Under conditions of genotoxic stress, cancer cells strongly rely on efficient DNA repair to survive and proliferate. The human BRCA2 tumor suppressor protein is indispensable for the repair of DNA double-strand breaks by homologous recombination (HR) by virtue of its ability to promote RAD51 loading onto single-stranded DNA. Therefore, blocking the interaction between BRCA2 and RAD51 could significantly improve the efficacy of conventional anticancer therapies. However, targeting protein–protein interaction (PPI) interfaces has proven challenging because flat and large PPI surfaces generally do not support binding of small-molecule inhibitors. In contrast, peptides are more potent for targeting PPIs but are otherwise difficult to deliver into cells. Here, we report that a synthetic 16-mer peptide derived from the BRC4 repeat motif of BRCA2 is capable of blocking RAD51 binding to BRCA2. Efficient noncytotoxic cellular uptake of a nona-arginine (R9)-conjugated version of the BRC4 peptide interferes with DNA damage–induced RAD51 foci formation and HR. Moreover, transduction of the BRC4 peptide impairs replication fork–protective function of BRCA2 and triggers MRE11-dependent degradation of nascent DNA in response to DNA replication stress. Finally, the BRC4 cell-penetrating peptide (CPP) confers selective hypersensitivity to PARP inhibition in cancer cells but spares noncancerous cells. Taken together, our data highlight an innovative approach to develop novel peptide-based DNA repair inhibitors and establish BRCA2-derived CPPs as promising anticancer agents. Mol Cancer Ther; 17(7); 1392–404. ©2018 AACR.

Double-strand breaks (DSB) are highly detrimental DNA lesions because, if left unrepaired or misrepaired, they can trigger cell death and genomic instability, ultimately causing cancer (1). To circumvent this threat, cells are equipped with diverse DSB repair mechanisms, including nonhomologous end joining (NHEJ) and homologous recombination (HR) as the two major pathways (2). Furthermore, recent work has established that in response to DNA replication stress, several key HR factors play a crucial role in protecting stalled DNA replication forks from nucleolytic degradation (3). Because rapidly dividing cancer cells rely on efficient DSB repair and fork protection mechanisms for their survival, inhibiting HR represents an attractive strategy for the development of novel therapeutic drugs, in particular when used in combination with DNA-damaging agents (4, 5).

The human BRCA2 protein plays an essential role in HR by promoting homology search and stimulating strand invasion into the sister chromatid (6). Specifically, following DNA-end resection, BRCA2 directs RAD51 filament nucleation onto RPA-coated single-stranded DNA (ssDNA; ref. 7). RAD51 interacts with two distinct regions in BRCA2, the BRC repeat motifs and a C-terminal domain (8–10). Importantly, the eight evolutionarily conserved BRC repeats, each consisting of about 35 amino acids, significantly differ in their capacity to bind RAD51 with BRC4 displaying the highest affinity (11, 12). Consequently, it was proposed that BRC repeats 1–4 facilitate nucleation of RAD51 by binding monomeric RAD51 and reducing its ATPase activity (11, 13). Structural analysis of the BRC4 repeat identified residues 1523-GFHTASG-1529 of BRCA2 to structurally mimic the self-oligomerization motif of RAD51 (14). In addition to the FHTA motif, a second consensus tetrameric module in BRC4, denoted as LFDE motif, was shown to bind to a distinct pocket in RAD51 distant from the oligomerization interface (15). In contrast to the BRC repeats, the C-terminal domain does not bind monomeric RAD51 but instead stabilizes RAD51 nucleoprotein filaments (9, 10). Taken together, compounds that selectively and efficiently block BRCA2–RAD51 interaction could advance into the clinic as bona fide HR inhibitors for both monotherapy and add-on therapy with DNA-damaging agents.

The physical nature of protein–protein interaction (PPI) interfaces often renders them unable to support binding of small-molecule inhibitors (SMI). Instead, peptide therapeutics offer an alternative way to target PPIs with key advantages over SMIs, including their direct similarity to protein fragments and the coverage of extensive PPI interfaces (16). However, poor membrane permeability has previously limited their use to extracellular targets (17). Thus, hydrophilic peptides are reliant on a permeation-enhancing strategy that facilitates targeting of intracellular molecules (18). Recently, cell-penetrating peptides (CPP) have been developed to enhance the cellular uptake and nuclear translocation of membrane-impermeable cargo molecules (19). They comprise a highly diverse class of short, primarily cationic peptides that combine a limited cytotoxicity and the ability to mediate receptor-independent transport of cargoes across cell membranes (20). Notably, the nona-arginine (R9) peptide is one of the most potent CPPs, giving a high transduction efficiency combined with low cytotoxicity (21).

Here, we design a CPP comprised of a 16-amino acid stretch of the BRCA2 BRC4 repeat able to inhibit BRCA2–RAD51 interaction. Our detailed functional analysis reveals that an R9-fused BRC4 CPP prevents RAD51 loading onto ssDNA, resulting in defective homology-mediated repair of DSBs as well as increased MRE11-dependent degradation of stalled DNA replication forks. Consequently, peptide incubation renders cells hypersensitive to the PARP inhibitor olaparib, providing a potential use for BRCA2-derived peptides in the treatment of certain types of cancer.

Cell culture

HeLa, U2OS, RPE1, MRC5 (all from ATCC), and HeLa DR-GFP were cultured in DMEM (Gibco) supplemented with 10% FCS (Sigma-Aldrich), 100 U/mL penicillin, and 100 μg/mL streptomycin (Life Technologies). PEO1 and PEO4 cells were purchased from the Health Protection Agency Culture Collections and cultured in RPMI medium (Gibco) supplemented with 10% FCS, 2 mmol/L sodium pyruvate (Gibco), and penicillin/streptomycin. MCF10A cells were purchased from ATCC and cultured in DMEM/F12 (Gibco) containing 5% Horse Serum (Gibco), 20 ng/mL human EGF (Sigma-Aldrich), 0.5 mg/mL hydrocortisone (Sigma-Aldrich), 10 μg/mL insulin, and penicillin/streptomycin. Stable U2OS cells expressing GFP-RAD51 (22) were grown in DMEM supplemented with 10% Tet system approved FCS (GIBCO) and penicillin/streptomycin. To induce GFP-RAD51 expression, cells were treated with 1 μg/mL doxycycline (Sigma-Aldrich) for 24 hours. All cell lines were confirmed to be free of mycoplasma contamination on a regular basis (PCR Mycoplasma Test Kit, AppliChem). Cells were passaged for no longer than 2 months after thawing of early-passage stocks. For cells that have been received from secondary sources, no cell line authentication was performed. Irradiation was performed using a Faxitron X-ray machine.

Chemicals and peptides

Camptothecin, RAD51 inhibitor B02 (23), cycloheximide, hydroxyurea (HU), and mirin (24) were purchased from Sigma-Aldrich. Olaparib (AZD2281) was provided by Selleck Chemicals. Thymidine analogs CIdU, IdU, and EdU were purchased from Sigma-Aldrich and Life Technologies, respectively. Custom-designed peptides were purchased from Bachem AG and, if not specified, synthesized according to standard practice (l-amino acids, N-terminal tag or acetylation, C-terminal amidation). Lyophilized peptides were dissolved in PBS at 1 mg/mL.

Antibodies

A detailed list of all primary and secondary antibodies can be found in Supplementary Tables S1 and S2, respectively.

siRNA

A detailed list of siRNA oligonucleotide sequences used in this study can be found in Supplementary Table S3. siRNA oligos were used at a final concentration of 10 nmol/L and transfected using Lipofectamine RNAiMAX (Invitrogen) according to the manufacturer's instructions.

Recombinant protein expression

BRCA2 GST-fusion plasmids (GST-BRC 1-2, GST-BRC3-5, GST-BRC6-8, GST-C-term) have been described before (25). BRCA2 GST-fusion proteins were expressed in BL21-CodonPlus-RIL Escherichia coli (E. coli) by growing them overnight at 18°C using 100 μmol/L isopropyl β-d-thiogalactoside. Recombinant full-length RAD51 was prepared as described previously (26).

Immunoblotting

If not specified otherwise, cells were lysed in Laemmli buffer (4% SDS, 20% glycerol, 120 mmol/L Tris-HCl pH 6.8) and resolved by Tris-glycine SDS-PAGE. To probe for BRCA2, 3%–6% NuPAGE Tris-Acetate gels (Thermo Fisher Scientific) were run according to the manufacturer's instructions. After transfer to nitrocellulose membranes, immunoblotting was performed with indicated primary antibodies overnight at 4°C and secondary antibodies for 1 hour at room temperature. Stained proteins were visualized using the Advansta WesternBright ECL reagent and the VilberLourmat Fusion Solo S imaging system.

Pull-down assays

For peptide pull-downs, 30 μL streptavidin-coupled Dynabeads (Life Technologies) were incubated with 5 μg (2.7 nmol) of biotinylated BRC4 peptides or biotin analogue d-desthiobiotin (Sigma-Aldrich) in 1 mL PBS-T (0.1% Triton X-100) for 1 hour at 4°C. Beads were washed three times with PBS-T and blocked for 30 minutes with 0.3% BSA in PBS at 4°C. A total of 50 ng (1.35 pmol) recombinant RAD51 together with 2 μmol/L ATP was added to the beads and incubated for 2 hours in 700 μL PBS-T. The beads were washed four times with NTEN300 (20 mmol/L Tris pH 7.4, 0.1 mmol/L EDTA, 300 mmol/L NaCl, 0.5% NP-40) and once with TEN100 (20 mmol/L Tris pH 7.4, 0.1 mmol/L EDTA, 100 mmol/L NaCl) before complexes were boiled in 2× SDS sample buffer (10 mmol/L Tris pH 6.8, 20% glycerol, 3% SDS, 200 mmol/L DTT, 0.04% bromophenol blue) and subjected to immunoblotting. For GST pull-down assays, glutathione sepharose beads (GE Healthcare) were incubated for 1 hour at 4°C with equalized amounts of BL21 E. coli soluble extracts expressing one of the four GST-BRCA2 fusion constructs in TEN100. Beads were washed three times with NTEN300 buffer and once with TEN100 before adding either 1 mg HeLa nuclear extracts (CilBiotech) or 50 ng purified RAD51 supplemented with varying amounts of BRC4wt or BRC4mut peptides filled up to 1 mL with TEN100. After 2 hours of incubation, beads were washed twice with NTEN500 (20 mmol/L Tris pH 7.4, 0.1 mmol/L EDTA, 500 mmol/L NaCl, 0.5% NP-40), twice with NTEN300, and twice with TEN100 buffer before boiling in SDS sample buffer and protein analysis by immunoblotting.

Coimmunoprecipitation

Cell extracts were prepared using NP-40 extraction buffer [50 mmol/L Tris-HCl pH 7.5, 120 mmol/L NaCl, 1 mmol/L EDTA, 6 mmol/L EGTA, 15 mmol/L sodium pyrophosphate and 1% NP-40 supplemented with phosphatase inhibitors (20 mmol/L NaF, 1 mmol/L sodium orthovanadate), and protease inhibitors (1 mmol/L benzamidine and 0.1 mmol/L PMSF)]. After benzonase (Novagen) digestion for 30 minutes at 4°C, cell extracts were cleared by centrifugation. Lysates (2 mg) were supplemented with increasing amounts of peptides filled up to 1 mL with NP-40 extraction buffer and incubated for 1 hour at 4°C before adding 20 μL GFP-Trap agarose beads (ChromoTek) for 1 hour at 4°C. Beads were subsequently washed three times with GFP-IP buffer (100 mmol/L NaCl, 0.2% NP-40, 1 mmol/L MgCl2, 10% glycerol, 5 mmol/L NaF, 50 mmol/L Tris-HCl pH 7.5) and boiled in SDS sample buffer for analysis by immunoblotting.

Peptide transfection

Cells were seeded either into 8-well chamber imaging slides (μ-Slide 8 Well, ibidi), 24-well plates, 6-well plates, or 6-cm culture dishes (Sarstedt) and grown to around 80% confluence at day of peptide transfection. Cells were washed at least once with PBS to remove residual FCS and incubated with indicated peptide concentrations in appropriate serum-free medium for 1 hour at 37°C. If not specified otherwise, the following incubation volumes were used: 0.3 mL for 8-well chamber imaging slides and 24-well plates, 0.5 mL for 12-well plates and 2 mL for 6-well plates and 6-cm culture dishes.

Confocal microscopy

A total of 4 × 104 cells were seeded into 8-well chamber imaging slides and grown overnight. After 30 minutes of staining with 0.5 μg/mL Hoechst 33342 (Life Technologies), cells were washed twice with PBS and incubated with indicated peptide concentrations. Cells were washed twice with PBS and imaged in Live Cell Imaging solution (Thermo Fisher Scientific). Images were taken with CLSM SP5 Mid UV-VIS Leica with 63× objective at 37°C at ambient CO2 concentrations.

Flow cytometry

EdU incorporation was analyzed using the Click-it EdU technology (Thermo Fisher Scientific) according to the manufacturer's instructions. For peptide uptake studies, 1 × 105 cells were seeded into 12-well plates. The day after, peptide transfection was performed and cells were harvested by trypsinization to remove membrane-bound peptides. After one wash with PBS, cells were resuspended in PBS and subjected to flow cytometry analysis. To quantify intracellular peptide stability, same cells were released for indicated time points in DMEM + 10% FCS and fixed with 4% formaldehyde (w/v) in PBS for 15 minutes at room temperature. To measure the TAMRA fluorescence intensity, the LSR II Fortessa equipped with a 561 nm laser line and a 586/15 band-pass filter was used. Of note, the fluorescence intensity of TAMRA-labeled R9-BRC4mut peptides was corrected for quenching by multiplying measured TAMRA intensity with a quenching factor. The quenching factor was calculated by loading 10 pmol of freshly solubilized peptides on Tricine SDS-PAGE gel and quantifying TAMRA intensity (see Fig. 2C, lane 5). Fluorescein intensity was measured with Attune Nxt Flow Cytometer equipped with 488 laser and 530/30 band-pass filter. For each condition, 20,000 events were recorded. MACS Quant Calibration Beads (MACS Miltenyi Biotec) were applied for voltage standardization to exclude any machine-dependent variations between measurements.

Tricine SDS-PAGE

To resolve low molecular weight peptides, Tricine SDS-PAGE was performed as described previously (27). For peptide separation, Laemmli lysates were loaded onto 16% Tricine SDS-PAGE gel containing 6 mol/L urea. For peptide detection via fluorescence, gels were scanned using a Typhoon FLA 9500 FluorImager.

HR reporter assay

HR frequency was measured as described previously (28, 29). Briefly, following siRNA transfection, 1 × 105 HeLa cells containing a stably integrated DR-GFP reporter construct were seeded into 12-well plates. The day after, cells were either mock-transfected or transfected with 0.6 μg I-SceI expression plasmid (pCBASce) using jetPrime transfection reagent (Polyplus transfection). Four hours later, medium was exchanged and either a 1-hour peptide incubation or second siRNA transfection was performed. Peptide incubations were repeated 24 and 34 hours post–I-SceI transfection. After each peptide incubation, 0.5 mL of DMEM + 20% FCS was directly added to the peptide/DMEM mix. Forty-eight hours after I-SceI transfection, cells were harvested and directly analyzed for GFP expression by flow cytometry using an Attune Nxt Flow Cytometer.

Immunofluorescence microscopy

Twenty-four hours after siRNA transfection, 8 × 104 cells were seeded on coverslips in 24-well plates. The day after, cells were treated either with 100 nmol/L camptothecin for 1 hour or irradiated and incubated for another hour with the peptides, before releasing them for 3 hours by directly adding 1 mL of DMEM + 14% FCS. Alternatively, cells were transfected with the peptides, followed by 1-hour camptothecin treatment and direct processing. Cells were preextracted for 5 minutes on ice (25 mmol/L HEPES pH 7.4, 50 mmol/L NaCl, 1 mmol/L EDTA, 3 mmol/L MgCl2, 300 mmol/L sucrose, 0.5% Triton X-100), fixed with 4% formaldehyde (w/v) in PBS for 15 minutes at room temperature, before incubating them with indicated primary and appropriate secondary antibodies for 1 hour. Afterwards, coverslips were mounted with Vectashield (Vector Laboratories) containing DAPI and sealed. Images were acquired on Leica DMI6000 widefield fluorescence microscope with a 63× objective.

DNA fiber analysis

DNA fiber analyses were carried out as described previously (30). In brief, U2OS cells were seeded into 6-well plates at a confluence of 40%. Twenty-four hours later, cells were pulse-labeled with 33 μmol/L CIdU for 30 minutes, followed by 340 μmol/L IdU for 30 minutes prior to incubation with 2 mmol/L HU and peptides (10 μmol/L) for 4 hours in serum-free medium. Cells were lysed (200 mmol/L Tris-HCl pH 7.4, 50 mmol/L EDTA, 0.5% SDS) and DNA fibers were stretched onto glass slides before fixation in methanol–acetic acid (3:1, Merck) overnight. Rehydration in PBS was followed by denaturation in 2.5 mol/L HCl for 1 hour, a PBS wash, and blockage (2% BSA (w/v) PBS, 0.1% Tween 20) for 40 minutes. CIdU and IdU staining was performed using anti-BrdU primary and secondary antibodies for 2.5 hours. Coverslips were mounted using Antifade Gold (Invitrogen). Images were acquired on Olympus microscope IX81 with ×60 magnification, and analysis was carried out using ImageJ software.

Colony formation assay

Indicated cell lines were plated in poly-l-lysine (Sigma-Aldrich) coated 24-well plates at low cell dilutions of 200 cells/well in technical triplicates. PEO1 and PEO4 cells were seeded at 500 and 1,000 cells/well, respectively. Twenty-four hours later, cells were washed once with PBS and incubated with olaparib in presence or absence of peptides. After 1 hour, 1 mL of appropriate medium containing 14% FCS with indicated olaparib concentrations was directly added to the cells without removing the peptide solution. For MCF10A cells, FCS concentration of culture medium was increased to 7%. Alternatively, HeLa cells were treated for 1 hour with 1 μmol/L camptothecin, washed twice with PBS, and peptide transfection was carried out for 1 hour before directly adding 1 mL of DMEM + 14% FCS. Cells were grown for 10 days before fixation with crystal violet solution [0.5 % crystal violet, 20% ethanol (w/v)]. For analysis, plates were scanned and analyzed with the ImageJ Plugin ColonyArea using the parameter colony intensity, integrating the percentage of the covered area and staining intensity (31).

Statistical analysis

All results were confirmed in at least two independent experiments. Quantitative data are displayed as mean ± SD, and statistical analyses were performed using GraphPad Prism 7. P values <0.05 were considered statistically significant.

BRC4 peptide inhibits BRCA2–RAD51 interaction

It is well established that BRCA2 binds RAD51 through its BRC repeats composed of two highly conserved tetrameric motifs (Supplementary Fig. S1A). Among the eight BRC repeats, BRC4 was reported to display the highest affinity for RAD51, mainly using its FHTA sequence to bind the RAD51 oligomerization motif. To specifically target the BRCA2–RAD51 protein interaction interface, we therefore synthesized a 16-mer peptide mimicking the N-terminal half of BRC4 comprising the FHTA hydrophobic motif (Fig. 1A). In addition to the “wild type” BRC4 peptide (BRC4wt), we included a “mutated” BRC4 peptide (BRC4mut) harboring an inverted FHTA sequence (Fig. 1A). Employing N-terminally biotinylated peptides, recombinant RAD51 was efficiently pulled down by BRC4wt but to a much lesser extent by BRC4mut (Fig. 1B). In agreement with previous reports (32), we observed that GST-tagged BRCA2 fusion proteins spanning BRC 1-2, BRC 3-5, or the C-terminal (C-term) domain were able to interact with RAD51 (Supplementary Fig. S1B; Fig. 1C). Remarkably, BRC4wt was able to outcompete each of these individual interactions in a concentration-dependent manner (Fig. 1D). Interestingly, the BRC4 peptide was more effective in outcompeting RAD51 binding to the BRCA2 C-term and BRC repeats 1-2 than to BRC repeats 3-5 (Fig. 1D), indicating that the BRC4 repeat of BRCA2 exhibits the highest binding affinity for RAD51. Similar GST pull-down results were obtained using HeLa nuclear extracts as a source for RAD51 (Supplementary Fig. S1C). Given that the BRCA2 C-terminal region exclusively binds to assembled RAD51 oligomers, we reasoned that the BRC4 peptide is able to disrupt RAD51 multimers present in solution, which is in agreement with binding of the FHTA cluster to the RAD51 oligomerization motif (9, 10, 14). Most importantly, coimmunoprecipitation experiments in U2OS cells inducibly expressing GFP-tagged RAD51 demonstrated that BRC4wt, but not BRC4mut, is capable of interfering both with BRCA2 binding to RAD51 and RAD51 oligomerization (Fig. 1E; Supplementary Fig. S1D and S1E). Taken together, our results indicated that a short, synthetic BRC4-derived peptide is proficient in blocking BRCA2–RAD51 protein–protein interaction.

Figure 1.

A short BRC4 peptide inhibits BRCA2–RAD51 interaction. A, Amino acid sequences of BRCA2-derived 16-mer peptides used in this study. In addition to the wild-type BRC4 peptide (BRC4wt), a mutant version (BRC4mut) was included, characterized by an inverted FHTA tetrapeptide motif. B, Biotinylated BRC4 peptides coupled to streptavidin beads were used to perform pull-down assays using bacterially expressed and purified human RAD51. Inputs and pulled down protein complexes were analyzed by immunoblotting. SE and LE denote short and long exposure times of the same anti-RAD51 immunoblot. C, Different GST-BRCA2 fusion proteins coupled to glutathione–sepharose beads were used to perform pull-down assays using recombinant RAD51. Inputs and pulled down protein complexes were analyzed by immunoblotting. D, GST-BRCA2 pull-down assays with recombinant RAD51 in the absence or presence of increasing amounts of BRC4wt peptides. Inputs and pulled down protein complexes were analyzed by immunoblotting. E, Lysates of U2OS cells inducibly expressing GFP-RAD51 were supplemented with 50 μg of the indicated BRC4 peptides and subjected to immunoprecipitation (IP) using anti-GFP affinity resin. Inputs and coimmunoprecipitated protein complexes were analyzed by immunoblotting.

Figure 1.

A short BRC4 peptide inhibits BRCA2–RAD51 interaction. A, Amino acid sequences of BRCA2-derived 16-mer peptides used in this study. In addition to the wild-type BRC4 peptide (BRC4wt), a mutant version (BRC4mut) was included, characterized by an inverted FHTA tetrapeptide motif. B, Biotinylated BRC4 peptides coupled to streptavidin beads were used to perform pull-down assays using bacterially expressed and purified human RAD51. Inputs and pulled down protein complexes were analyzed by immunoblotting. SE and LE denote short and long exposure times of the same anti-RAD51 immunoblot. C, Different GST-BRCA2 fusion proteins coupled to glutathione–sepharose beads were used to perform pull-down assays using recombinant RAD51. Inputs and pulled down protein complexes were analyzed by immunoblotting. D, GST-BRCA2 pull-down assays with recombinant RAD51 in the absence or presence of increasing amounts of BRC4wt peptides. Inputs and pulled down protein complexes were analyzed by immunoblotting. E, Lysates of U2OS cells inducibly expressing GFP-RAD51 were supplemented with 50 μg of the indicated BRC4 peptides and subjected to immunoprecipitation (IP) using anti-GFP affinity resin. Inputs and coimmunoprecipitated protein complexes were analyzed by immunoblotting.

Close modal

The CPP R9 facilitates intracellular delivery of BRC4

Native peptides do not readily cross cell membranes. To enhance cellular uptake, we decided to conjugate BRC4 peptides with a nona-arginine (R9) CPP. In addition, a red fluorescent dye (TAMRA) was N-terminally attached to R9-BRC4 peptides to analyze cellular uptake. Using confocal microscopy, we observed robust cytoplasmic and nuclear TAMRA signals in HeLa and U2OS cells upon transfection with R9-fused peptides (Fig. 2A). Moreover, we found that the concentration threshold for efficient R9-BRC4 cell penetration was above 10 μmol/L (Supplementary Fig. S2A). Flow cytometry analyses further confirmed that BRC4 peptide delivery reached a transduction efficiency of almost 100% when fused to R9 (Fig. 2B; Supplementary Fig. S2B). Importantly, when replacing the TAMRA label with a green fluorescent dye (Fluorescein), we observed very similar subcellular localization patterns and fluorescent intensities of the R9-BRC4 peptides (Supplementary Fig. S2C and S2D). As peptides are prone to proteolytic degradation upon cellular uptake, we next determined the amount of intact peptides being delivered to the cells. To differentiate between full-length and degraded peptides, whole-cell lysates of HeLa and U2OS cells incubated with fluorescently labeled peptides were subjected to SDS-PAGE designed for resolving very low molecular weight protein species (27). Using this approach, we detected significant amounts of intact TAMRA- and Fluorescein-labeled R9-BRC4 peptides being effectively delivered to HeLa and U2OS cells (Fig. 2C; Supplementary Fig. S2E). Comparing the band intensity of TAMRA signals between 10 pmol of freshly solubilized peptides directly loaded onto the gel and those of peptide-transduced cell lysates, we calculated a delivery rate of approximately 107 peptides per cell, yielding an estimated intracellular peptide concentration of around 1 to 10 μmol/L (Fig. 2C).

Figure 2.

The CPP nona-arginine (R9) mediates efficient delivery of BRC4 peptides into human cells. A, HeLa and U2OS cells were stained for 30 minutes with Hoechst 33342 to visualize nuclei and washed with PBS before adding 10 μmol/L of the indicated TAMRA-labeled peptides. After 1-hour incubation in serum-free medium, cells were washed thoroughly to remove any membrane-bound peptides, and live cells were directly imaged on a confocal microscope. Gains were adjusted for each condition to reduce background fluorescence and to visualize strong and weak signals at the same time. Scale bar, 20 μm. B, HeLa and U2OS cells were transfected with 10 μmol/L of the indicated TAMRA-labeled peptides in serum-free medium for 1 hour. Cells were harvested by trypsinization, and TAMRA signal intensities were recorded on a flow cytometer by scoring 20,000 events. Bar graph depicts median TAMRA fluorescence intensity. a.u., arbitrary units. Data, mean ± SD (n = 3). C, HeLa and U2OS cells were incubated with 10 μmol/L of the indicated TAMRA-labeled peptides. After 1 hour, cells were harvested by trypsinization. Freshly solubilized peptides (10 pmol; lanes 2–5) and whole-cell lysates of 70,000 peptide-transfected cells (lanes 6–15) were loaded onto Tricine SDS-PAGE gel, and TAMRA signals were visualized using a FluoroImager.

Figure 2.

The CPP nona-arginine (R9) mediates efficient delivery of BRC4 peptides into human cells. A, HeLa and U2OS cells were stained for 30 minutes with Hoechst 33342 to visualize nuclei and washed with PBS before adding 10 μmol/L of the indicated TAMRA-labeled peptides. After 1-hour incubation in serum-free medium, cells were washed thoroughly to remove any membrane-bound peptides, and live cells were directly imaged on a confocal microscope. Gains were adjusted for each condition to reduce background fluorescence and to visualize strong and weak signals at the same time. Scale bar, 20 μm. B, HeLa and U2OS cells were transfected with 10 μmol/L of the indicated TAMRA-labeled peptides in serum-free medium for 1 hour. Cells were harvested by trypsinization, and TAMRA signal intensities were recorded on a flow cytometer by scoring 20,000 events. Bar graph depicts median TAMRA fluorescence intensity. a.u., arbitrary units. Data, mean ± SD (n = 3). C, HeLa and U2OS cells were incubated with 10 μmol/L of the indicated TAMRA-labeled peptides. After 1 hour, cells were harvested by trypsinization. Freshly solubilized peptides (10 pmol; lanes 2–5) and whole-cell lysates of 70,000 peptide-transfected cells (lanes 6–15) were loaded onto Tricine SDS-PAGE gel, and TAMRA signals were visualized using a FluoroImager.

Close modal

Next, to more precisely determine the intracellular residence time and stability of our BRC4 CPPs, we modified a previously established method (33) and monitored the Fluorescein signal intensity in HeLa cells over a time course of 24 hours after peptide incubation. SDS-PAGE analysis revealed that BRC4wt as well as BRC4mut cargo peptides were gradually degraded with an approximate half-life of 2 hours (Fig. 3A). Strikingly, using the same experimental setup, flow cytometry analysis of Fluorescein-R9-BRC4wt indicated a rather heterogeneous degradation pattern, which was most pronounced after 8 hours with a large proportion of cells still showing moderate to high fluorescent intensities (Fig. 3B).

Figure 3.

R9-BRC4 peptides have an approximate half-life of 2 to 4 hours upon cell entry. A, HeLa cells were transfected with 10 μmol/L of the indicated Fluorescein-labeled peptides in serum-free DMEM for 1 hour, followed by a release in serum-containing medium for the indicated time points. Top, 10 pmol of freshly solubilized peptides (lane 2) and whole-cell lysates of peptide-transduced cells (lanes 3–7) were loaded onto Tricine SDS-PAGE gel, and Fluorescein signals were analyzed using a FluoroImager. Bottom, quantification of relative Fluorescein signal intensities of intact full-length R9-BRC4wt and R9-BRC4mut peptides. Data, mean ± SD (n = 4). B, HeLa cells incubated with Fluorescein-R9-BRCwt peptides as in A were analyzed by flow cytometry recording 20,000 events for each condition. The intensity of the Fluorescein signal is plotted against the side-scatter area (SSC-A). Gates display percentages of cells showing moderate to high fluorescent intensities.

Figure 3.

R9-BRC4 peptides have an approximate half-life of 2 to 4 hours upon cell entry. A, HeLa cells were transfected with 10 μmol/L of the indicated Fluorescein-labeled peptides in serum-free DMEM for 1 hour, followed by a release in serum-containing medium for the indicated time points. Top, 10 pmol of freshly solubilized peptides (lane 2) and whole-cell lysates of peptide-transduced cells (lanes 3–7) were loaded onto Tricine SDS-PAGE gel, and Fluorescein signals were analyzed using a FluoroImager. Bottom, quantification of relative Fluorescein signal intensities of intact full-length R9-BRC4wt and R9-BRC4mut peptides. Data, mean ± SD (n = 4). B, HeLa cells incubated with Fluorescein-R9-BRCwt peptides as in A were analyzed by flow cytometry recording 20,000 events for each condition. The intensity of the Fluorescein signal is plotted against the side-scatter area (SSC-A). Gates display percentages of cells showing moderate to high fluorescent intensities.

Close modal

Collectively, we concluded that intracellular uptake of R9-BRC4wt and R9-BRC4mut was efficient and comparable as determined by confocal microscopy, flow cytometry, and SDS-PAGE analysis, thus providing a solid basis for further mechanistic investigations.

BRC4 peptide specifically inhibits RAD51-mediated HR

Conditional overexpression of full-length BRC4 was previously shown to inhibit DNA damage–induced RAD51 foci formation in MCF7 (breast cancer) and chicken DT40 cells (12, 34). Therefore, we wanted to investigate whether the 16-mer BRC4 peptide fused to R9 was able to mimic this phenotype, indicative of effective disruption of RAD51 binding to BRCA2. To this end, HeLa or U2OS cells were first incubated with our peptides and subsequently treated with the DNA topoisomerase I poison camptothecin to induce replication-dependent DSBs. Importantly, we did not detect any significant change in camptothecin-induced ATM and CHK2 phosphorylation as well as γH2AX foci formation upon peptide addition, indicating regular activation of apical DNA damage response kinases (Fig. 4A; Supplementary Fig. S3A and S3B). In contrast, camptothecin-induced RAD51 foci formation in both cell lines was specifically compromised in the presence of BRC4wt but not BRC4mut CPPs (Fig. 4B and C; Supplementary Fig. S3C and S3D). In line with this observation, BRC4wt significantly suppressed RAD51 foci formation following ionizing radiation (Fig. 4D). Notably, intrinsic RAD51 protein stability was not affected by BRC4wt cellular uptake (Supplementary Fig. S3E). To directly examine the impact of BRC4 peptides on DSB repair by HR, we performed repair reporter assays (DR-GFP) in HeLa cells and observed a significant decrease in HR frequency upon repetitive BRC4wt transfections (Fig. 4E). Throughout all experiments reported in this section, we observed that siRNA-mediated BRCA2 depletion conferred much stronger phenotypes compared with delivery of the BRC4wt peptide.

Figure 4.

BRC4 peptide inhibits RAD51 foci formation and HR. A, HeLa cells were either transfected with control (CNTL) or BRCA2 siRNA oligos for 48 hours, or incubated with 10 μmol/L of the indicated CPPs for 1 hour before treatment with 100 nmol/L camptothecin (CPT) for 1 hour. Whole-cell lysates were analyzed by immunoblotting using the indicated antibodies. B and C, Same cells as in A were treated with 100 nmol/L camptothecin for 1 hour before transfecting them with the depicted peptides (10 μmol/L) for 1 hour and releasing them for 3 hours in serum-containing medium. Cells were preextracted, fixed, and immunostained for RAD51. B, Scale bars, 10 and 5 μm (Zoom). C, The graph illustrates the percentage of cells displaying more than 10 RAD51 foci/nucleus. For each condition, at least 100 cells were scored. D, Same cells as in A were irradiated with increasing doses of IR, peptide-transfected (10 μmol/L), and released for 3 hours in serum-containing medium prior to preextraction, fixation, and immunostaining for RAD51. The graph illustrates the percentage of cells displaying more than 10 RAD51 foci/nucleus. For each condition, at least 100 cells were scored. E, HeLa DR-GFP cells were transfected with the I-SceI expression plasmid 48 hours post-siRNA transfection. Alternatively, 1-hour incubation periods with 10 μmol/L peptides took place 4, 24, and 34 hours post–I-SceI transfection of HeLa DR-GFP cells. A second siRNA transfection was performed 24 hours after I-SceI transfection. Cells were harvested 48 hours post–I-SceI transfection, and GFP-positive cells were scored by flow cytometry. A total of 20,000 events were recorded. The graph illustrates the percentage of GFP-positive cells as a readout for HR efficiency. Data in C–E display the mean ± SD (n ≥ 3). Statistical significance (***, P ≤ 0.001; ns, nonsignificant) was calculated with Tukey multiple comparison test using two-way ANOVA (C and D) and one-way ANOVA (E).

Figure 4.

BRC4 peptide inhibits RAD51 foci formation and HR. A, HeLa cells were either transfected with control (CNTL) or BRCA2 siRNA oligos for 48 hours, or incubated with 10 μmol/L of the indicated CPPs for 1 hour before treatment with 100 nmol/L camptothecin (CPT) for 1 hour. Whole-cell lysates were analyzed by immunoblotting using the indicated antibodies. B and C, Same cells as in A were treated with 100 nmol/L camptothecin for 1 hour before transfecting them with the depicted peptides (10 μmol/L) for 1 hour and releasing them for 3 hours in serum-containing medium. Cells were preextracted, fixed, and immunostained for RAD51. B, Scale bars, 10 and 5 μm (Zoom). C, The graph illustrates the percentage of cells displaying more than 10 RAD51 foci/nucleus. For each condition, at least 100 cells were scored. D, Same cells as in A were irradiated with increasing doses of IR, peptide-transfected (10 μmol/L), and released for 3 hours in serum-containing medium prior to preextraction, fixation, and immunostaining for RAD51. The graph illustrates the percentage of cells displaying more than 10 RAD51 foci/nucleus. For each condition, at least 100 cells were scored. E, HeLa DR-GFP cells were transfected with the I-SceI expression plasmid 48 hours post-siRNA transfection. Alternatively, 1-hour incubation periods with 10 μmol/L peptides took place 4, 24, and 34 hours post–I-SceI transfection of HeLa DR-GFP cells. A second siRNA transfection was performed 24 hours after I-SceI transfection. Cells were harvested 48 hours post–I-SceI transfection, and GFP-positive cells were scored by flow cytometry. A total of 20,000 events were recorded. The graph illustrates the percentage of GFP-positive cells as a readout for HR efficiency. Data in C–E display the mean ± SD (n ≥ 3). Statistical significance (***, P ≤ 0.001; ns, nonsignificant) was calculated with Tukey multiple comparison test using two-way ANOVA (C and D) and one-way ANOVA (E).

Close modal

In summary, our findings suggested that delivery of a synthetic BRC4 peptide potently inhibits RAD51 foci formation and HR, most likely as a result of defective BRCA2-mediated RAD51 loading onto resected DSBs.

BRC4 peptide causes MRE11-dependent degradation of stalled replication forks

BRCA2-dependent RAD51 loading is not only critical for HR, but has also been established to protect stalled replication forks from nucleolytic degradation by MRE11 (35, 36). Thus, we examined a potential effect of the BRC4 peptide on replication fork protection by performing dual-labeling DNA fiber assays in the presence of HU, which stalls fork progression. Strikingly, we found that the BRC4wt CPP resulted in shortening of nascent DNA tracts, indicative of increased fork degradation (Fig. 5A). BRC4 intracellular uptake did not interfere with global replication rates, as we could not observe any differences in EdU incorporation (Supplementary Fig. S4). Similar to what has been shown for BRCA2-deficient cells, fork degradation in BRC4 peptide–transduced cells was completely rescued both by mirin, an MRE11 SMI (Fig. 5B), and by siRNA-mediated MRE11 depletion (Fig. 5C). Of note, BRC4 delivery did not result in additive or synergistic effects when delivered into BRCA2-depleted cells, indicating that peptide-mediated fork degradation resulted most likely from specific targeting of the BRCA2–RAD51 interaction (Fig. 5C).

Figure 5.

BRC4 peptide provokes MRE11-dependent degradation of stalled DNA replication forks. A, U2OS cells were transfected with indicated siRNA oligos for 48 hours or left untransfected. Subsequently, cells were labeled with CIdU and IdU before adding hydroxyurea (HU) together with indicated peptides (10 μmol/L) in serum-free medium for 4 hours followed by DNA fiber spreading. Top, experimental setup, representative DNA fiber images, and their corresponding IdU/CIdU ratios, respectively; bottom, IdU/CIdU ratios for each condition illustrated by a dot plot with mean ratios denoted as a red line. B, Same cells as in A were additionally treated with the MRE11 inhibitor mirin. C, U2OS cells were transfected with the indicated siRNA oligos for 48 hours before dual labeling and HU treatment in combination with 10 μmol/L R9 or R9-BRC4wt peptides. Representative immunblot is shown to indicate individual knockdown efficiencies. Approximately 250 DNA replication tracks were scored, and numbers indicate the mean ± SD (n = 2). Statistical differences (****, P ≤ 0.0001; ns, nonsignificant) were determined by Mann–Whitney U test.

Figure 5.

BRC4 peptide provokes MRE11-dependent degradation of stalled DNA replication forks. A, U2OS cells were transfected with indicated siRNA oligos for 48 hours or left untransfected. Subsequently, cells were labeled with CIdU and IdU before adding hydroxyurea (HU) together with indicated peptides (10 μmol/L) in serum-free medium for 4 hours followed by DNA fiber spreading. Top, experimental setup, representative DNA fiber images, and their corresponding IdU/CIdU ratios, respectively; bottom, IdU/CIdU ratios for each condition illustrated by a dot plot with mean ratios denoted as a red line. B, Same cells as in A were additionally treated with the MRE11 inhibitor mirin. C, U2OS cells were transfected with the indicated siRNA oligos for 48 hours before dual labeling and HU treatment in combination with 10 μmol/L R9 or R9-BRC4wt peptides. Representative immunblot is shown to indicate individual knockdown efficiencies. Approximately 250 DNA replication tracks were scored, and numbers indicate the mean ± SD (n = 2). Statistical differences (****, P ≤ 0.0001; ns, nonsignificant) were determined by Mann–Whitney U test.

Close modal

BRC4 peptide confers hypersensitivity to PARP inhibition in cancer cell lines

On the basis of our molecular analyses providing robust evidence of the inhibitory effect of the BRC4 peptide on BRCA2 functions in HR and fork protection, we next sought to determine whether it also sensitizes cells to DNA-damaging agents. Profound hypersensitivity of BRCA2-mutant cells to PARP inhibitors has become an emerging therapeutic paradigm known as synthetic lethality (37). Therefore, we performed clonogenic survival assays in peptide-transfected HeLa cells treated with the PARP inhibitor olaparib (38). Indeed, targeting of the BRCA2–RAD51 interaction by BRC4wt CPPs resulted in a significant reduction in cellular viability in response to chronic PARP inhibition (Fig. 6A and B). Consistent with impaired RAD51 foci formation, delivery of BRC4wt CPPs also sensitized HeLa cells to camptothecin (Supplementary Fig. S5A). Upon treatment with olaparib, we estimated the IC50 value of BRC4 to be around 10 μmol/L for HeLa cells (Fig. 6C). Importantly, R9-BRC4wt alone did not decrease cell survival in otherwise undamaged cells (Fig. 6C). Moreover, despite its relatively short half-life, the efficacy of BRC4 CPPs in sensitizing HeLa cells to PARP inhibition was comparable with that of the B02 small-molecule compound, which specifically inhibits RAD51 binding to DNA (Fig. 6D; refs. 23, 39). To further exclude potential off-target effects of the BRC4 CPP in stimulating olaparib-induced cytotoxicity, we employed PEO1 (BRCA2 null) and PEO4 (BRCA2 revertant to wild type) isogenic ovarian cancer cell lines (40). Notably, we observed a strong synergy between PARP inhibition and R9-BRC4 peptides in PEO4 cells, whereas PEO1 cells did not exhibit BRC4-mediated sensitivity toward olaparib (Supplementary Fig. S5B). Differential peptide uptake could be excluded as both cell lines displayed comparable uptake efficiency and intracellular peptide localization (Supplementary Fig. S5B). Finally, in addition to HeLa cells, we observed that BRC4 CPPs elicited olaparib-induced cell death of U2OS human osteosarcoma cells but not of noncancerous cell lines, including hTERT-immortalized human retinal pigment epithelial cells (hTERT-RPE1), SV40-immortalized human fetal lung fibroblasts (MRC5) and human breast epithelial cells (MCF10A; Fig. 6E; Supplementary Fig. S6A). This finding could at least in part be explained by an overall low cellular uptake efficiency and predominant endosome trapping of BRC4 peptides in RPE1, MRC5, and MCF10A nontumorigenic cell lines (Supplementary Fig. S6B and S6C).

Figure 6.

BRC4 peptide sensitizes cancer cells to PARP inhibition. A, HeLa cells were transfected with the indicated siRNAs or left untransfected for standard peptide transduction. Twenty-four hours later, cells were plated at low cell density into 24-well plates. Another 24 hours later, non-siRNA transfected cells were incubated for 1 hour with 10 μmol/L of peptides in serum-free medium supplemented with 1 μmol/L olaparib. Finally, serum-containing medium supplemented with 1 μmol/L olaparib was added, cells were grown for 10 days, and colonies were stained using crystal violet. Representative images of a colony formation assay are shown. Clonogenic survival was determined by quantifying the colony intensity of olaparib-treated relative to untreated and is depicted as bar graph. Data, mean ± SD (n = 4). Statistical significance (***, P ≤ 0.001; ns, nonsignificant) was calculated with Tukey multiple comparison test using one-way ANOVA. B, Peptide-transfected HeLa cells as in A were treated with increasing concentrations of olaparib, and survival was determined by colony formation assay. C, Same cells as in A were treated with 1 μmol/L olaparib in combination with increasing concentrations of BRC4 peptides, and survival was determined by colony formation assay. D, Same cells as in B were treated with 10 μmol/L of the RAD51 inhibitor BO2 under identical experimental conditions as used for the BRC4 peptide, and survival was determined by colony formation assay. Data in B–D are represented as mean ± SD (n ≤ 3). E, HeLa, U2OS, RPE1, MRC5, and MCF10A were identically treated as in A, and survival was determined by colony formation assay. Bottom, knockdown efficiencies were analyzed by immunoblotting of whole-cell lysates of indicated cell lines; Top, relative clonogenic survival is depicted as mean ± SD (n = 4). Statistical significance (****, P ≤ 0.0001; **, P ≤ 0.01; ns, nonsignificant) was calculated with Tukey multiple comparison test using one-way ANOVA.

Figure 6.

BRC4 peptide sensitizes cancer cells to PARP inhibition. A, HeLa cells were transfected with the indicated siRNAs or left untransfected for standard peptide transduction. Twenty-four hours later, cells were plated at low cell density into 24-well plates. Another 24 hours later, non-siRNA transfected cells were incubated for 1 hour with 10 μmol/L of peptides in serum-free medium supplemented with 1 μmol/L olaparib. Finally, serum-containing medium supplemented with 1 μmol/L olaparib was added, cells were grown for 10 days, and colonies were stained using crystal violet. Representative images of a colony formation assay are shown. Clonogenic survival was determined by quantifying the colony intensity of olaparib-treated relative to untreated and is depicted as bar graph. Data, mean ± SD (n = 4). Statistical significance (***, P ≤ 0.001; ns, nonsignificant) was calculated with Tukey multiple comparison test using one-way ANOVA. B, Peptide-transfected HeLa cells as in A were treated with increasing concentrations of olaparib, and survival was determined by colony formation assay. C, Same cells as in A were treated with 1 μmol/L olaparib in combination with increasing concentrations of BRC4 peptides, and survival was determined by colony formation assay. D, Same cells as in B were treated with 10 μmol/L of the RAD51 inhibitor BO2 under identical experimental conditions as used for the BRC4 peptide, and survival was determined by colony formation assay. Data in B–D are represented as mean ± SD (n ≤ 3). E, HeLa, U2OS, RPE1, MRC5, and MCF10A were identically treated as in A, and survival was determined by colony formation assay. Bottom, knockdown efficiencies were analyzed by immunoblotting of whole-cell lysates of indicated cell lines; Top, relative clonogenic survival is depicted as mean ± SD (n = 4). Statistical significance (****, P ≤ 0.0001; **, P ≤ 0.01; ns, nonsignificant) was calculated with Tukey multiple comparison test using one-way ANOVA.

Close modal

PPIs represent attractive but at the same time challenging targets for pharmacologic intervention in cancer therapy. The emergence of successful peptide-based PPI inhibitors seemed promising, yet drug development was faced with the low membrane translocation ability of native peptides. The identification of a wide range of carrier molecules, commonly termed as CPPs or protein transduction domains, provided a possible solution for this major limitation (19).

Here, we designed a 16-amino acid short peptide derived from the N-terminal half of the human BRCA2 BRC4 repeat able to occupy the RAD51 oligomerization interface and thereby potently inhibiting the interaction between BRCA2 and RAD51. Covalent fusion of the BRC4 peptide with the cationic polyarginine R9 CPP resulted in efficient uptake into both cytoplasm and nucleus, causing defects in HR repair, increased fork degradation, and, ultimately, hypersensitivity to DNA-damaging agents.

Detailed biochemical studies have demonstrated that the full-length BRC4 repeat covering both FHTA and LFDE tetrameric cluster motifs selectively interacts with RAD51 and inhibits RAD51–DNA complex formation (15, 41). To increase the likelihood of intracellular peptide uptake (42), we reduced the length of the native BRC4 peptide sequence from 35 to 16 amino acids, including only the first FHTA module, which was shown to mimic the RAD51 oligomerization motif (14). In line with our findings, a similar 17-mer N-terminal BRC4 peptide was reported to compete with full-length BRC4 for RAD51 binding, albeit with modest potency (15). We further corroborated the importance of the FHTA sequence by showing that BRC4 16-mer peptides harboring a mirrored ATHF sequence display greatly reduced RAD51 binding affinity.

Importantly, intracellular BRC4 peptide uptake in the low micromolar range was sufficient to impair BRCA2–RAD51 complex formation. A high uptake efficiency was likely achieved through a combination of the potent CPP R9 with the beneficial biochemical properties of the BRC4 peptide, including its positive net charge, hydrophobicity, and three-dimensional conformation (18). CPP internalization involves both endocytosis and direct translocation, whereby direct cell penetration is preferred because endocytosis can lead to the trapping of peptides in endosomes and their ultimate degradation (18). Multiple factors, including the CPP, cargo identity, concentration, cell type, and differentiation status, influence the balance between both uptake mechanisms (18). We primarily observed a diffuse fluorescent staining and only minor punctuate patterns, indicative of direct translocation as the preferred uptake mechanism. The substantial number of peptide molecules reaching the intracellular space might compensate for the short half-life. Even though only 5% of intact peptides are present 24 hours after transfection, the absolute peptide number per cell at this stage still adds up to 106 functional peptides. According to a recent proteomics study in the U2OS cell line, most proteins involved in DSB repair processes, including BRCA2, were shown to be of low to moderate abundance (102–104 copies/cell; ref. 43). On the basis of these numbers, we speculated that BRC4 peptides are present in high molar excess over their endogenous BRCA2 target protein even 24 hours after uptake.

RAD51 and its property to self-assemble are under dynamic control to enable faithful homology-directed DSB repair. Yu and colleagues proposed the presence of three RAD51 fractions in the cell: a mobile fraction of RAD51 monomers, an immobile oligomerized fraction, and an immobile BRCA2-bound fraction (44). Strikingly, they found the BRCA2-bound fraction to be selectively mobilized upon DNA damage and suggested a dual BRCA2 function of RAD51 sequestration and mobilization. Importantly, the other fractions did not change, suggesting that only the BRCA2-bound RAD51 fraction is involved in DNA repair. Notably, R9-BRC4wt did not affect viability under unstressed growth conditions of conventionally cultured cancer cell lines. Thus, we hypothesize that under normal, undamaged conditions, RAD51 is stably bound to BRCA2 and the peptide's affinity for RAD51 is not high enough to disassemble the protein complex. Accordingly, the BRC4 mimetic peptide is only able to disassemble mobilized BRCA2–RAD51 complexes. We speculate that upon mobilization and ultimate RAD51 loading onto ssDNA, the protein complex becomes increasingly unstable due to an equilibrium change during DNA loading. In this context, the peptide is potent enough to sequester RAD51 transiently dissociated from BRCA2. We further presume that at low levels of DNA damage, requiring only a small fraction of BRCA2-bound RAD51, repair processes are not significantly disturbed by the BRC4 peptide. However, upon increasing DNA damage load and replication stress, when there is a high demand of functional BRCA2–RAD51 complexes, a major pool of RAD51 is sequestered away from BRCA2 by the BRC4 peptide.

Despite its potentially rather high dissociation constant compared with SMIs, BRC4 peptides provided an unexpectedly high degree of specificity. CPPs have often been reported to cause unspecific membrane disruption, which has led to false conclusions regarding their specificity for a number of peptide therapeutics. Because we do not observe a decrease in cell viability upon BRC4 peptide uptake, we can exclude such unspecific CPP-mediated toxicity (45). Furthermore, our finding that MRE11 depletion fully rescued the peptide-induced replication fork degradation phenotype and the lack of synergy between peptides and PARP inhibition in BRCA2-deficient cancer cells argue against potential off-target activity of the BRC4 CPP. Nevertheless, when comparing the outcomes of siRNA-mediated BRCA2 depletion versus R9-BRC4wt peptide delivery, depletion is commonly yielding much more pronounced phenotypes. A possible explanation for this relates to the high susceptibility of peptides to extra- and intracellular protease attacks (17). To improve their pharmacokinetic potential, possible peptide optimizations include various backbone modifications and cyclization (17, 46). Moreover, the BRCA2–RAD51 interaction might not be completely abolished by the BRC4 peptide as it only comprises the first FHTA but lacks the second LFDE tetrameric cluster motif.

The lack of cell-type specificity is commonly reported to present another major limitation of CPP-mediated drug delivery (19, 20). Strikingly, we could not find a synergistic relationship between the PARP inhibitor olaparib and the R9-BRC4 peptide in normal cell lines as compared with cancer cell lines. Interestingly, our peptide uptake data suggest an overall decreased intracellular BRC4 concentration. It is reasonable to speculate that efficient intracellular peptide delivery relies on specific membrane components or lipid compositions that are more uptake responsive in cancer versus noncancerous cell lines. Numerous reports classify lipid metabolic reprogramming as a major source of cell transformation (47). Specifically, the sphingolipid metabolism, which was suggested to influence R9-mediated peptide uptake, was reported to significantly alter during transformation (48). In addition, cancer cells possess a more negatively charged membrane than normal cells, which is partly caused by a loss in membrane symmetry and the exposure of anionic phosphatidylcholine on the outer leaflet (49). This feature, in combination with higher membrane fluidity (50), might favor the uptake of the cationic R9-BRC4 peptide.

Ultimately, the BRC4 peptide could emerge as a potent radio- and chemosensitizer. Specifically, BRC4 peptide–induced HR deficiency could represent a promising strategy for expanding the utility of PARP inhibitors, successfully applied in breast and ovarian cancer patients with a germline BRCA1 or BRCA2 mutation (51, 52), to BRCA-proficient cancers. Similarly, the BRC4 peptide could resensitize BRCA1/2-mutated tumors that acquired chemoresistance to PARP inhibitors by restoring HR (53). Moreover, malignant cells are frequently compromised in genome stability maintenance pathways that are synthetically lethal with HR deficiency (54). Consequently, monotherapy with the identified BRCA2 peptide inhibitor could provide a promising option for the treatment of these tumors. The BRCA2–BRC4 peptide described in this study could also be applied in combination with proton irradiation, which was shown to be highly advantageous over conventional photon therapy. Notably, it was found that HR-deficient tumor cells exhibit an enhanced susceptibility toward proton versus photon irradiation (55). Drug-induced HR deficiency with the BRC4 peptide inhibitor could make this advantageous effect accessible for patients with HR-proficient tumors. Finally, we reveal a straightforward approach to study distinct PPIs in a biological context without the need of elaborate mutagenesis methodologies and provide a potent research tool to study BRCA2-dependent RAD51 loading to ssDNA in different biological contexts.

No potential conflicts of interest were disclosed.

Conception and design: A. Trenner, A.A. Sartori

Development of methodology: A. Trenner, A.A. Sartori

Acquisition of data (provided animals, acquired and managed patients, provided facilities, etc.): A. Trenner, J. Godau,

Analysis and interpretation of data (e.g., statistical analysis, biostatistics, computational analysis): A. Trenner, A.A. Sartori

Writing, review, and/or revision of the manuscript: A. Trenner, J. Godau, A.A. Sartori

Study supervision: A.A. Sartori

The authors would like to thank Roland Brock (Institute of Molecular Life Sciences, Radboud University Nijmengen, the Netherlands), Martin Pruschy (Clinic of Radiation Oncology, University Hospital Zurich, Zurich, Switzerland), and Stefano Ferrari (Institute of Molecular Cancer Research, University of Zurich, Zurich, Switzerland) for scientific input. We are grateful to Pavel Janscak (Institute of Molecular Cancer Research, University of Zurich) for providing purified, recombinant RAD51 protein and anti-RAD51 rabbit polyclonal antibody and to Fumiko Esashi (Sir William Dunn School of Pathology, University of Oxford, Oxford, United Kingdom) for providing the GST-tagged BRCA2 fusion plasmids. Furthermore, we are grateful to Ross Chapman (Nuffield Department of Medicine, University of Oxford) for providing hTERT-RPE1 cells, Minoru Takata (Radiation Biology Center, Kyoto University, Kyoto, Japan) for the U2OS GFP-RAD51 cell line, Marcel van Vugt (Medical Oncology Department, University Medical Centre Groningen, Groningen, the Netherlands) for HeLa DR-GFP cells, and Steve Jackson (Gurdon Institute, Cambridge, United Kingdom) for MRC5 cells. We also want to thank Antonio Porro (Institute of Molecular Cancer Research, University of Zurich) for critical reading of the manuscript. This study was supported by research grants from the Krebsliga Schweiz (KFS-3845-02-2016, to A.A. Sartori), Swiss National Science Foundation (31003A_156023 and 31003A_176161, to A.A. Sartori), Promedica Stiftung (1317/M to A.A. Sartori), and Novartis Foundation for Medical-Biological Research (#17C155 to A.A. Sartori).

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.

1.
Negrini
S
,
Gorgoulis
VG
,
Halazonetis
TD
. 
Genomic instability–an evolving hallmark of cancer
.
Nat Rev Mol Cell Biol
2010
;
11
:
220
8
.
2.
Ceccaldi
R
,
Rondinelli
B
,
D'Andrea
AD
. 
Repair pathway choices and consequences at the double-strand break
.
Trends Cell Biol
2016
;
26
:
52
64
.
3.
Kolinjivadi
AM
,
Sannino
V
,
de Antoni
A
,
Técher
H
,
Baldi
G
,
Costanzo
V
. 
Moonlighting at replication forks - a new life for homologous recombination proteins BRCA1, BRCA2 and RAD51
.
FEBS Lett
2017
;
591
:
1083
100
.
4.
Hühn
D
,
Bolck
HA
,
Sartori
AA
. 
Targeting DNA double-strand break signalling and repair: recent advances in cancer therapy
.
Swiss Med Wkly
2013
;
143
:
w13837
.
5.
Carvalho
JFS
,
Kanaar
R
. 
Targeting homologous recombination-mediated DNA repair in cancer
.
Expert Opin Ther Targets
2014
;
18
:
427
58
.
6.
Fradet-Turcotte
A
,
Sitz
J
,
Grapton
D
,
Orthwein
A
. 
BRCA2 functions: from DNA repair to replication fork stabilization
.
Endocr Relat Cancer
2016
;
23
:
T1
T17
.
7.
Liu
J
,
Doty
T
,
Gibson
B
,
Heyer
WD
. 
Human BRCA2 protein promotes RAD51 filament formation on RPA-covered single-stranded DNA
.
Nat Struct Mol Biol
2010
;
17
:
1260
2
.
8.
Wong
AK
,
Pero
R
,
Ormonde
PA
,
Tavtigian
SV
,
Bartel
PL
. 
RAD51 interacts with the evolutionarily conserved BRC motifs in the human breast cancer susceptibility gene brca2
.
J Biol Chem
1997
;
272
:
31941
4
.
9.
Esashi
F
,
Galkin
VE
,
Yu
X
,
Egelman
EH
,
West
SC
. 
Stabilization of RAD51 nucleoprotein filaments by the C-terminal region of BRCA2
.
Nat Struct Mol Biol
2007
;
14
:
468
74
.
10.
Davies
OR
,
Pellegrini
L
. 
Interaction with the BRCA2 C terminus protects RAD51-DNA filaments from disassembly by BRC repeats
.
Nat Struct Mol Biol
2007
;
14
:
475
83
.
11.
Carreira
A
,
Kowalczykowski
SC
. 
Two classes of BRC repeats in BRCA2 promote RAD51 nucleoprotein filament function by distinct mechanisms
.
Proc Natl Acad Sci U S A
2011
;
108
:
10448
53
.
12.
Chen
CF
,
Chen
PL
,
Zhong
Q
,
Sharp
ZD
,
Lee
WH
. 
Expression of BRC repeats in breast cancer cells disrupts the BRCA2-Rad51 complex and leads to radiation hypersensitivity and loss of G(2)/M checkpoint control
.
J Biol Chem
1999
;
274
:
32931
5
.
13.
Carreira
A
,
Hilario
J
,
Amitani
I
,
Baskin
RJ
,
Shivji
MK
,
Venkitaraman
AR
, et al
The BRC repeats of BRCA2 modulate the DNA-binding selectivity of RAD51
.
Cell
2009
;
136
:
1032
43
.
14.
Pellegrini
L
,
Yu
DS
,
Lo
T
,
Anand
S
,
Lee
M
,
Blundell
TL
, et al
Insights into DNA recombination from the structure of a RAD51-BRCA2 complex
.
Nature
2002
;
420
:
287
93
.
15.
Rajendra
E
,
Venkitaraman
AR
. 
Two modules in the BRC repeats of BRCA2 mediate structural and functional interactions with the RAD51 recombinase
.
Nucleic Acids Res
2009
;
38
:
82
96
.
16.
Ivanov
AA
,
Khuri
FR
,
Fu
H
. 
Targeting protein-protein interactions as an anticancer strategy
.
Trends Pharmacol Sci
2013
;
34
:
393
400
.
17.
Tsomaia
N
. 
Peptide therapeutics: targeting the undruggable space
.
Eur J Med Chem
2015
;
94
:
459
70
.
18.
Kristensen
M
,
Birch
D
,
Mørck Nielsen
H
. 
Applications and challenges for use of cell-penetrating peptides as delivery vectors for peptide and protein cargos
.
Int J Mol Sci
2016
;
17
:
pii
:
185
.
19.
Guidotti
G
,
Brambilla
L
,
Rossi
D
. 
Cell-penetrating peptides: from basic research to clinics
.
Trends Pharmacol Sci
2017
;
38
:
406
24
.
20.
Raucher
D
,
Ryu
JS
. 
Cell-penetrating peptides: strategies for anticancer treatment
.
Trends Mol Med
2015
;
21
:
560
70
.
21.
Tünnemann
G
,
Ter-Avetisyan
G
,
Martin
RM
,
Stöckl
M
,
Herrmann
A
,
Cardoso
MC
. 
Live-cell analysis of cell penetration ability and toxicity of oligo-arginines
.
J Pept Sci
2008
;
14
:
469
76
.
22.
Sato
K
,
Shimomuki
M
,
Katsuki
Y
,
Takahashi
D
,
Kobayashi
W
,
Ishiai
M
, et al
FANCI-FANCD2 stabilizes the RAD51-DNA complex by binding RAD51 and protects the 5′-DNA end
.
Nucleic Acids Res
2016
;
44
:
10758
71
.
23.
Huang
F
,
Motlekar
NA
,
Burgwin
CM
,
Napper
AD
,
Diamond
SL
,
Mazin
AV
. 
Identification of specific inhibitors of human RAD51 recombinase using high-throughput screening
.
ACS Chem Biol
2011
;
6
:
628
35
.
24.
Dupré
A
,
Boyer-Chatenet
L
,
Sattler
RM
,
Modi
AP
,
Lee
JH
,
Nicolette
ML
, et al
A forward chemical genetic screen reveals an inhibitor of the Mre11-Rad50-Nbs1 complex
.
Nat Chem Biol
2008
;
4
:
119
25
.
25.
Lee
M
,
Daniels
MJ
,
Venkitaraman
AR
. 
Phosphorylation of BRCA2 by the Polo-like kinase Plk1 is regulated by DNA damage and mitotic progression
.
Oncogene
2003
;
23
:
865
72
.
26.
Piwko
W
,
Mlejnkova
LJ
,
Mutreja
K
,
Ranjha
L
,
Stafa
D
,
Smirnov
A
, et al
The MMS22L-TONSL heterodimer directly promotes RAD51-dependent recombination upon replication stress
.
EMBO J
2016
;
35
:
2584
601
.
27.
Schägger
H
. 
Tricine-SDS-PAGE
.
Nat Protoc
2006
;
1
:
16
22
.
28.
Steger
M
,
Murina
O
,
Hühn
D
,
Ferretti
LP
,
Walser
R
,
Hänggi
K
, et al
Prolyl isomerase PIN1 regulates DNA double-strand break repair by counteracting DNA end resection
.
Mol Cell
2013
;
50
:
333
43
.
29.
Krajewska
M
,
Fehrmann
RSN
,
Schoonen
PM
,
Labib
S
,
de Vries
EG
,
Franke
L
, et al
ATR inhibition preferentially targets homologous recombination-deficient tumor cells
.
Oncogene
2015
;
34
:
3474
81
.
30.
Merrick
CJ
,
Jackson
D
,
Diffley
JFX
. 
Visualization of altered replication dynamics after DNA damage in human cells
.
J Biol Chem
2004
;
279
:
20067
75
.
31.
Guzmán
C
,
Bagga
M
,
Kaur
A
,
Westermarck
J
,
Abankwa
D
. 
ColonyArea: an ImageJ plugin to automatically quantify colony formation in clonogenic assays
.
PLoS One
2014
;
9
:
e92444
.
32.
Esashi
F
,
Christ
N
,
Gannon
J
,
Liu
Y
,
Hunt
T
,
Jasin
M
, et al
CDK-dependent phosphorylation of BRCA2 as a regulatory mechanism for recombinational repair
.
Nature
2005
;
434
:
598
604
.
33.
Ruttekolk
IR
,
Witsenburg
JJ
,
Glauner
H
,
Bovee-Geurts
PH
,
Ferro
ES
,
Verdurmen
WP
, et al
The intracellular pharmacokinetics of terminally capped peptides
.
Mol Pharm
2012
;
9
:
1077
86
.
34.
Abe
T
,
Branzei
D
. 
High levels of BRC4 induced by a Tet-On 3G system suppress DNA repair and impair cell proliferation in vertebrate cells
.
DNA Repair
2014
;
22
:
153
64
.
35.
Schlacher
K
,
Christ
N
,
Siaud
N
,
Egashira
A
,
Wu
H
,
Jasin
M
. 
Double-strand break repair-independent role for BRCA2 in blocking stalled replication fork degradation by MRE11
.
Cell
2011
;
145
:
529
42
.
36.
Hashimoto
Y
,
Ray Chaudhuri
A
,
Lopes
M
,
Costanzo
V
. 
Rad51 protects nascent DNA from Mre11-dependent degradation and promotes continuous DNA synthesis
.
Nat Struct Mol Biol
2010
;
17
:
1305
11
.
37.
Farmer
H
,
McCabe
N
,
Lord
CJ
,
Tutt
AN
,
Johnson
DA
,
Richardson
TB
, et al
Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy
.
Nature
2005
;
434
:
917
21
.
38.
Menear
KA
,
Adcock
C
,
Boulter
R
,
Cockcroft
XL
,
Copsey
L
,
Cranston
A
, et al
4-[3-(4-cyclopropanecarbonylpiperazine-1-carbonyl)-4-fluorobenzyl]-2H-phthalazin-1-one: a novel bioavailable inhibitor of poly(ADP-ribose) polymerase-1
.
J Med Chem
2008
;
51
:
6581
91
.
39.
Huang
F
,
Mazina
OM
,
Zentner
IJ
,
Cocklin
S
,
Mazin
AV
. 
Inhibition of homologous recombination in human cells by targeting RAD51 recombinase
.
J Med Chem
2012
;
55
:
3011
20
.
40.
Sakai
W
,
Swisher
EM
,
Jacquemont
C
,
Chandramohan
KV
,
Couch
FJ
,
Langdon
SP
, et al
Functional restoration of BRCA2 protein by secondary BRCA2 mutations in BRCA2-mutated ovarian carcinoma
.
Cancer Res
2009
;
69
:
6381
6
.
41.
Davies
AA
,
Masson
JY
,
McIlwraith
MJ
,
Stasiak
AZ
,
Stasiak
A
,
Venkitaraman
AR
, et al
Role of BRCA2 in control of the RAD51 recombination and DNA repair protein
.
Mol Cell
2001
;
7
:
273
82
.
42.
Renukuntla
J
,
Vadlapudi
AD
,
Patel
A
,
Boddu
SH
,
Mitra
AK
. 
Approaches for enhancing oral bioavailability of peptides and proteins
.
Int J Pharm
2013
;
447
:
75
93
.
43.
Beck
M
,
Schmidt
A
,
Malmstroem
J
,
Claassen
M
,
Ori
A
,
Szymborska
A
, et al
The quantitative proteome of a human cell line
.
Mol Syst Biol
2011
;
7
:
549
.
44.
Yu
DS
,
Sonoda
E
,
Takeda
S
,
Huang
CL
,
Pellegrini
L
,
Blundell
TL
, et al
Dynamic control of Rad51 recombinase by self-association and interaction with BRCA2
.
Mol Cell
2003
;
12
:
1029
41
.
45.
El-Andaloussi
S
,
Järver
P
,
Johansson
HJ
,
Langel
Ü
. 
Cargo-dependent cytotoxicity and delivery efficacy of cell-penetrating peptides: a comparative study
.
Biochem J
2007
;
407
:
285
92
.
46.
Wójcik
P
,
Berlicki
Ł
. 
Peptide-based inhibitors of protein-protein interactions
.
Bioorg Med Chem Lett
2016
;
26
:
707
13
.
47.
Beloribi-Djefaflia
S
,
Vasseur
S
,
Guillaumond
F
. 
Lipid metabolic reprogramming in cancer cells
.
Oncogenesis
2016
;
5
:
e189
9
.
48.
Verdurmen
WP
,
Thanos
M
,
Ruttekolk
IR
,
Gulbins
E
,
Brock
R
. 
Cationic cell-penetrating peptides induce ceramide formation via acid sphingomyelinase: implications for uptake
.
J Control Release
2010
;
147
:
171
9
.
49.
Jobin
ML
,
Bonnafous
P
,
Temsamani
H
,
Dole
F
,
Grélard
A
,
Dufourc
EJ
, et al
The enhanced membrane interaction and perturbation of a cell penetrating peptide in the presence of anionic lipids: toward an understanding of its selectivity for cancer cells
.
Biochim Biophys Acta
2013
;
1828
:
1457
70
.
50.
Sok
M
,
Sentjurc
M
,
Schara
M
,
Stare
J
,
Rott
T
. 
Cell membrane fluidity and prognosis of lung cancer
.
Ann Thorac Surg
2002
;
73
:
1567
71
.
51.
Fong
PC
,
Boss
DS
,
Yap
TA
,
Tutt
A
,
Wu
P
,
Mergui-Roelvink
M
, et al
Inhibition of poly(ADP-ribose) polymerase in tumors from BRCA mutation carriers
.
N Engl J Med
2009
;
361
:
123
34
.
52.
Robson
M
,
Im
SA
,
Senkus
E
,
Xu
B
,
Domchek
SM
,
Masuda
N
, et al
Olaparib for metastatic breast cancer in patients with a germline BRCA mutation
.
N Engl J Med
2017
;
377
:
523
33
.
53.
Lord
CJ
,
Ashworth
A
. 
Mechanisms of resistance to therapies targeting BRCA-mutant cancers
.
Nat Med
2013
;
19
:
1381
8
.
54.
Chernikova
SB
,
Game
JC
,
Brown
JM
. 
Inhibiting homologous recombination for cancer therapy
.
Cancer Biol Ther
2012
;
13
:
61
8
.
55.
Fontana
AO
,
Augsburger
MA
,
Grosse
N
,
Guckenberger
M
,
Lomax
AJ
,
Sartori
AA
, et al
Differential DNA repair pathway choice in cancer cells after proton- and photon-irradiation
.
Radiother Oncol
2015
;
116
:
374
80
.