Inhibitors of PARP: Number crunching and structure gazing

Johannes Rudolph, Karen Jung, Karolin Luger, Johannes Rudolph, Karen Jung, Karolin Luger

Abstract

SignificancePARP is an important target in the treatment of cancers, particularly in patients with breast, ovarian, or prostate cancer that have compromised homologous recombination repair (i.e., BRCA-/-). This review about inhibitors of PARP (PARPi) is for readers interested in the development of next-generation drugs for the treatment of cancer, providing insights into structure-activity relationships, in vitro vs. in vivo potency, PARP trapping, and synthetic lethality.

Keywords: HPF1; cancer drugs; drug specificity; inhibitor of Parp; synthetic lethality.

Conflict of interest statement

The authors declare no competing interest.

Figures

Fig. 1.
Fig. 1.
(A) Domain structure of PARP1 and PARP2. PARP1/2 contain DNA-binding domains and catalytic domains. In PARP1, the DNA-binding domain includes three zinc fingers (Zn1, Zn2, and Zn3), a breast cancer susceptibility protein-1 C terminus (BRCT) domain, and the tryptophan-glycine-arginine–rich (WGR) domain. The locations of the predominant automodification sites of PARP1 (D387, E488, E491, S499, S507, and S519) are indicated by red uptick bars. In PARP2, the DNA-binding domain includes an unstructured N-terminal region and the WGR domain. The catalytic domains of both PARP1/2 are composed of an alpha-helical subdomain (HD) and the ADP ribosyltransferase subdomain (CAT). The locations of the three known automodification sites of PARP2 (S47, S76, and S281) are indicated by red uptick bars. (B) The mechanism of synthetic BRCA1/2 deficiency and PARPi. During DNA replication, single-strand DNA is vulnerable to breakage. When a single-strand break (SSB) occurs, the replication fork is stalled. PARP1/2 (orange) arrive at the site of the damage and recruit other repair factors (blue cloud). In the presence of PARP1/2 inhibitor (red hexagon), however, PARP1/2 are trapped at the damage site, inhibiting the restart of replication. As a result of unresolved replication stress, a double-strand break (DSB) can arise. Homologous recombination (HR) is one of the most faithful ways to repair a DSB. Cells with functioning BRCA1/2 can carry out HR and restore the genomic integrity, but cancer cells with defective BRCA1/2 cannot. Thus, synthetic lethality selectively kills HR-deficient cancer cells in conjunction with a PARPi.
Fig. 2.
Fig. 2.
Chemical mechanism of PARylation by PARP1/2. In an initiation reaction, the ADPR moiety of the NAD+ substrate is attached to an amino acid side chain on PARP1/2. Elongation occurs wherein additional ADPRs are added to the existing ADPR, resulting in chain lengths up to 200. Branching of the PAR chains can also occur at the site indicated by an arrow. Additionally, PARP1 can simply hydrolyze NAD+ to yield free ADPR and nicotinamide in a reaction known as treadmilling.
Fig. 3.
Fig. 3.
Summary of inhibition measurements for PARPi with PARP1 and PARP2 and comparison to cell-based data. (A) For PARP1, each reported IC50 and KI value is shown as a point and the line indicates the median value. Our reported measurements for all these inhibitors using a method that avoids the tight-binding limit problem are indicated by black arrows. (B) For PARP2, each reported IC50 and KI value is shown as a point and the line indicates the median value. (C) The ratio of the median IC50 value for BRCT+/+ vs. matched BRCT−/− cells is plotted against the KI as determined in ref. , since these values were determined with proper consideration of the tight-binding limit. All raw values for the points in these plots along with literature references can be found in SI Appendix, Tables S1 and S2.
Fig. 4.
Fig. 4.
(A) Distribution of potencies for triggering cell death in 1,000 different cell lines by olaparib. The leftmost column includes all 1,000 cell lines studied in the Genomics of Drug Sensitivity in Cancer. Subsequent columns to the right display cancer-specific cell types wherein the breast cancer cell lines are not predominantly BRCA−/−. The line indicates the median value. The data were obtained from https://www.cancerrxgene.org. (BF) Comparison of the potency of olaparib to other PARPi in a panel of 1,000 different cell types. The data were obtained from https://www.cancerrxgene.org. The IC50 values for each inhibitor were downloaded in an array and compared to the IC50 values for the other inhibitors for each cell line for which the database contained a value. Each data point represents the log IC50 value of a PARPi (B, trametinib; C, talazoparib; D, niraparib; E, rucaparib; F, veliparib) vs. the log IC50 value of olaparib for a different cell line. To facilitate interpretation and comparisons between the different graphs, the axes are “square” (one log unit is the same length on both x and y axes), and the scales are the same for all five graphs. The black line represents a best-fit linear correlation between the two inhibitors and the Pearson correlation coefficient and slope value are indicated for each graph.
Fig. 5.
Fig. 5.
Overview of binding pocket for NAD+ in the catalytic domain of PARP1. (Left) The ribbon structure of the catalytic domain of PARP1 is shown in cyan with the nonhydrolyzable NAD+ analog (BAD) shown in green (Protein Data Bank [PDB] ID 6BHV). (Right) The pocket for binding BAD is visualized with the protein surface shown in cyan. Note that the nicotinamide ring sits in the deepest pocket of the active site.
Fig. 6.
Fig. 6.
(AF) Interaction diagrams for PARPi with PARP1. (A) Nicotinamide ring from BAD (PDB ID 6BHV). (B) Olaparib (PDB ID 7KK4). (C) Talazoparib (PDB ID 7KK3). (D) Rucaparib (PDB ID 6VVK). (E) Niraparib (PDB ID 7KK5). (F) Veliparib (PDB ID 7KK6). For simplicity and easier comparison with the PARPi, only the nicotinamide end of BAD is shown here. The full interaction diagram for BAD is included in SI Appendix, Fig. S1. All diagrams were prepared in ChemDraw using the deposited PDB entries indicated.

References

    1. Lodish H., et al. , Molecular Biology of the Cell (Freeman, New York, NY, 2016).
    1. Jeggo P. A., Pearl L. H., Carr A. M., DNA repair, genome stability and cancer: A historical perspective. Nat. Rev. Cancer 16, 35–42 (2016).
    1. Wood R. D., Mitchell M., Sgouras J., Lindahl T., Human DNA repair genes. Science 291, 1284–1289 (2001).
    1. Gibson B. A., Kraus W. L., New insights into the molecular and cellular functions of poly(ADP-ribose) and PARPs. Nat. Rev. Mol. Cell Biol. 13, 411–424 (2012).
    1. Bai P., Biology of poly(ADP-ribose) polymerases: The factotums of cell maintenance. Mol. Cell 58, 947–958 (2015).
    1. Langelier M. F., Eisemann T., Riccio A. A., Pascal J. M., PARP family enzymes: Regulation and catalysis of the poly(ADP-ribose) posttranslational modification. Curr. Opin. Struct. Biol. 53, 187–198 (2018).
    1. Suskiewicz M. J., Palazzo L., Hughes R., Ahel I., Progress and outlook in studying the substrate specificities of PARPs and related enzymes. FEBS J. 288, 2131–2142 (2020).
    1. Teloni F., Altmeyer M., Readers of poly(ADP-ribose): Designed to be fit for purpose. Nucleic Acids Res. 44, 993–1006 (2016).
    1. Krastev D. B., et al. , Coupling bimolecular PARylation biosensors with genetic screens to identify PARylation targets. Nat. Commun. 9, 2016 (2018).
    1. Dasovich M., et al. , Identifying poly(ADP-ribose)-binding proteins with photoaffinity-based proteomics. J. Am. Chem. Soc. 143, 3037–3042 (2021).
    1. Wang M., et al. , PaxDb, a database of protein abundance averages across all three domains of life. Mol. Cell. Proteomics 11, 492–500 (2012).
    1. Franzese E., et al. , PARP inhibitors in ovarian cancer. Cancer Treat. Rev. 73, 1–9 (2019).
    1. Muthurajan U. M., et al. , Automodification switches PARP-1 function from chromatin architectural protein to histone chaperone. Proc. Natl. Acad. Sci. U.S.A. 111, 12752–12757 (2014).
    1. David K. K., Andrabi S. A., Dawson T. M., Dawson V. L., Parthanatos, a messenger of death. Front. Biosci. 14, 1116–1128 (2009).
    1. Amé J. C., et al. , PARP-2, a novel mammalian DNA damage-dependent poly(ADP-ribose) polymerase. J. Biol. Chem. 274, 17860–17868 (1999).
    1. Schreiber V., et al. , Poly(ADP-ribose) polymerase-2 (PARP-2) is required for efficient base excision DNA repair in association with PARP-1 and XRCC1. J. Biol. Chem. 277, 23028–23036 (2002).
    1. Beck C., Robert I., Reina-San-Martin B., Schreiber V., Dantzer F., Poly(ADP-ribose) polymerases in double-strand break repair: Focus on PARP1, PARP2 and PARP3. Exp. Cell Res. 329, 18–25 (2014).
    1. Chen Q., Kassab M. A., Dantzer F., Yu X., PARP2 mediates branched poly ADP-ribosylation in response to DNA damage. Nat. Commun. 9, 3233 (2018).
    1. Barkauskaite E., Jankevicius G., Ahel I., Structures and mechanisms of enzymes employed in the synthesis and degradation of PARP-dependent protein ADP-ribosylation. Mol. Cell 58, 935–946 (2015).
    1. Hoch N. C., Polo L. M., ADP-ribosylation: From molecular mechanisms to human disease. Genet. Mol. Biol. 43 (suppl. 1), e20190075 (2019).
    1. Dawicki-McKenna J. M., et al. , PARP-1 activation requires local unfolding of an autoinhibitory domain. Mol. Cell 60, 755–768 (2015).
    1. Eustermann S., et al. , Structural basis of detection and signaling of DNA single-strand breaks by human PARP-1. Mol. Cell 60, 742–754 (2015).
    1. Rudolph J., Mahadevan J., Luger K., Probing the conformational changes associated with DNA binding to PARP1. Biochemistry 59, 2003–2011 (2020).
    1. Rudolph J., et al. , The BRCT domain of PARP1 binds intact DNA and mediates intrastrand transfer. Mol. Cell 81, 4994–5006.e5 (2021).
    1. Obaji E., Haikarainen T., Lehtiö L., Characterization of the DNA dependent activation of human ARTD2/PARP2. Sci. Rep. 6, 34487 (2016).
    1. Bilokapic S., Suskiewicz M. J., Ahel I., Halic M., Bridging of DNA breaks activates PARP2-HPF1 to modify chromatin. Nature 585, 609–613 (2020).
    1. Gaullier G., et al. , Bridging of nucleosome-proximal DNA double-strand breaks by PARP2 enhances its interaction with HPF1. PLoS One 15, e0240932 (2020).
    1. Farmer H., et al. , Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature 434, 917–921 (2005).
    1. Bryant H. E., et al. , Specific killing of BRCA2-deficient tumours with inhibitors of poly(ADP-ribose) polymerase. Nature 434, 913–917 (2005).
    1. Bridges C. B., The origin of variations in sexual and sex-limited characters. Am. Nat. 56, 51–63 (1922).
    1. Dobzhansky T., Genetics of natural populations; recombination and variability in populations of Drosophila pseudoobscura. Genetics 31, 269–290 (1946).
    1. Hartwell L. H., Szankasi P., Roberts C. J., Murray A. W., Friend S. H., Integrating genetic approaches into the discovery of anticancer drugs. Science 278, 1064–1068 (1997).
    1. Kuchenbaecker K. B., et al. ; BRCA1 and BRCA2 Cohort Consortium, Risks of breast, ovarian, and contralateral breast cancer for BRCA1 and BRCA2 mutation carriers. JAMA 317, 2402–2416 (2017).
    1. Tai Y. C., Domchek S., Parmigiani G., Chen S., Breast cancer risk among male BRCA1 and BRCA2 mutation carriers. J. Natl. Cancer Inst. 99, 1811–1814 (2007).
    1. Han J., Huang J., DNA double-strand break repair pathway choice: The fork in the road. Genome Instab. Dis. 1, 10–19 (2020).
    1. Trego K. S., et al. , Non-catalytic roles for XPG with BRCA1 and BRCA2 in homologous recombination and genome stability. Mol. Cell 61, 535–546 (2016).
    1. Powell S. N., Kachnic L. A., Roles of BRCA1 and BRCA2 in homologous recombination, DNA replication fidelity and the cellular response to ionizing radiation. Oncogene 22, 5784–5791 (2003).
    1. Scully R., Livingston D. M., In search of the tumour-suppressor functions of BRCA1 and BRCA2. Nature 408, 429–432 (2000).
    1. Venkitaraman A. R., Cancer susceptibility and the functions of BRCA1 and BRCA2. Cell 108, 171–182 (2002).
    1. Evers B., Helleday T., Jonkers J., Targeting homologous recombination repair defects in cancer. Trends Pharmacol. Sci. 31, 372–380 (2010).
    1. Helleday T., The underlying mechanism for the PARP and BRCA synthetic lethality: Clearing up the misunderstandings. Mol. Oncol. 5, 387–393 (2011).
    1. Curtin N. J., Szabo C., Poly(ADP-ribose) polymerase inhibition: Past, present and future. Nat. Rev. Drug Discov. 19, 711–736 (2020).
    1. Vikas P., Borcherding N., Chennamadhavuni A., Garje R., Therapeutic potential of combining PARP inhibitor and immunotherapy in solid tumors. Front. Oncol. 10, 570 (2020).
    1. Spiegel J. O., Van Houten B., Durrant J. D., PARP1: Structural insights and pharmacological targets for inhibition. DNA Repair (Amst.) 103, 103125 (2021).
    1. Pommier Y., O’Connor M. J., de Bono J., Laying a trap to kill cancer cells: PARP inhibitors and their mechanisms of action. Sci. Transl. Med. 8, 362ps17 (2016).
    1. Zhao Y., et al. , The ups and downs of poly(ADP-ribose) polymerase-1 inhibitors in cancer therapy: Current progress and future direction. Eur. J. Med. Chem. 203, 112570 (2020).
    1. Yi M., et al. , Advances and perspectives of PARP inhibitors. Exp. Hematol. Oncol. 8, 29 (2019).
    1. Berger N. A., et al. , Opportunities for the repurposing of PARP inhibitors for the therapy of non-oncological diseases. Br. J. Pharmacol. 175, 192–222 (2018).
    1. Noordermeer S. M., van Attikum H., PARP inhibitor resistance: A tug-of-war in BRCA-mutated cells. Trends Cell Biol. 29, 820–834 (2019).
    1. Crawford K., Bonfiglio J. J., Mikoč A., Matic I., Ahel I., Specificity of reversible ADP-ribosylation and regulation of cellular processes. Crit. Rev. Biochem. Mol. Biol. 53, 64–82 (2018).
    1. Alvarez-Gonzalez R., Jacobson M. K., Characterization of polymers of adenosine diphosphate ribose generated in vitro and in vivo. Biochemistry 26, 3218–3224 (1987).
    1. Desmarais Y., Ménard L., Lagueux J., Poirier G. G., Enzymological properties of poly(ADP-ribose)polymerase: Characterization of automodification sites and NADase activity. Biochim. Biophys. Acta 1078, 179–186 (1991).
    1. Murai J., et al. , Trapping of PARP1 and PARP2 by clinical PARP inhibitors. Cancer Res. 72, 5588–5599 (2012).
    1. Murai J., et al. , Stereospecific PARP trapping by BMN 673 and comparison with olaparib and rucaparib. Mol. Cancer Ther. 13, 433–443 (2014).
    1. Kurgina T. A., Anarbaev R. O., Sukhanova M. V., Lavrik O. I., A rapid fluorescent method for the real-time measurement of poly(ADP-ribose) polymerase 1 activity. Anal. Biochem. 545, 91–97 (2018).
    1. Rudolph J., Roberts G., Muthurajan U. M., Luger K., HPF1 and nucleosomes mediate a dramatic switch in activity of PARP1 from polymerase to hydrolase. eLife 10, e65773 (2021).
    1. Holtlund J., Jemtland R., Kristensen T., Two proteolytic degradation products of calf-thymus poly(ADP-ribose) polymerase are efficient ADP-ribose acceptors. Implications for polymerase architecture and the automodification of the polymerase. Eur. J. Biochem. 130, 309–314 (1983).
    1. Bauer P. I., Hakam A., Kun E., Mechanisms of poly(ADP-ribose) polymerase catalysis; mono-ADP-ribosylation of poly(ADP-ribose) polymerase at nanomolar concentrations of NAD. FEBS Lett. 195, 331–338 (1986).
    1. Alvarez-Gonzalez R., 3′-Deoxy-NAD+ as a substrate for poly(ADP-ribose)polymerase and the reaction mechanism of poly(ADP-ribose) elongation. J. Biol. Chem. 263, 17690–17696 (1988).
    1. Ménard L., Thibault L., Poirier G. G., Reconstitution of an in vitro poly(ADP-ribose) turnover system. Biochim. Biophys. Acta 1049, 45–58 (1990).
    1. Ruf A., de Murcia G., Schulz G. E., Inhibitor and NAD+ binding to poly(ADP-ribose) polymerase as derived from crystal structures and homology modeling. Biochemistry 37, 3893–3900 (1998).
    1. Kawaichi M., Ueda K., Hayaishi O., Multiple autopoly(ADP-ribosyl)ation of rat liver poly(ADP-ribose) synthetase. Mode of modification and properties of automodified synthetase. J. Biol. Chem. 256, 9483–9489 (1981).
    1. Cheung A., Zhang J., A scintillation proximity assay for poly(ADP-ribose) polymerase. Anal. Biochem. 282, 24–28 (2000).
    1. Penning T. D., et al. , Discovery and SAR of 2-(1-propylpiperidin-4-yl)-1H-benzimidazole-4-carboxamide: A potent inhibitor of poly(ADP-ribose) polymerase (PARP) for the treatment of cancer. Bioorg. Med. Chem. 16, 6965–6975 (2008).
    1. Putt K. S., Hergenrother P. J., An enzymatic assay for poly(ADP-ribose) polymerase-1 (PARP-1) via the chemical quantitation of NAD(+): Application to the high-throughput screening of small molecules as potential inhibitors. Anal. Biochem. 326, 78–86 (2004).
    1. Jarmoskaite I., AlSadhan I., Vaidyanathan P. P., Herschlag D., How to measure and evaluate binding affinities. eLife 9, e57264 (2020).
    1. Rudolph J., Roberts G., Luger K., Histone parylation factor 1 contributes to the inhibition of PARP1 by cancer drugs. Nat. Commun. 12, 736 (2021).
    1. Hopkins T. A., et al. , Mechanistic dissection of PARP1 trapping and the impact on in vivo tolerability and efficacy of PARP inhibitors. Mol. Cancer Res. 13, 1465–1477 (2015).
    1. Hopkins T. A., et al. , PARP1 trapping by PARP inhibitors drives cytotoxicity in both cancer cells and healthy bone marrow. Mol. Cancer Res. 17, 409–419 (2019).
    1. Goldstein B., Coombs D., He X., Pineda A. R., Wofsy C., The influence of transport on the kinetics of binding to surface receptors: Application to cells and BIAcore. J. Mol. Recognit. 12, 293–299 (1999).
    1. Shen Y., et al. , BMN 673, a novel and highly potent PARP1/2 inhibitor for the treatment of human cancers with DNA repair deficiency. Clin. Cancer Res. 19, 5003–5015 (2013).
    1. Lord C. J., Ashworth A., In the clinic. Science 355, 1152–1158 (2017).
    1. Yang W., et al. , Genomics of drug sensitivity in cancer (GDSC): A resource for therapeutic biomarker discovery in cancer cells. Nucleic Acids Res. 41, D955–D961 (2013).
    1. Garnett M. J., et al. , Systematic identification of genomic markers of drug sensitivity in cancer cells. Nature 483, 570–575 (2012).
    1. Jaspers J. E., et al. , Loss of 53BP1 causes PARP inhibitor resistance in Brca1-mutated mouse mammary tumors. Cancer Discov. 3, 68–81 (2013).
    1. Pilié P. G., Gay C. M., Byers L. A., O’Connor M. J., Yap T. A., PARP inhibitors: Extending benefit beyond BRCA-mutant cancers. Clin. Cancer Res. 25, 3759–3771 (2019).
    1. Zandarashvili L., et al. , Structural basis for allosteric PARP-1 retention on DNA breaks. Science 368, eaax6367 (2020).
    1. Satoh M. S., Lindahl T., Role of poly(ADP-ribose) formation in DNA repair. Nature 356, 356–358 (1992).
    1. Kedar P. S., Stefanick D. F., Horton J. K., Wilson S. H., Increased PARP-1 association with DNA in alkylation damaged, PARP-inhibited mouse fibroblasts. Mol. Cancer Res. 10, 360–368 (2012).
    1. Rudolph J., Mahadevan J., Dyer P., Luger K., Poly(ADP-ribose) polymerase 1 searches DNA via a ‘monkey bar’ mechanism. eLife 7, e37818 (2018).
    1. Krüger A., Bürkle A., Hauser K., Mangerich A., Real-time monitoring of PARP1-dependent PARylation by ATR-FTIR spectroscopy. Nat. Commun. 11, 2174 (2020).
    1. Prokhorova E., et al. , Serine-linked PARP1 auto-modification controls PARP inhibitor response. Nat. Commun. 12, 4055 (2021).
    1. Purnell M. R., Whish W. J. D., Novel inhibitors of poly(ADP-ribose) synthetase. Biochem. J. 185, 775–777 (1980).
    1. Langelier M.-F., Zandarashvili L., Aguiar P. M., Black B. E., Pascal J. M., NAD+ analog reveals PARP-1 substrate-blocking mechanism and allosteric communication from catalytic center to DNA-binding domains. Nat. Commun. 9, 844 (2018).
    1. Marsischky G. T., Wilson B. A., Collier R. J., Role of glutamic acid 988 of human poly-ADP-ribose polymerase in polymer formation. Evidence for active site similarities to the ADP-ribosylating toxins. J. Biol. Chem. 270, 3247–3254 (1995).
    1. Thorsell A. G., et al. , Structural basis for potency and promiscuity in poly(ADP-ribose) polymerase (PARP) and tankyrase inhibitors. J. Med. Chem. 60, 1262–1271 (2017).
    1. Gibbs-Seymour I., Fontana P., Rack J. G. M., Ahel I., HPF1/C4orf27 is a PARP-1-interacting protein that regulates PARP-1 ADP-ribosylation activity. Mol. Cell 62, 432–442 (2016).
    1. Wahlberg E., et al. , Family-wide chemical profiling and structural analysis of PARP and tankyrase inhibitors. Nat. Biotechnol. 30, 283–288 (2012).
    1. Vyas S., et al. , Family-wide analysis of poly(ADP-ribose) polymerase activity. Nat. Commun. 5, 4426 (2014).
    1. Bonfiglio J. J., et al. , Serine ADP-ribosylation depends on HPF1. Mol. Cell 65, 932–940.e6 (2017).
    1. Palazzo L., et al. , Serine is the major residue for ADP-ribosylation upon DNA damage. eLife 7, 34334 (2018).
    1. Suskiewicz M. J., et al. , HPF1 completes the PARP active site for DNA damage-induced ADP-ribosylation. Nature 579, 598–602 (2020).
    1. Sun F. H., et al. , HPF1 remodels the active site of PARP1 to enable the serine ADP-ribosylation of histones. Nat. Commun. 12, 1028 (2021).
    1. Brustel J., et al. , Linking DNA repair and cell cycle progression through serine ADP-ribosylation of histones. Nat. Commun. 13, 185 (2022).
    1. Zhou P., Wang J., Mishail D., Wang C. Y., Recent advancements in PARP inhibitors-based targeted cancer therapy. Precis. Clin. Med. 3, 187–201 (2020).
    1. Mateo J., et al. , DNA-repair defects and olaparib in metastatic prostate cancer. N. Engl. J. Med. 373, 1697–1708 (2015).
    1. Zimmermann M., et al. , CRISPR screens identify genomic ribonucleotides as a source of PARP-trapping lesions. Nature 559, 285–289 (2018).
    1. Fugger K., et al. , Targeting the nucleotide salvage factor DNPH1 sensitizes BRCA-deficient cells to PARP inhibitors. Science 372, 156–165 (2021).
    1. Hewitt G., et al. , Defective ALC1 nucleosome remodeling confers PARPi sensitization and synthetic lethality with HRD. Mol. Cell 81, 767–783.e11 (2021).
    1. Juhász S., et al. , The chromatin remodeler ALC1 underlies resistance to PARP inhibitor treatment. Sci. Adv. 6, eabb8626 (2020).
    1. Lo C. S. Y., et al. , SMARCAD1-mediated active replication fork stability maintains genome integrity. Sci. Adv. 7, eabe7804 (2021).
    1. Sinha G., Downfall of iniparib: A PARP inhibitor that doesn’t inhibit PARP after all. J. Natl. Cancer Inst. 106, djt447 (2014).

Source: PubMed

3
Předplatit