivan), and acetylated histone H3 antibody (Upstate Biotechnology, Chicago, IL). Cell lines. SKBr-3, MCF-7, KM12C, A375, BT-474, and MDA-MB361 cells were purchased from the American Type Culture Collection (Manassas, VA). Doxorubicin-resistant MCF-7 cells (MCF:Dox) were a kind gift from Dr. Fred Hausheer. Cell lines were maintained in DMEM supplemented with 10% heat-inactivated fetal bovine serum (FBS), 2 mmol/L glutamine, and 50 units/mL penicillin, and 50 Ag/mL streptomycin (Life Technologies Bethesda Research Laboratories, Carlsbad, CA). Cells were incubated in a humidified atmosphere with 5% CO2 at 37jC. Cytotoxicity assays. Topoisomerase II poison cytotoxicity was evaluated in the presence of HDACi by apoptotic and clonogenic assays. Apoptosis was scored by the presence of nuclear chromatin condensation and DNA fragmentation and evaluated with fluorescence microscopy using bis-benzimide staining. Briefly, cells were treated with an HDACi for 48 hours, after which the media was removed and replaced with media containing the indicated concentrations of a topoisomerase II poison. After 4 hours, the topoisomerase II poison was removed, and the cells were cultured for an additional 48 hours. Cells were harvested using a Cell Scraper (Fisher, Hampton, NH), fixed in 4% paraformaldehyde for 10 minutes at room temperature, and washed with PBS. Cell nuclei were stained with 0.5 Ag/mL of bisbenzimide trihydrochloride (Hoechst #33258, Molecular Probes, Eugene, OR). Two hundred cells were counted for each experiment and evaluated for apoptotic scores (apoptotic nuclei/all nuclei  100). Each experiment was repeated thrice, and the SE was calculated. For clonogenic assays, cells were plated on six-well dishes at a density of 150 per well and allowed to adhere for 24 hours. For drug combination studies, cells were incubated with medium containing 0.5 mmol/L valproic acid. At this concentration of valproic acid, valproic acid had no discernable effects on growth or apoptosis. After a 48-hour treatment with valproic acid, the medium was removed and replaced with medium containing 0, 1, 10, 25, 50, or 100 nmol/L epirubicin for

Clin Cancer Res 2005;11(23) December 1, 2005

4 hours. Epirubicin was then removed, and colonies were allowed to grow for 14 to 21 days to a maximal size of 2 to 3 mm in diameter, stained with 2% crystal violet in methanol, and counted. Colonies were included in the assessment if measuring at least 0.2 mm. For single-drug samples and untreated controls, saline was used in lieu of valproic acid or epirubicin or both. All experiments were done in duplicates and repeated at least thrice. The concentration of epirubicin required for IC50 was determined using the CalcuSyn software as described previously (14). Synergistic effects versus additive effects were also determined by the CalcuSyn program. The fractional inhibition of valproic acid was depicted as the number of observed cells (treatment group) divided by the number of expected cells (untreated control) with a range of 0 to 1. Microarray. Expression levels of topoisomerase IIa mRNA were evaluated by microarray analysis using Affymetrix Genechips (Affymetrix, Santa Clara, CA) by standard protocols (Moffitt, Molecular Biology Core). Hybridization to Affymetrix chips was analyzed using Affymetrix Microarray Suite 5.0 software. Signal intensity was scaled to an average intensity of 500 before comparison analysis. The MAS 5.0 software uses a statistical algorithm to assess changes in mRNA abundance in a direct comparison between two samples (Statistical algorithms description document. http://www.affymetrix. com/support/technical/whitepapers.affx). This analysis is based on the behavior of 16 different oligonucleotide probes designed to detect the same gene. Using the programmed default values, probe sets that yielded a change P < 0.04 were identified as changed (increased or decreased), and those that yielded a P between 0.04 and 0.06 were identified as marginally changed. Western blot analysis. Samples were prepared using SDS lysis buffer [2% SDS, 10% glycerol, 0.06 mol/L Tris (pH 6.8)] and evaluated for protein concentration using the bicinchoninic acid method (Pierce, Rockford, IL). Proteins (50 Ag) were separated on 8% SDS-PAGE gels and transferred to nitrocellulose membranes. Membranes were blocked in tris-buffered saline containing 0.05% Tween 20 (TBST), 5% nonfat milk and incubated with primary antibody in TBST, 5% nonfat milk, overnight at 4jC. Membranes were washed thrice for 10 minutes with TBST and incubated with the appropriate secondary antibody in TBST, 5% nonfat milk for 90 minutes at room temperature. Antibody binding was visualized by chemiluminescence on autoradiography film. Relative expression of proteins was determined by densitometry analysis of at least two scanned autoradiography films from independent experiments using Photoshop software. Immunofluorescence. MCF-7 cells were treated with 2 mmol/L valproic acid for 48 hours. Cells were harvested by trypsinization, washed in PBS, and adhered to glass slide using Cytospin Funnels (Shandon, Pittsburgh, PA) at a density of 1  105/mL. Slides were blocked with 2% bovine serum albumin (BSA) in PBS for 1 hour and probed with antibodies specific for topoisomerase IIa, KiS1 (monoclonal, 1:50), and acetylated histone H3 (polyclonal, 1:100) in 1% BSA/PBS for 1 hour at room temperature in a humidified chamber. Slides were washed in PBS and incubated with anti-rabbit Alexa Fluor 488 and anti-mouse Alexa Fluor 546 (Molecular Probes) diluted 1:100 in 1% BSA/PBS containing goat serum for 1 hour at room temperature in a humidified chamber. Slides were washed with PBS and coverslipped using Prolong Gold mounting media (Molecular Probes). Images were acquired by confocal microscopy. The square pixel surface area for the respective protein expression was analyzed in at least 50 cells per treatment using Photoshop software. For each measured nucleus, the background staining was measured in the immediate vicinity and subtracted from the average square pixel surface area. Evaluation of secondary antibody staining only served as internal staining control. Levels for the fluorescence detection were set accordingly. Experiments were repeated at least twice. Antisense/small interfering RNA. Topoisomerase IIa and topoisomerase IIh protein expression was blocked with antisense oligonucleotides and small interfering RNA (siRNA) duplexes. Antisense: The DNA sequences for the oligonucleotide probes were

8468

www.aacrjournals.org

Interaction of HDAC and Topoisomerase II Inhibitors

as follows: 5V-CTGCAATGGTGACACTTCCAT-3V for topoisomerase IIa and 5V-TTTGTAGTGGACAGAAACACAGTA-3V for topoisomerase IIh as described by Towatari et al. (21). Oligonucleotide (1 Ag) was suspended in 100 AL Optimem (Life Technologies Bethesda Research Laboratories) containing 2 AL Superfect transfection reagent (Qiagen, Valencia, CA). The oligonucleotide suspension was raised to a total volume of 700 AL by the addition of DMEM containing 10% fetal bovine serum and incubated on cell monolayers (1  105 cells) for 24 hours. The oligonucleotide mixture was replaced the following day, and the cells were incubated for an additional 4 hours before experimental procedures. Controls included topoisomerase IIa and topoisomerase IIh sense oligonucleotides as well as incubation of cells with Superfect in the absence of oligonucleotide. siRNA: RNA duplexes for topoisomerase IIa (sense, GGUAUUCCUGUUGUUGAAC) and topoisomerase IIh (sense, GGUUACCUUUGUGCCAGGU) were purchased from Ambion (Austin, TX). Cells were suspended in 0.1 mL siPort electroporation buffer (3  106/mL, Ambion), mixed with 1 Ag siRNA, and pulsed with 300 V for 0.5 milliseconds. Pulsed cells were incubated at 37jC for 15 minutes before experimentation. The Silencer Negative Control #2 siRNA (Ambion), a nonsense siRNA duplex, was used as a control. Trapped-in-agarose DNA immunostaining assay. The trapped-inagarose DNA immunostaining assay (TARDIS) was done as described by Willmore et al. (22) with modifications. Cells (2  104 per slide) were imbedded in agarose and lysed [2.5 mol/L NaCl, 100 mmol/L EDTA (pH 10), 10 mmol/L Tris, 1% sarkosyl, 1% Triton X-100] in the presence of protease inhibitors (2 Ag/mL pepstatin, 10 Ag/mL aprotinin, 10 Ag/mL leupeptin, 0.2 mmol/L Na3VO4, and 1 mmol/L phenylmethylsulfonyl fluoride) for 30 minutes. Imbedded cells were washed for 20 minutes with 1 mol/L NaCl in the presence of protease inhibitors and incubated in 2% BSA in PBS for 1 hour. Immunostaining consisted of anti-topoisomerase IIa (454, polyclonal, 1:200) and anti-topoisomerase IIh (monoclonal, 1:100) in 0.5% BSA for 1 hour at room temperature in a humidified chamber. Slides were washed twice for 10 minutes in 0.5% BSA and incubated with the appropriate secondary antibodies (anti-rabbit 546 and anti-mouse 488; Molecular Probes) at a concentration of 1:200 in 0.5% BSA containing goat serum (1:100) for 1 hour at room temperature. Slide were washed with 0.5%

Fig. 1. HDACi potentiate topoisomerase II inhibitor cytotoxicity. MCF_7 breast cancer cells were treated with 2 mmol/L valproic acid (VPA) or 0.5 Amol/L SAHA for 48 hours followed by exposure to 0.5 Amol/L epirubicin or mitoxantrone for 4 hours. Cells were harvested 24 hours later and evaluated for nuclear condensation and fragmentation. Columns, mean from at least three replicate experiments; bars, SE.

www.aacrjournals.org

BSA for 10 minutes, dried, and counterstained with 0.5 Ag/mL of bisbenzimide trihydrochloride (Hoechst #33258, Molecular Probes). Images were acquired by confocal microscopy and the square pixel surface area of the respective protein was analyzed for a minimum of 50 cells per treatment group. Statistical analysis was done by ANOVA. Experiments were repeated at least twice. Band depletion. The band depletion assay was done as described by Xiao et al. (8). Briefly, 5  105 cells were lysed in alkaline lysis solution (200 mmol/L NaOH, 2 mmol/L EDTA), and the lysate was neutralized [neutralization buffer: 1 mol/L HCl, 600 mmol/L Tris (pH 8.0)]. The neutralized lysate was then mixed with 3 SDS sample buffer [150 mmol/L Tris-HCl (pH 6.8), 6 mmol/L EDTA, 45% sucrose, 9% SDS, 10% h-mercaptoethanol], and the lysates were separated on 8% SDSPAGE gels.

Results Interactions of topoisomerase II poisons and histone deacetylase inhibitors. We previously reported an increased sensitivity of tumor cells to topoisomerase II – targeting agents after pretreatment with an HDACi in vitro and in vivo (14, 20). Treatment with HDACi led to dose- and time-dependent histone acetylation and subsequent modulation of genes and proteins essential for the maintenance of heterochromatin. The ensuing chromatin decondensation was associated with an increased binding of topoisomerase II poisons to the DNA substrate and in the presence of topoisomerase II, with increased DNA damage and cell death. The sensitivity of cancer cells to topoisomerase II – targeting agents has been linked to expression of the topoisomerase IIa isoform, whereas studies on the role of topoisomerase IIh have been more limited. Our experimental data indicate that preexposure of tumor cells to HDACi potentiates the apoptosis induced by the topoisomerase II poisons epirubicin and mitoxantrone (Fig. 1A). Potentiation was not limited to a certain class of HDACi and was observed with selective (SAHA and trichostatin A) as well as with nonselective (valproic acid) HDACi. Similarly, potentiation occurred with other topoisomerase II – targeting agents, such as doxorubicin, etoposide, and VM-26 (see below; data not shown; and refs. 14, 20). Densitometry analysis of Western blots (n = 3) evaluating cancer cells derived from different organs, including breast, colon, and melanoma, showed that the examined cell lines with decreased or mutated topoisomerase IIa, required higher concentration of epirubicin for IC50 (Table 1) as assessed by colony-forming assays. As reported by other investigators, these data suggest that not organ specificity but rather expression of the topoisomerase II isoforms was predictive for response (13). A 48-hour preexposure of these cells to the HDACi valproic acid resulted in a decrease in the IC50 of epirubicin. The concentration of valproic acid used for these experiments affected 5 Amol/L) that were shown to result in a G2-M arrest, the cell cycle phase where topoisomerase IIa is increased (23). Relevance of topoisomerase IIb as a target. Potentiation of topoisomerase II targeting agents by HDACi in the setting of topoisomerase IIa depletion argued for the involvement of the alternative topoisomerase II isoform topoisomerase IIh. Synergy between an HDACi and a topoisomerase II – targeting agent was preserved when using the topoisomerase IIh – specific poison, XK469, signifying a potential role of topoisomerase IIh as a target for drug therapy. A potentiation of XK469 by preexposure to valproic acid was found in several examined breast cancer cell lines, including SKBr-3, MCF-7, and the topoisomerase IIa – depleted MCF:Dox cells (Fig. 3A). To further define the involvement of topoisomerase IIh in the interaction between HDACi and topoisomerase II poisons, both topoisomerase II isoforms were alternatively depleted by siRNA or antisense and exposed to HDACi before administration of a topoisomerase II poison. We have shown above that treatment with HDACi resulted in a down-regulation of topoisomerase IIa (Fig. 2). Treatment of MCF-7 cells with siRNA resulted in significant depletion of topoisomerase IIa and topoisomerase IIh protein levels without affecting viability (Fig. 3B; data not shown). Topoisomerase IIa was further depleted with the addition of valproic acid (Fig. 3B). Depletion of topoisomerase IIa by siRNA did not effect the potentiation of epirubicin, mitoxantrone, or VM-26 by valproic acid (Fig. 3C). In contrast, the synergistic activity between HDACi and these topoisomerase II – targeting drugs was abrogated by the depletion of topoisomerase IIh (Fig. 3C). These findings were mirrored when antisense was used to deplete topoisomerase IIa or topoisomerase IIh or when SAHA was used in lieu of valproic acid (data not shown). Although the topoisomerase IIa siRNA was exquisitely specific, the topoisomerase IIh siRNA showed a slight offtarget effect on the expression of topoisomerase IIa. This offtarget effect unlikely contributed to the abrogation of synergy between valproic acid and topoisomerase II poisons because

8470

www.aacrjournals.org

Interaction of HDAC and Topoisomerase II Inhibitors

the direct depletion of topoisomerase IIa by siRNA in the absence of HDACi effects and the further depletion of topoisomerase IIa by the addition of valproic acid did not sensitize cells to topoisomerase II poisons. Hence, in the presence of an HDACi, topoisomerase II poison cytotoxicity seems mediated through topoisomerase IIh. Histone deacetylase inhibitor increases the formation of DNA/ topoisomerase IIb cleavable complexes. Treatment with HDACi result in chromatin decondensation and increased access of topoisomerase II – targeting agents to the DNA substrate

Fig. 3. Topoisomerase IIh (topo IIb) is the prime target of topoisomerase II inhibitors. A, % apoptotic nuclei in cells treated with 2 mmol/L valproic acid (VPA) for 48 hours followed by 0, 0.1, and 0.5 mmol/L of the topoisomerase IIh ^ specific inhibitor XK469 for 48 hours in the cell lines MCF-7 and SKBr-3 and the topoisomerase IIa ^ depleted MCF:Dox. B, expression of topoisomerase IIa and topoisomerase IIh in MCF-7 cells after exposure to siRNA and 2 mmol/L valproic acid for 48 hours. Densitometry analysis of the protein bands is listed as % expression relative to control samples. C, % apoptotic nuclei in cells treated with 2 mmol/L valproic acid followed by 0.5 Amol/L epirubicin, 0.5 Amol/L mitoxantrone, or 1 Amol/L VM-26 for 4 hours, after depletion of topoisomerase IIa and topoisomerase IIh by siRNA. Columns, mean from at least three different replicate experiments; bars, SE.

Fig. 2. HDACi exposure depletes topoisomerase IIa (topo IIa) but not topoisomerase IIh. A, Western blot depicting changes in the expression of topoisomerase IIa in log-phase MCF-7 cells after exposure to 2 Amol/L SAHA for the indicated times. B, expression of topoisomerase IIa and topoisomerase IIh in log-phase MCF-7 cells treated with 2 mmol/L valproic acid (VPA), 0.5 Amol/L SAHA, 0.05 Amol/L trichostatin A (TSA), and 2 mmol/L NaB for 48 hours. C, microarray analysis showing the relative topoisomerase IIa mRNA expression after 0, 4, 24, and 48 hours of incubation with 2 mmol/L valproic acid. D, expression and location of topoisomerase IIa and acetylation of histone H3 in MCF-7 cells after exposure to 2 mmol/L valproic acid for 48 hours using confocal microscopy at a 63 magnification.

www.aacrjournals.org

(14, 20). Topoisomerase II poisons are thought to confer cytotoxicity through the stabilization of topoisomerase II/DNA cleavable complexes. We evaluated whether the presence of HDACi resulted in the involvement of a specific topoisomerase II isoform in the cleavable complexes stabilized by topoisomerase II poisons. The cleavable complexes involving topoisomerase IIa or topoisomerase IIh induced by epirubicin in the presence and absence of a 48-hour pretreatment with 2 mmol/L valproic acid were evaluated by trapped-in-agarose DNA immunostaining assay. Immunofluorescence analysis of confocal microscopy images showed that at concentrations of epirubicin with minimal effect on the stabilization of cleavable complexes involving topoisomerase IIa (P > 0.05), there was a 3-fold increase in the accumulation of topoisomerase IIh – containing cleavable complexes stabilized by epirubicin in the presence of valproic acid (Fig. 4A and B). This difference reached statistical

8471

Clin Cancer Res 2005;11(23) December 1, 2005

Cancer Therapy: Preclinical

significance (*, P < 0.001). Similar results were observed when mitoxantrone was used in lieu of epirubicin (data not shown). Valproic acid by itself did not induce the formation of cleavable complexes. As described above, valproic acid treatment results in depletion of topoisomerase IIa. To further establish whether these findings were intrinsic to the anthracycline family members (intercalating topoisomerase II poisons), band depletion assays were performed using the nonintercalating topoisomerase II poison, VM-26. This drug has been reported to induce cell death with the preferential formation of topoisomerase IIa – containing cleavable complexes (24). The band pattern of topoisomerase IIa and topoisomerase IIh was evaluated in MCF-7 cells exposed to VM-26 and the HDACi valproic acid (Fig. 4C). Band depletion assays allow detection of topoisomerase IIa and topoisomerase IIh complexed with DNA. The topoisomerase enzymes that are covalently bound to DNA (cleavable complexes) are prohibited from entering the acrylamide gel resulting in a band depletion, whereas topoisomerase II enzymes not bound to DNA readily enter the gel. Exposure of cells to VM-26 resulted in a 52% depletion of topoisomerase IIh band, whereas no effects were seen with valproic acid alone. In cells preexposed to 2 mmol/L valproic acid, VM-26 induced an 81% reduction in the topoisomerase IIh band, suggesting an increase in cleavable complex formation Fig. 4C (bottom). VM-26 induced a 51% depletion in the topoisomerase IIa band. There was a 46% decrease in the topoisomerase IIa band after treatment with valproic acid alone. In cells treated with valproic acid before VM-26, there was a 92% depletion of the topoisomerase IIa band (Fig. 4C, top). However, as discussed above, treatment with valproic acid resulted in a downregulation of topoisomerase IIa protein expression (Fig. 2A and B and Fig. 3B). The further reduction in the topoisomerase IIa band by VM-26 in the presence of valproic acid resulted from the additive effects of the HDACi-induced downregulation of the protein and the loss of topoisomerase IIa band due to cleavable complex formation induced by VM-26. This is further supported by the trapped-in-agarose DNA immunostaining assay showing no increase in the topoisomerase IIa – cleavable complexes induced by valproic acid alone. These findings suggest that treatment of cell with valproic acid enhanced the formation of cleavable complexes involving topoisomerase IIh but not topoisomerase IIa in the presence of VM-26.

Discussion Topoisomerase II is an essential enzyme for cell survival, and the increased expression of topoisomerase II in many cancer cells has made these enzymes attractive targets for cancer therapy. Topoisomerase II exists in two isoforms, topoisomerase IIa and topoisomerase IIh, with partial functional redundancy (9, 25 – 28). In yeast, either isoform may substitute for the endogenous topoisomerase II enzyme (29). Murine transgenic studies have shown that although topoisomerase IIa may compensate for a loss of topoisomerase IIh during embryogenesis (30), the loss of topoisomerase IIa leads to embryonic death (31). However, topoisomerase IIh seems essential for neuromuscular development after birth (30). In adult mammalian cells, disruption of topoisomerase IIh by siRNA has no detrimental effects on cell cycle progression. In

Clin Cancer Res 2005;11(23) December 1, 2005

addition, topoisomerase IIh may compensate, albeit only partially, for topoisomerase IIa during chromatin condensation and cell segregation (9, 32). Despite similar catalytic activities of the topoisomerase II isoenzymes, cell survival has been correlated with the expression of topoisomerase IIa (12, 33). The role of topoisomerase II as drug targets for topoisomerase II poisons has been evaluated in several studies; and it has been suggested that topoisomerase IIa is the relevant target of topoisomerase II poisons. Disruption or mutation of topoisomerase IIa was associated with drug resistance (34 – 36). In particular, there are several reports suggesting topoisomerase IIa to be an important target for epirubicin, the drug used predominantly for this study; however, much less is known about the role of topoisomerase IIh as a drug target for epirubicin (17). There currently are several drugs approved for clinical use that predominantly target topoisomerase IIa, whereas the clinical utility of topoisomerase IIh – specific drugs is still under investigation (37). Here, we provide evidence that HDACi render topoisomerase IIh a relevant target and effector substrate for topoisomerase II poisons. We have shown that preexposure of tumor cells to HDACi led to histone acetylation and down-regulation of proteins essential for the maintenance of heterochromatin (20). The ensuing chromatin decondensation was associated with increased binding of topoisomerase II poisons to the DNA substrate (14, 20). The kinetics of the HDACi-induced chromatin decondensation suggested that a 48-hour preexposure was optimal for synergistic activity. We now show that exposure of tumor cells to HDACi significantly reduced the concentrations of topoisomerase II poisons required for growth inhibition, even in cells manipulated for target-specific resistance to topoisomerase II – targeting agents. Sensitization was not tissue specific but was observed in breast, colon, leukemia, melanoma, and sarcoma cell lines. We observed a time-dependent reduction in the expression of topoisomerase IIa mRNA and protein in the presence of HDACi that was maximal at 48 hours. In contrast, the effects of HDACi on topoisomerase IIh were minimal. Modulation of the topoisomerase II isoforms with siRNA and antisense showed that depletion of topoisomerase IIh abrogated the HDACi-induced potentiation of topoisomerase II poisons, whereas the depletion of topoisomerase IIa did not affect this synergistic interaction. Furthermore, topoisomerase II poison cytotoxicity in the presence of HDACi was associated with an accumulation of cleavable complexes involving topoisomerase IIh. The depletion of topoisomerase IIa and increase in topoisomerase IIh – containing cleavable complexes provide a strong argument for a more prominent role of topoisomerase IIh in the synergistic interaction between HDACi and topoisomerase II – targeting agents. Combinations of HDACi and topoisomerase II – targeting agents have been reported by several investigators. Although Johnson et al. observed a decrease in sensitivity of leukemia cells to etoposide after preexposure to an HDACi (38), Tsai et al. observed a sensitization of leukemia cells to etoposide (18). This discrepancy is most like explained by differences in HDACi preexposure and drug concentration. Where Johnson et al. preexposed HL-60 cells to 100 nmol/L trichostatin A for 0.5 hours, Tsai et al. used 400 nmol/L trichostatin A for 4 hours. Since then, other groups have reported increased

8472

www.aacrjournals.org

Interaction of HDAC and Topoisomerase II Inhibitors

sensitivity of several tumor cell lines to topoisomerase II targeting agents after preexposure to an HDACi (14, 19, 20, 39, 40). Synergy was dependent on sequence and timing of drug administration and associated with HDACi-induced chromatin decondensation (14, 19). Although topoisomerase IIa has been proposed as the effector molecule for cytotoxicity for topoisomerase II poison, the specific roles of topoisomerase IIa and topoisomerase IIh in the interaction between HDACi and topoisomerase II poisons have not been defined.

Fraser et al. reported an initial activation of the topoisomerase IIa gene promoter with a transient up-regulation of topoisomerase IIa protein expression in leukemia cells treated with the HDACi NaB for 18 to 24 hours. This promoter activation was reversed with longer HDACi exposure, resulting in topoisomerase IIa promoter activity repression (41). In contrast, Kim et al. found no change in the expression or activity of topoisomerase IIa in glioblastoma cells exposed to the HDACi SAHA (19); however, the exposure times may have differed. We observed an HDACi-induced depletion of

Fig. 4. Topoisomerase IIh (topo IIb) is the effector protein. A, trapped-in-agarose DNA immunostaining analysis of cleavable complexes involving topoisomerase IIa (red) and topoisomerase IIh (green) in cells treated with 0.5 Amol/L epirubicin (Epi) for 4 hours in the presence and absence of a 48-hour pretreatment with 2 mmol/L valproic acid (VPA) by confocal microscopy. B, immunofluorescence analysis of square pixel surface area per replicate sample depicting changes in topoisomerase IIa and topoisomerase IIh cleavable complexes for the indicated treatment groups. Squares, mean; square brackets, 95% confidence interval. *, P V 0.001. C, band depletion assay of topoisomerase IIa and topoisomerase IIh in MCF-7 cells incubated with 50 Amol/L VM-26 after pretreatment with 0 or 2 mmol/L valproic acid and densitometry analysis relative to untreated cells. Topoisomerase IIa and topoisomerase IIh were depleted in the presence of VM-26, suggesting the formation of cleavable complexes. Topoisomerase IIh was further depleted when cells were exposed to valproic acid beforeVM-26, suggesting an enhancement of cleavable complex formation. The decrease in topoisomerase IIa induced by valproic acid was not due to the formation of cleavable complexes but due to down-regulation of protein expression (see Fig. 2B and Fig. 3B). Therefore, the depletion of topoisomerase IIa in the VPA:VM-26 sample was a reflection of protein down-regulation by valproic acid and cleavable complex formation induced byVM-26 alone and did not signify an enhancement of VM-26-induced cleavable complex formation.

www.aacrjournals.org

8473

Clin Cancer Res 2005;11(23) December 1, 2005

Cancer Therapy: Preclinical

topoisomerase IIa mRNA and protein. Maximal effects only occurred after a prolonged exposure to the HDACi (48 hours). Topoisomerase IIa depletion was not limited to a specific HDACi class but was observed with both short-chain fatty acid and hydroxamic acid HDACi. In part, the differences between the findings presented in this report and those presented by others may be explained by differences in the duration of HDACi exposure. More importantly, these findings may also suggest a limited role of topoisomerase IIa in the observed synergy between an HDACi and a topoisomerase II poison. This is further supported by our findings showing that nearcomplete depletion of topoisomerase IIa by siRNA or antisense in the presence of an HDACi did not inhibit the potentiation of topoisomerase II poisons by HDACi. The HDACi-induced depletion of topoisomerase IIa mRNA and protein expression was maximal at 48 hours, which correlated to maximal HDACiinduced chromatin decondensation and the optimal timing of synergy between HDACi and topoisomerase II – targeting agents. Although topoisomerase IIa was reported to be involved in chromatin condensation (42), a mechanistic link between the topoisomerase IIa depletion and chromatin decondensation cannot be established by the presented data. More evidence suggesting topoisomerase IIh as a relevant target in the potentiation of topoisomerase II – targeting agents by HDACi emerges from studies using the topoisomerase IIh – specific poison XK469. XK469 was shown to induce cleavable complexes in several cell lines in vitro; however, meaningful effects on tumor burden in vivo were not achievable at tolerable doses (43). We found that XK469 by itself only caused minimal apoptosis in breast cancer cells; however, when administered in a sequence-specific combination with HDACi, we observed a potentiation of XK469 induced apoptosis by HDACi. These experiments collectively imply that topoisomerase IIh is a relevant target in the interaction between HDACi and

topoisomerase II – targeting agents. This is further supported by the findings that depletion of topoisomerase IIh abrogated the synergistic interaction, whereas depletion of topoisomerase IIa did not affect synergy. Although topoisomerase IIh may be a target of selected topoisomerase II poisons (44 – 46), most reports suggest that the cytotoxicity for the clinically relevant topoisomerase II – targeting agents is determined by topoisomerase IIa levels (47 – 50). The data presented here show that HDACi potentiates the cytotoxicity of topoisomerase IIa and topoisomerase IIh – specific topoisomerase II poisons by recruiting topoisomerase IIh as a target. The results of this study may have several clinical implications. HDACi may lower the concentration of topoisomerase II targeting agents required for activity, thereby limiting adverse effects. Emerging drug resistance due to a loss or mutation of topoisomerase IIa may be circumvented by engaging topoisomerase IIh as an alternative target. A potentiation of cytotoxicity was not observed in fibroblast exposed to an HDACi before exposure to a topoisomerase II poison (data not shown). Furthermore, preliminary results from an ongoing phase I trial studying a combination of valproic acid and epirubicin did not suggest a potentiation of epirubicin-induced toxicity (Munster et al, ASCO 2005 #A3084). The differential effects of this combination on somatic versus tumor cells may limit tissue-specific toxicities, such as anthracycline-associated cardiotoxicity. In addition, HDACi may increase the clinical utility of h-specific drugs.

Acknowledgments We thank Dr. Fred Hausheer (Bionumerick Pharmaceuticals Inc., San Antonio, TX) for the MCF:Dox cell line and Dr. Robert Snapka (Department of Radiology, Ohio State University, Columbus, OH) for XK469.

References 1. Lang AJ, Mirski SE, Cummings HJ, et al. Structural organization of the human TOP2A and TOP2B genes. Gene 1998;221:255 ^ 66. 2. Drake FH, Hofmann GA, Bartus HF, et al. Biochemical and pharmacological properties of p170 and p180 forms of topoisomerase II. Biochemistry 1989;28: 8154 ^ 60. 3. Larsen AK, Skladanowski A, Bojanowski K.The roles of DNA topoisomerase II during the cell cycle. Prog Cell Cycle Res 1996;2:229 ^ 39. 4. Burden DA, Osheroff N. Mechanism of action of eukaryotic topoisomerase II and drugs targeted to the enzyme. Biochim Biophys Acta 1998;1400:139 ^ 54. 5. Heck MM, Hittelman WN, Earnshaw WC. Differential expression of DNA topoisomerases I and II during the eukaryotic cell cycle. Proc Natl Acad Sci U S A 1988; 85:1086 ^ 90. 6. Prosperi E, Sala E, Negri C, et al. Topoisomerase II a and h in human tumor cells grown in vitro and in vivo. Anticancer Res 1992;12:2093 ^ 9. 7. Woessner RD, Mattern MR, Mirabelli CK, Johnson RK, Drake FH. Proliferation- and cell cycle-dependent differences in expression of the 170 kilodalton and 180 kilodalton forms of topoisomerase II in NIH-3T3 cells. Cell Growth Differ 1991;2:209 ^ 14. 8. Xiao H, MaoY, Desai SD, et al. The topoisomerase IIh circular clamp arrests transcription and signals a 26S proteasome pathway. Proc Natl Acad Sci U S A 2003;100:3239 ^ 44. 9. Sakaguchi A, Kikuchi A. Functional compatibility between isoform a and h of type II DNA topoisomerase. J Cell Sci 2004;117:1047 ^ 54.

10. Berger JM, Gamblin SJ, Harrison SC, Wang JC. Structure and mechanism of DNA topoisomerase II. Nature 1996;379:225 ^ 32. 11. Capranico G,Tinelli S, Zunino F. Formation, resealing and persistence of DNA breaks produced by 4-demethoxydaunorubicin in P388 leukemia cells. Chem Biol Interact 1989;72:113 ^ 23. 12. PommierY, Pourquier P, FanY, Strumberg D. Mechanism of action of eukaryotic DNA topoisomerase I and drugs targeted to the enzyme. Biochim Biophys Acta 1998;1400:83 ^ 105. 13. JarvinenTA, Tanner M, Rantanen V, et al. Amplification and deletion of topoisomerase IIa associate with ErbB- 2 amplification and affect sensitivity to topoisomerase II inhibitor doxorubicin in breast cancer. Am J Pathol 2000;156:839 ^ 47. 14. Marchion DC, Bicaku E, Daud AI, et al. Sequencespecific potentiation of topoisomerase II inhibitors by the histone deacetylase inhibitor suberoylanilide hydroxamic acid. J Cell Biochem 2004;92:223 ^ 37. 15. Gudkov AV, Zelnick CR, Kazarov AR, et al. Isolation of genetic suppressor elements, inducing resistance to topoisomerase II-interactive cytotoxic drugs, from human topoisomerase II cDNA. Proc Natl Acad Sci U S A 1993;90:3231 ^ 5. 16. Asano T, An T, Mayes J, Zwelling LA, Kleinerman ES. Transfection of human topoisomerase II a into etoposide-resistant cells: transient increase in sensitivity followed by down-regulation of the endogenous gene. Biochem J 1996;319:307 ^ 13. 17. MacGrogan G, Rudolph P, Mascarel Id I, et al. DNA topoisomerase IIa expression and the response topri-

Clin Cancer Res 2005;11(23) December 1, 2005

8474

mary chemotherapy in breast cancer. Br J Cancer 2003;89:666 ^ 71. 18. Tsai SC, Valkov N, Yang WM, et al. Histone deacetylase interacts directly with DNA topoisomerase II. Nat Genet 2000;26:349 ^ 53. 19. Kim MS, Blake M, BaekJH, et al. Inhibition of histone deacetylase increases cytotoxicity to anticancer drugs targeting DNA. Cancer Res 2003;63:7291 ^ 300. 20. Marchion DC, Bicaku E, Daud AI, Sullivan DM, Munster PN. Valproic acid alters chromatin structure by regulation of chromatin modulation proteins. Cancer Res 2005;65:3815 ^ 22. 21. Towatari M, Adachi K, Marunouchi T, Saito H. Evidence for a critical role of DNA topoisomerase IIa in drug sensitivity revealed by inducible antisense RNA in a human leukaemia cell line. Br J Haematol 1998;101:548 ^ 51. 22. Willmore E, Frank AJ, Padget K, Tilby MJ, Austin CA. Etoposide targets topoisomerase IIa and IIh in leukemic cells: isoform-specific cleavable complexes visualized and quantified in situ by a novel immunofluorescence technique. Mol Pharmacol 1998;54: 78 ^ 85. 23. Munster PN, Troso-Sandoval T, Rosen N, et al. The histone deacetylase inhibitor suberoylanilide hydroxamic acid induces differentiation of human breast cancer cells. Cancer Res 2001;61:8492 ^ 7. 24. Qiu J, Catapano CV, Fernandes DJ. Formation of topoisomerase II a complexes with nascent DNA is related to VM-26-induced cytotoxicity. Biochemistry 1996;35:16354 ^ 60. 25. Christensen MO, Larsen MK, Barthelmes HU, et al.

www.aacrjournals.org

Interaction of HDAC and Topoisomerase II Inhibitors Dynamics of human DNA topoisomerases IIa and IIh in living cells. J Cell Biol 2002;157:31 ^ 44. 26. Null AP, Hudson J, Gorbsky GJ. Both a and h isoforms of mammalian DNA topoisomerase II associate with chromosomes in mitosis. Cell Growth Differ 2002;13:325 ^ 33. 27. Marsh KL, Willmore E, Tinelli S, et al. Amsacrinepromoted DNA cleavage site determinants for the two human DNA topoisomerase II isoforms a and h. Biochem Pharmacol 1996;52:1675 ^ 85. 28. Austin CA, Marsh KL. Eukaryotic DNA topoisomerase II h. Bioessays 1998;20:215 ^ 26. 29. Jensen S, Redwood CS, Jenkins JR, Andersen AH, Hickson ID. Human DNA topoisomerases II a and II h can functionally substitute for yeast TOP2 in chromosome segregation and recombination. Mol Gen Genet 1996;252:79 ^ 86. 30. Yang X, Li W, Prescott ED, Burden SJ, Wang JC. DNA topoisomerase IIh and neural development. Science 2000;287:131 ^ 4. 31. Akimitsu N, Kamura K, Tone S, et al. Induction of apoptosis by depletion of DNA topoisomerase IIa in mammalian cells. Biochem Biophys Res Commun 2003;307:301 ^ 7. 32. Grue P, Grasser A, Sehested M, et al. Essential mitotic functions of DNA topoisomerase IIa are not adopted by topoisomerase IIh in human H69 cells. J Biol Chem 1998;273:33660 ^ 6. 33. Corbett AH, Osheroff N. When good enzymes go bad: conversion of topoisomerase II to a cellular toxin by antineoplastic drugs. Chem Res Toxicol 1993;6: 585 ^ 97. 34. Jensen LH,Wessel I, Moller M, et al. N-terminal and core-domain random mutations in human topoisomerase II a conferring bisdioxopiperazine resistance. FEBS Lett 2000;480:201 ^ 7.

www.aacrjournals.org

35. Matsumoto Y, Takano H, Kunishio K, Nagao S, FojoT. Incidence of mutation and deletion in topoisomerase II a mRNA of etoposide and mAMSA-resistant cell lines. Jpn J Cancer Res 2001;92:1133 ^ 7. 3 6. Kruczynski A, Barret JM, Van Hille B, et al. Decreased nucleotide excision repair activity and alterations of topoisomerase IIa are associated with the in vivo resistance of a P388 leukemia subline to F11782, a novel catalytic inhibitor of topoisomerases I and II. Clin Cancer Res 2004;10:3156 ^ 68. 37. Kellner U, Rudolph P, Parwaresch R. Human DNAtopoisomerases: diagnostic and therapeutic implications for cancer. Onkologie 2000;23:424 ^ 30. 38. Johnson CA, Padget K, Austin CA, Turner BM. Deacetylase activity associates with topoisomerase II and is necessary for etoposide-induced apoptosis. J Biol Chem 2001;276:4539 ^ 42. 39. Niitsu N, KasukabeT,Yokoyama A, et al. Anticancer derivative of butyric acid (Pivalyloxymethyl butyrate) specifically potentiates the cytotoxicity of doxorubicin and daunorubicin through the suppression of microsomal glycosidic activity. Mol Pharmacol 2000;58: 27 ^ 36. 40. Kurz EU, Wilson SE, Leader KB, et al. The histone deacetylase inhibitor sodium butyrate induces DNA topoisomerase II a expression and confers hypersensitivity to etoposide in human leukemic cell lines. Molecular CancerTherapeutics 2002;1:121 ^ 31. 41. Fraser DJ, Brandt TL, Kroll DJ. Topoisomerase II a promoter trans-activation early in monocytic differentiation of HL-60 human leukemia cells. Mol Pharmacol 1995;47:696 ^ 706. 42. Cuvier O, HiranoT. A role of topoisomerase II in linking DNA replication to chromosome condensation. J Cell Biol 2003;160:645 ^ 55. 43. Mensah-Osman EJ, Al-Katib AM, Dandashi MH,

8475

Mohammad RM. 2-[4-(7-chloro-2-quinoxalinyloxy)phenoxy]-propionic acid (XK469) inhibition of topoisomerase IIh is not sufficient for therapeutic response in humanWaldenstrom’s macroglobulinemia xenograft model. Mol CancerTher 2002;1:1315 ^ 20. 44. Errington F,Willmore E,Tilby MJ, et al. Murine transgenic cells lacking DNA topoisomerase IIh are resistant to acridines and mitoxantrone: analysis of cytotoxicity and cleavable complex formation. Mol Pharmacol 1999;56:1309 ^ 16. 45. Harker WG, Slade DL, Drake FH, Parr RL. Mitoxantrone resistance in HL-60 leukemia cells: reduced nuclear topoisomerase II catalytic activity and druginduced DNA cleavage in association with reduced expression of the topoisomerase II h isoform. Biochemistry 1991;30:9953 ^ 61. 46. Gao H, Huang KC, Yamasaki EF, et al. XK469, a selective topoisomerase IIh poison. Proc Natl Acad Sci U S A 1999;96:12168 ^ 73. 47. Feldhoff PW, Mirski SE, Cole SP, Sullivan DM. Altered subcellular distribution of topoisomerase II a in a drug-resistant human small cell lung cancer cell line. Cancer Res 1994;54:756 ^ 62. 48. BeckWT, Danks MK, Wolverton JS, Kim R, Chen M. Drug resistance associated with altered DNA topoisomerase II. Adv Enzyme Regul 1993;33:113 ^ 27. 49. Nitiss JL, Liu YX, Hsiung Y. A temperature sensitive topoisomerase II allele confers temperature dependent drug resistance on amsacrine and etoposide: a genetic system for determining the targets of topoisomerase II inhibitors. Cancer Res 1993;53:89 ^ 93. 50. Kobayashi M, Adachi N, Aratani Y, Kikuchi A, Koyama H. Decreased topoisomerase IIa expression confers increased resistance to ICRF-193 as well as VP-16 in mouse embryonic stem cells. Cancer Lett 2001;166:71 ^ 7.

Clin Cancer Res 2005;11(23) December 1, 2005