Next Article in Journal
Novel COX11 Mutations Associated with Mitochondrial Disorder: Functional Characterization in Patient Fibroblasts and Saccharomyces cerevisiae
Previous Article in Journal
Ceramides Mediate Insulin-Induced Impairments in Cerebral Mitochondrial Bioenergetics in ApoE4 Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A New Drug Discovery Platform: Application to DNA Polymerase Eta and Apurinic/Apyrimidinic Endonuclease 1

1
XPose Therapeutics, Inc., San Carlos, CA 94070, USA
2
Accelero Biostructures, Inc., San Carlos, CA 94070, USA
3
Mitchell Cancer Institute and Department of Pharmacology, University of South Alabama, Mobile, AL 36604, USA
4
Department of Pathology & Laboratory Medicine, Warrant Alpert Medical School & Legorreta Cancer Center, Brown University, Providence, RI 02912, USA
5
Department of Biochemistry and Molecular Biology, Indiana University School of Medicine, Indianapolis, IN 46202, USA
6
Mid-Atlantic BioTherapeutics, Inc., Doylestown, PA 18902, USA
7
Department of BioSciences, Rice University, Houston, TX 77251, USA
8
Biomedical Research Institute, Hasselt University, 3500 Diepenbeek, Belgium
9
Belgium & Boost Scientific, 3550 Heusden-Zolder, Belgium
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(23), 16637; https://doi.org/10.3390/ijms242316637
Submission received: 29 August 2023 / Revised: 26 October 2023 / Accepted: 27 October 2023 / Published: 23 November 2023
(This article belongs to the Section Molecular Genetics and Genomics)

Abstract

:
The ability to quickly discover reliable hits from screening and rapidly convert them into lead compounds, which can be verified in functional assays, is central to drug discovery. The expedited validation of novel targets and the identification of modulators to advance to preclinical studies can significantly increase drug development success. Our SaXPyTM (“SAR by X-ray Poses Quickly”) platform, which is applicable to any X-ray crystallography-enabled drug target, couples the established methods of protein X-ray crystallography and fragment-based drug discovery (FBDD) with advanced computational and medicinal chemistry to deliver small molecule modulators or targeted protein degradation ligands in a short timeframe. Our approach, especially for elusive or “undruggable” targets, allows for (i) hit generation; (ii) the mapping of protein–ligand interactions; (iii) the assessment of target ligandability; (iv) the discovery of novel and potential allosteric binding sites; and (v) hit-to-lead execution. These advances inform chemical tractability and downstream biology and generate novel intellectual property. We describe here the application of SaXPy in the discovery and development of DNA damage response inhibitors against DNA polymerase eta (Pol η or POLH) and apurinic/apyrimidinic endonuclease 1 (APE1 or APEX1). Notably, our SaXPy platform allowed us to solve the first crystal structures of these proteins bound to small molecules and to discover novel binding sites for each target.

1. Introduction

Cancer will directly affect the lives of over one-third of the global population. The process of carcinogenesis involves (at least) six biological hallmarks [1]: sustaining proliferative signaling, evading growth suppressors, resisting cell death, enabling replicative immortality, inducing angiogenesis, and activating invasion and metastasis. Many, if not all, of these hallmarks can be attributed to genomic instability that arises from excessive DNA damage or defects in the DNA damage response (DDR) components. DDR is an intricate system, which involves damage recognition, DNA damage repair, cell cycle regulation, and cell fate determination, playing a prominent role in cancer etiology and therapy [2]. DDR pathways are frequently upregulated in cancer cells as a compensatory mechanism to adapt to elevated background levels of DNA damage from rapid cell division and increased metabolism [3] or genotoxic stress induced by many anti-cancer agents, including radiotherapy and certain forms of chemotherapy [4,5]. The realization that such intrinsic changes in the DDR (i.e., sporadic inactivation or upregulation) offer therapeutic opportunities has led to advances in cancer treatment efficacy.
The discovery that homologous recombination repair (HRR)-defective breast and ovarian cancers are uniquely sensitive to poly-(ADP-ribose) polymerase (PARP) inhibitors has driven the renewed interest in the design of DDR inhibitors to combat both intrinsic, cancer-specific characteristics and acquired drug resistance [6,7]. Exploiting such so-called synthetic lethality (SL), where PARP inhibitors drive the accumulation of cytotoxic DNA intermediates that are normally resolved via HRR, has motivated improved drug design/application and has led to better outcomes for many of these cancer-affected individuals [8]. Moreover, developed drugs will widen the repertoire of initial treatment options and have potential utility in re-sensitizing cells to genotoxic therapies that have failed due to the upregulation of DDR pathways. Thus, several efforts around the world are focused on the development of novel small molecule inhibitor(s) (SMI) or targeted protein degradation (TPD) cancer therapeutics against DDR proteins, with an eye on SL and combinatorial treatment opportunities [9]. Here, we give a brief overview of some current drug discovery efforts before presenting a novel strategy, termed SaXPy (SAR by X-ray Poses Quickly), and its application to two DDR proteins, DNA polymerase eta (Pol η or POLH) and apurinic/apyrimidinic endonuclease 1 (APE1 or APEX1).

2. Overview of Drug Discovery

Early drug discovery towards preclinical studies includes one or more of the following components: hit generation, hit-to-lead expansion, and lead optimization. Hit generation involves screening drug targets against compound libraries (e.g., fragment, scaffold, small molecule libraries, and DELs—DNA-Encoded Libraries) using different methods, the conventional ones being biochemical screening (HTS—High Throughput Screening), biophysical screening (NMR—Nuclear Magnetic Resonance; SPR—Surface Plasmon Resonance; and TSA—Thermal Shift Assay), and computational screening (also known as virtual or in silico screening) [10,11,12,13]. Hit-to-lead expansion and lead optimization involve cycles of medicinal, computational, and synthetic chemistry interwoven with interactive cycles of validating biochemical, biophysical, and cellular assays, as well as structural biology for structure-based drug discovery (SBDD) [14,15], amongst other discovery activities (e.g., ADME/PK, in vivo pharmacology, toxicology, and formulation activities). In many cases, biopharmaceutical companies have also built their own proprietary drug discovery platforms incorporating one or more of these foundational methods with new innovations built on top.

3. SaXPy Platform

The SaXPy platform is designed to steer a fragment-based drug discovery (FBDD) and SBDD program and encompasses two technological components: (i) hit generation via high-throughput X-ray crystallography-based screening of a fragment library, one fragment at a time, and (ii) hit-to-lead expansion using computational and medicinal chemistry with analog scoping, scaffold hopping, and fragment growth. Directly using high-throughput X-ray crystallography as a primary screen allows one to immediately assess target engagement via the direct visualization of the hit(s) binding site, binding pose, and protein–ligand interactions. Such information guides the quick advances of hits to functional lead compounds. The SaXPy platform is coupled to tests of in vitro functional activity using target-specific in vitro biochemical assays, guiding the identification of the most effective molecules. Validated compounds obtained using the SaXPy strategy can then be used either as SMIs or as ligands for TPD drugs that target either an orthosteric or non-orthosteric/allosteric site. We describe herein the principles of the SaXPy platform and provide an overview of the first crystal structures of APE1 and POLH bound to small-molecule drug-like fragments, highlighting some of the critical parts of our approach and the results obtained that have enabled us to pursue novel drug development.

4. POLH Background

POLH is a member of the Y family of DNA polymerases [16,17,18]. It is a translesion DNA polymerase that can bypass certain blocking lesions, such as those generated via ultraviolet radiation (UVR) or cisplatin, and it is deployed to replicate foci for translesion synthesis (TLS) as part of the DDR. Inherited defects in the gene encoding POLH (XPV) are associated with the rare, sun-sensitive, cancer-prone disorder, xeroderma pigmentosum, due to the loss of the ability of POLH to accurately bypass UVR-induced thymine dimers. In standard-of-care cancer therapies that involve platinum-based clinical agents, e.g., cisplatin or oxaliplatin, POLH can also bypass platinum-DNA adducts, negating the benefits of the treatment and enabling drug resistance [19,20,21,22,23,24,25,26,27]. Moreover, POLH has been implicated in resistance to nucleoside analogs, such as gemcitabine and cytarabine [28], and other studies suggest that POLH plays an important role in oxidative stress resistance, likely by carrying out the TLS [29,30] of bulky oxidative base lesions, such as cyclopurines [31,32,33]. Beyond its canonical TLS functions, recent studies found that POLH is important for resistance against the chemotherapeutic alkylating agent, temozolomide (TMZ), seemingly via a TLS-independent mechanism [34]. Finally, POLH has been shown to exhibit the capacity to promote the RNA-templated error-free repair of DNA double-strand breaks (DSBs), a finding that has important implications for carcinogenesis and cancer therapy considerations [35].
POLH overexpression has been linked to the development of chemoresistance in several cancers, including lung, ovarian, and bladder cancers [16]. Consistent with the known biochemistry, elevated POLH expression is correlated with reduced cisplatin sensitivity in models of lung and bladder cancer [19]. The strategic downregulation of POLH in these cases re-sensitizes cancer cells to cisplatin treatment, supporting the targeting of the polymerase in certain situations of acquired drug resistance. The suppression of POLH expression also enhances the cisplatin-induced apoptosis of cancer stem cells isolated from both ovarian cancer cell lines and primary tumors [21]. Furthermore, studies indicate that POLH is a predictive factor for treatment response and the survival of metastatic gastric adenocarcinoma patients receiving oxaliplatin-based first-line chemotherapy [36]. In addition to its well-established role in platin drug resistance, preclinical studies indicate that POLH-deficient cells are 3-fold more sensitive to the nucleoside analogs, B-D-arabinofuranosylcytosine and gemcitabine. POLH-deficient cells are even more sensitive (10-fold) to gemcitabine/cisplatin combination treatment [37], which is a commonly used clinical regimen for treating a wide spectrum of cancers, including bladder, pancreatic, ovarian, cervical, and non-small-cell lung cancers. Additional investigations have revealed that the co-inhibition of POLH and ATR, a protein that is central to the replicative stress response, offers an SL approach for the treatment of a range of cancer types [38,39]. Notably, ATR inhibitors are progressing well in the clinic [40,41,42]. ATR haploinsufficiency, arising due to somatic mutations in one allele, is frequent in certain cancers [43], presenting therapeutic opportunities for POLH inhibition. Very recent work involving functional screening has implicated POLH as an important target for the discovery and development of inhibitors in HORMAD1-positive triple-negative breast cancer (TNBC) [44], a defined disease group that comprises ~60% of TNBC cases, further encouraging the pursuit of POLH inhibitors.
With the value of targeting POLH in the context of new oncology therapeutics, it is not surprising that some attempts have been made in this direction to develop POLH inhibitors: compounds derived from N-aryl-substituted indole barbituric acid (IBA), indole thiobarbituric acid (ITBA), and indole quinuclidine scaffolds [20,45], which are predicted to interfere with template DNA orientation; the computer-aided discovery of a novel class of chromone analogs [46]; and the design of amino acid- and carbohydrate-based compounds [47]. However, these compounds have yet to advance further, and our assessment based on the information available is that this could be due to (i) the precise target engagement/hit validation being unknown due to the absence of crystal structures, preventing further interaction-based optimization, and/or (ii) the suitability of these compounds for further chemistry tractability/optimization.

5. APE1 Background

APE1 (also known as APEX1) is a multifunctional protein with a primary function as a DNA repair nuclease and has a separate function as a regulator of transcription factor DNA binding via a redox mechanism [48,49]. As an apurinic/apyrimidinic (AP) endonuclease, APE1 is a central player in the base excision repair (BER) pathway, a process involved in resolving spontaneous, alkylative, and oxidative DNA damage [50]. Specifically, APE1 cleaves at AP sites generated spontaneously either via damage induction or through the action of DNA glycosylases, which initiate classic BER by excising a substrate (often damaged) base moiety [51]. Following the APE1-directed incision event, the remaining 5′-linked abasic fragment is removed, the gap is filled, and the nick is sealed [52]. Due to its prominent role in BER, and since rapidly proliferating cancer cells often upregulate DNA repair enzymes, such as APE1, the protein has emerged as a promising anti-cancer target [53].
Every day, more than 10,000 AP sites are created in each cell under normal metabolic conditions [54], and treatment with some chemotherapeutic agents, as well as ionizing radiation, increases the total number of genomic abasic lesions. If left unrepaired, non-coding AP sites can lead to mutations or replication fork collapse and DSBs that are highly cytotoxic [55]. The importance of APE1 in BER is highlighted by the facts that the enzyme is responsible for greater than 95% of the total AP endonuclease activity in human cells [56] and that the depletion of the protein leads to mammalian cell inviability [57,58], with some selectively for cancer cells [59]. In addition, the siRNA knockdown of APE1 increases the sensitivity of cells to several DNA-damaging agents, most notably alkylators, such as methyl methanesulfonate (MMS) [60,61,62,63]. Thus, APE1 is potentially a relevant therapeutic target for a number of cancers and has been validated in a xenograft model for ovarian cancer, with the knockdown of APE1 greatly reducing tumor growth [64].
APE1 may be particularly relevant in glioblastoma (GBM), for which the TMZ alkylator is the preferred chemotherapeutic agent, as the resistance to TMZ is in part mediated by elevated levels of APE1 [61,65,66,67,68,69,70]. A recent study on exceptional responders to TMZ treatment in GBM, following surgery and radiation, identified the inactivation of APE1 as a source for the exceptionally strong treatment response over 10 years, thus highlighting the promise of APE1 inhibitors in GBM treatment with TMZ [71]. The strategy of combining a DNA-damaging agent with a BER inhibitor, i.e., TMZ, with the PARP inhibitor, veliparib, has been explored in a phase I trial for the treatment of acute myeloid leukemia [72], showing that this combination treatment is well tolerated with efficacy in advanced acute myeloid leukemia. An even more effective strategy might be to employ an APE1 selective inhibitor with TMZ. In addition to the clinical potential of targeting APE1 in combinatorial therapies, it has been demonstrated that APE1 nuclease inhibitors induce the SL of cancer cells that are deficient in DNA DSB repair, i.e., BRCA or ataxia telangiectasia mutated (ATM) defective cell models [73]. The inhibition of APE1 results in the accumulation of DSBs and G2/M cell cycle arrest, presumably due to AP site accumulation and replication fork collapse, ultimately leading to genomic instability and cell death. A similar SL relationship has been observed with APE1 inactivation in PTEN-deficient melanoma cells [74], which likely suffer from analogous DSB repair defects. Another recent study has shown that APE1 inhibition sensitizes cells to inhibitors of DNA-dependent Protein Kinase (DNA-PK) and ATM- and Rad3-related (ATR) kinase, which are both DDR targets of very high therapeutic interest [75]. Thus, clinically effective inhibitors against APE1 nuclease activity have the potential to be used in combination with or as an alternative to DDR inhibitors in the treatment of a range of select DNA repair-deficient cancers. Interestingly, another protein, APE2, which has similarities to APE1 in terms of structure and function, although with a low sequence identity, has also garnered interest as a target for novel DDR therapeutics [76,77,78]. Thus, efforts for the discovery of new and effective APE1 inhibitors can also serve as a foundation for the discovery and development of APE2 inhibitors.
Efforts to date to identify high-affinity, selective small molecule APE1 inhibitors have been met with limited success as evidenced by the few follow-up studies that were conducted to improve the compounds. The published approaches have relied primarily on (i) screening commercially available compounds that were synthesized for other molecular targets, (ii) computational screening, and (iii) pharmacophore modeling [73,74,79,80,81,82,83,84,85,86,87,88,89]. The limited success of these efforts likely stems in part from the inherent difficulty in targeting a protein–nucleic acid interaction [90]. In the case of APE1, the DNA binding site is large, accommodating nine base pairs of duplex DNA, and polar [91,92]. Thus, a new approach is warranted for the discovery of novel and selective APE1 inhibitors.

6. SaXPy Platform and POLH/APE1 Hit-Bound Crystal Structures

Prior to our efforts reported herein, there were no available crystal structures of small molecules bound to POLH or APE1 to facilitate SBDD or compound optimization, and there were no experimental structural data on the binding sites or protein–ligand interactions from other published reports of screening and inhibitor development. Towards that end, we have employed our proprietary SaXPy platform, which entails the following steps: (1) isolation of crystallography-grade recombinant protein; (2) optimization of crystallization conditions; (3) fragment library screening, one fragment at a time; (4) determination of fragment-bound protein crystal structures; (5) mapping of the fragment-protein interaction; and (6) hit optimization through medicinal chemistry.
Using previously described protocols [93,94], POLH and APE1 recombinant proteins were purified for X-ray crystallography-based library screening and downstream bioassays. Initial hits were identified by screening a diverse fragment library in a single step using the ABS-OneStepTM crystallography-based library screening platform (Accelero Biostructures, San Carlos, CA, USA) [9] through a partnership. In brief, crystals of apo-POLH-DNA binary complex and apo-APE1 were generated [32,94], employing optimized conditions for reproducibility and scalability to generate several hundred crystals of uniform quality for library screening via ultra-high-throughput X-ray crystallography. Approximately 300 crystals of POLH-DNA or APE1 were used to screen the ABS-Real300TM (Accelero Biostructures, proprietary) 300-fragment library, 1 fragment at a time, typically at a ~10–20 mM fragment concentration and a ~2–4 h soak time. A total of approximately 300 individual X-ray diffraction data sets for each protein target were collected at 100 K using a Pilatus 6M detector (Dectris) at SSRL on beamline 9-2 and the BLU-ICE [95] data collection environment. The data were then processed within the ABS-OneStep platform using XDS [96] and CCP4 [97], with structure determination performed via molecular replacement using our high-resolution apo-POLH-DNA binary complex (1.5Å resolution, which was refined to a crystallographic R/Rfree of ~13/19%) and apo-APE1 structures (highest resolution of 1.35Å refined to an intermediate crystallographic R/Rfree of 18/20%) as the search templates. All crystal structures were in the 1.7–2.4 Å resolution range with reasonable crystallographic R/Rfree values. Due to intellectual property considerations and constraints at the time of publication, the disclosure of the high-resolution details of fragment binding sites and their engagement with POLH and APE1, chemical structures, and other specifics of fragment growth are not shown at this time.
POLH Inhibitor Development. For POLH, we obtained a screening hit rate of ~1.3%. Two distinct binding sites were identified on POLH, which we call Hit 1 and Hit 2. Hit 1 is proximal to the orthosteric site, i.e., the DNA binding groove near the polymerase active site. Hit 2 is distal to the orthosteric site and was previously found to be an allosteric site in an in vitro DNA polymerase biochemical assay panel [93,98]. The intrinsic knowledge afforded by our approach was quickly exploited in hit expansion and evolution to advance hits (Figure 1 and Figure 2). An initial small round of hit expansion and medicinal chemistry that included ~40 compounds resulted in inhibitors with a range of functional activity in the in vitro biochemical assay, leading to the rapid identification of inhibitors to advance to subsequent rounds of chemistry to generate a lead compound. A single round of Hit 1 evolution led to one compound with an IC50 that was improved by at least 10-fold (XPTx-0289; 230 µM, graph published previously) compared to the ~2 mM IC50 of Hit 1, as well as other backup compounds, i.e., about eight compounds with IC50 ~1–5 mM and one compound with an IC50 of ~8 mM. Hit 2 expansion led to the XPTx-0267 compound with an IC50 of ~2 mM. Importantly, our chemical matter is different from the traditional nucleoside analog-based compounds identified in targeting screens for DNA polymerases and are now being pursued in further optimization campaigns.
APE1 Inhibitor Development. For APE1, we obtained a screening hit rate of ~8%. Hit rates are target-dependent, and the higher hit rate in APE1 compared to POLH reflects differences in the protein size, shape, and solvent-accessible cavities where the ligands can bind. Similar to our POLH case, one of the APE1 binding sites (termed A site) is proximal to the orthosteric site or DNA binding groove (E site), which is where DNA strand incision takes place (Figure 3). We obtained over 20 hits and structures at the A site and show here the quality of 8 hits and corresponding electron density maps of the fragments (Figure 4). Further investigations are pending to understand if this novel binding site might be an allosteric site. Focusing on one of our E site hits, we performed a hit evolution and expansion to ~400 compounds. Using an in-house high-throughput in vitro biochemical assay for APE1 endonuclease activity based on [86], we identified two lead compounds, XPTx-0091 (Figure 5A) and XPTx-0387, with in vitro IC50s of ~500 nM and ~250 nM, respectively, as well as several weaker back-up compounds compared to the IC50 of the standard APE1 Inhibitor III (~2 µM). XPTx-0091 was also tested for specificity against Endonuclease IV, a non-homologous enzyme with shared AP endonuclease activity, and it was found to have essentially no non-specific activity in this secondary assay (Figure 5B). We then advanced XPTx-0091 to in vitro cell biology assays.
Cells overexpressing the BER-specific N-methylpurine DNA glycosylase (MPG), the enzyme preceding APE1 in BER, exhibit increased sensitivity to DNA alkylators such as MMS, which is likely due to the elevated production of toxic AP sites or DNA strand break repair intermediates. We tested the effects of XPTx-0091 in glioma cells overexpressing MPG (LN428/MPG) in the presence or absence of MMS (Figure 6). The LN428 cell line (originally provided by Ian Pollack, University of Pittsburgh, Pittsburgh, PA) is a glioblastoma-derived cell line that harbors a deletion in both p14ARF and p16, with mutations in the p53 gene [99,100]. Both the parental LN428 cell line and the derivative of LN428 modified for the elevated expression of MPG (LN428/MPG) were described previously [101]. To test the effects of the inhibitor, the LN428/MPG cells were seeded at a density of 1000 cells/well in 96-well tissue culture plates. The following day, the cells were counted using the Celigo S imaging cytometer’s (Nexcelom Bioscience, Lawrence, MA, USA) direct cell counting function. Wells with counts not matching the rest were excluded from the experiments. The wells were then treated with media alone or with media supplemented with 0.1 mM methyl methanesulfonate (MMS) and a dose range of XPTx-0091 (0.25 mM–10 mM) or DMSO as the control. Following 120 h of exposure, the cells were counted as above and graphed using GraphPad (Prism, V9) as the percent control vs. the DMSO-treated cells. XPTx-0091 enhances the cytotoxicity of MMS in a concentration-dependent manner, and this sensitization is observed at concentrations of the compound that are not cytotoxic (i.e., 1 µM; compare purple columns). These observations indicate that the inhibitors of APE1 will likely potentiate the activity of chemotherapeutic alkylators that promote AP site formation without causing broad toxicity.
To determine the molecular mechanism underlying the cytotoxicity observed in cells treated with XPTx-0091 (Figure 7), we measured the amount of DNA damage using the CometChip assay (Sykora et al. 2018). This assay measures DNA strand breaks (single and double) as well as alkaline-sensitive sites, e.g., AP lesions, which are converted to strand breaks via β-elimination. The migration of broken or fragmented DNA from the cell nucleus during an electrophoresis step generates a “tail” or “comet”, which can be visualized and quantified using imaging software. For these studies, LN428 and LN428/MPG cells were seeded at a density of 50,000 cells per well in 96-well tissue culture plates. The following day, the cells were treated with XPTx-0091 (5 mM and 10 mM) 30 min prior to +/− 0.5 mM MMS (LN428 cells) or +/− 0.25 mM MMS (LN428/MPG cells) exposure for 1 h. The cells in each well were then trypsinized and transferred to the CometChip apparatus and analyzed for DNA damage as previously detailed [102]. Consistent with our toxicity studies (see above), cells overexpressing MPG that are treated with XPTx-0091 and MMS exhibit a concentration-dependent increase in the DNA damage levels (Figure 7A), while the treatment of cells with XPTx-0091 alone does not result in elevated DNA damage. Additionally, in cells expressing extremely low levels of MPG (Figure 7B), MMS exposure has a lesser effect on the total AP site (DNA damage) levels, which are likely dependent on MPG-mediated base damage excision. Taken together, our data provide strong support for XPTx-0091 targeting APE1 in cells and enhancing alkylator-induced DNA damage and cytotoxicity—the intended mechanism of action of APE1 inhibitors.

7. Discussion

Given the success of PARP1 inhibitors in the treatment of certain HRR-defective cancers, fundamental and clinical researchers have been eyeing ways to exploit SL paradigms in the eradication of neoplastic disease. In light of the well-documented changes in the DDRs within cancer cells, coupled with the common use of genotoxins in therapeutic modalities, particular interest, including from a range of pharmaceutical companies, has revolved around the development of novel potent DDR inhibitors. FBDD provides significant benefits in the discovery and development of target-specific novel and active chemical matter, as evidenced by the advancement of several medicines to the clinic [103,104,105,106]. Additionally, SBDD approaches allow for a quick transition from initial hits to lead compounds, with the process being guided by structural insights intrinsic to the method. We describe herein how our SaXPy lead generation platform, which integrates X-ray crystallography-driven fragment library screening, can permit the quick conversion of initial hits to functional lead compounds from even weak hits. As SaXPy allows one to capture novel chemical matter during screening by using the widest detection range method (X-ray crystallography), the approach allows for the discovery of hits irrespective of where they lie on the potency spectrum from weak to strong, whereas a classic biochemical assay cannot typically pick them up, followed by efficient hit-to-lead progression via a structure-guided approach. Our approach simultaneously and uniquely provides experimental information in a high throughput about binding sites; binding poses; protein–ligand interactions; and the separation of hits at orthosteric or potentially allosteric sites and new binding hotspots.
As presented herein, we determined the first X-ray crystal structures of small, novel drug-like compounds bound to POLH and APE1, and rapidly progressed them from initial hit generation to early-stage lead compounds using our SaXPy platform. For APE1, we identified two lead compounds that are now entering lead optimization with in vitro IC50s of ~250 nM and ~500 nM. For POLH, while the measured potencies for XPTx-0289 (IC50 230 µM) and XPTx-0267 (2 mM) may appear to be low, such values, and even weaker values, are typical for starting hits in FBDD projects. Recent examples of programs successfully advancing fragments with initial low potencies (>2 mM Kd or IC50) include inhibitors against Cyclophilin D [107]; Mycobacterium tuberculosis InhA [108]; WDR5-Myc [109]; and our own DDR target, APE1. Notably, in our APE1 effort, the original fragment hit had undetectable activity in the target-specific biochemical assay. The rapid advancement of an initial hit to significantly improved congener inhibitors demonstrates the power of our SaXPy platform to rapidly execute hit-to-lead development campaigns in the design of target-specific inhibitors. Thus, we described herein that SaXPy is applicable to diverse targets, namely POLH and APE1, and that from a single, initial round of fragment growth and expansion, we can rapidly execute hit-to-lead conversion from the experimental knowledge that is intrinsic to the crystal structures.

Author Contributions

Conceptualization, D.D. and D.M.W.III; methodology, D.D., A.M.D., M.A.J.D., T.M.G., P.P., D.M.W.III, J.C., R.W.S., M.M.G., J.K.-U., C.C. and Y.G.; validation, D.D, A.M.D., M.A.J.D., P.P., D.M.W.III, J.C., R.W.S., J.K.-U. and Y.G.; formal analysis, D.D, A.M.D., M.A.J.D., P.P., D.M.W.III, J.C., R.W.S., J.K.-U. and Y.G.; investigation, D.D, A.M.D., M.A.J.D., P.P., J.C., R.W.S., J.K.-U. and Y.G.; resources, D.D, A.M.D., M.A.J.D., R.W.S., J.K.-U. and Y.G.; data curation, D.D, A.M.D., M.A.J.D., P.P., J.C., R.W.S., J.K.-U., C.C. and Y.G.; writing—original draft preparation, D.D. and D.M.W.III; writing—review and editing, D.D., D.M.W.III, A.M.D., M.A.J.D., T.M.G., P.P., J.C., R.W.S., M.M.G., J.K.-U., C.C., Y.G. and D.V.J.; visualization, D.D., A.M.D., M.A.J.D., P.P., J.C., R.W.S., J.K.-U., C.C. and Y.G.; supervision, D.D., A.M.D., M.A.J.D., P.P., R.W.S., J.K.-U. and Y.G.; project administration, D.D., A.M.D., M.A.J.D., P.P., R.W.S., J.K.-U. and Y.G.; funding acquisition, D.D., A.M.D., M.A.J.D., R.W.S. and Y.G. All authors have read and agreed to the published version of the manuscript.

Funding

The research included in this publication was supported by the National Center for Advancing Translational Sciences of the NIH under Award Number R43 TR001736 and the National Institute of General Medical Sciences of the NIH under Award Number R44 GM132796 to Accelero Biostructures Inc.; the National Cancer Institute of the NIH under Award Number R43 CA254552 to XPose Therapeutics, Inc.; and the Cancer Prevention & Research Institute of Texas (CPRIT) Award RR190046 and Welch Foundation Grant Number C-2033-20200401 to YG. CC was supported by a fellowship from the Houston Area Molecular Biophysics Program (NIH Grant No. T32 GM008280, Program Director Dr. Theodore Wensel). The research in the Sobol Lab on DNA repair, the analysis of DNA damage, and the impact of genotoxic exposure were funded by grants from the NIH (ES029518, CA148629, ES014811, ES028949, CA238061, CA236911, AG069740, and ES032522). Support was also provided by grants from the Breast Cancer Research Foundation of Alabama, from the Abraham A. Mitchell Distinguished Investigator Fund, from the Mitchell Cancer Institute Molecular & Metabolic Oncology Program Development Fund, and from the Legoretta Cancer Center Endowment Fund (to RWS).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Due to intellectual property considerations and constraints at the time of publication, the disclosure of the high-resolution details of fragment binding sites and their engagement with POLH and APE1, chemical structures, and other specifics of fragment growth are not shown at this time.

Acknowledgments

Synchrotron X-ray diffraction data collection and data processing were performed by Accelero Biostructures, Inc., California. Synchrotron data were collected at Stanford Synchrotron Radiation Lightsource (SSRL), SLAC National Accelerator Laboratory, Menlo Park, California. Use of the Stanford Synchrotron Radiation Lightsource (SSRL), SLAC National Accelerator Laboratory, was supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences under Contract No. DE-AC02-76SF00515. The SSRL Structural Molecular Biology Program is supported by the DOE Office of Biological and Environmental Research and by the National Institutes of Health, National Institute of General Medical Sciences (P41 GM103393). The contents of this publication are solely the responsibility of the authors and do not necessarily represent the official views of the NIGMS or NIH.

Conflicts of Interest

The authors Debanu Das, Matthew A. J. Duncton, Taxiarchis M. Georgiadis, Patricia Pellicena, Millie M. Georgiadis, John King-Underwood, David V. Jobes, Ashley M. Deacon, and David M. Wilson III received stock and cash compensation from XPose Therapeutics Inc. The authors Debanu Das and Ashley M. Deacon received stock and cash compensation from Accelero Biostructures Inc. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as potential conflicts of interest.

References

  1. Hanahan, D.; Weinberg, R.A. Hallmarks of Cancer: The next Generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [PubMed]
  2. Jackson, S.P.; Bartek, J. The DNA-Damage Response in Human Biology and Disease. Nature 2009, 461, 1071–1078. [Google Scholar] [CrossRef] [PubMed]
  3. O’Connor, M.J. Targeting the DNA Damage Response in Cancer. Mol. Cell 2015, 60, 547–560. [Google Scholar] [CrossRef]
  4. Madhusudan, S.; Hickson, I.D. DNA Repair Inhibition: A Selective Tumour Targeting Strategy. Trends Mol. Med. 2005, 11, 503–511. [Google Scholar] [CrossRef]
  5. Helleday, T.; Petermann, E.; Lundin, C.; Hodgson, B.; Sharma, R.A. DNA Repair Pathways as Targets for Cancer Therapy. Nat. Rev. Cancer 2008, 8, 193–204. [Google Scholar] [CrossRef]
  6. Bryant, H.E.; Schultz, N.; Thomas, H.D.; Parker, K.M.; Flower, D.; Lopez, E.; Kyle, S.; Meuth, M.; Curtin, N.J.; Helleday, T. Specific Killing of BRCA2-Deficient Tumours with Inhibitors of poly(ADP-Ribose) Polymerase. Nature 2005, 434, 913–917. [Google Scholar] [CrossRef] [PubMed]
  7. Farmer, H.; McCabe, N.; Lord, C.J.; Tutt, A.N.J.; Johnson, D.A.; Richardson, T.B.; Santarosa, M.; Dillon, K.J.; Hickson, I.; Knights, C.; et al. Targeting the DNA Repair Defect in BRCA Mutant Cells as a Therapeutic Strategy. Nature 2005, 434, 917–921. [Google Scholar] [CrossRef] [PubMed]
  8. Slade, D. PARP and PARG Inhibitors in Cancer Treatment. Genes Dev. 2020, 34, 360–394. [Google Scholar] [CrossRef]
  9. Wilson, D.M., 3rd; Deacon, A.M.; Duncton, M.A.J.; Pellicena, P.; Georgiadis, M.M.; Yeh, A.P.; Arvai, A.S.; Moiani, D.; Tainer, J.A.; Das, D. Fragment- and Structure-Based Drug Discovery for Developing Therapeutic Agents Targeting the DNA Damage Response. Prog. Biophys. Mol. Biol. 2020, 163, 130–142. [Google Scholar] [CrossRef]
  10. Shuker, S.B.; Hajduk, P.J.; Meadows, R.P.; Fesik, S.W. Discovering High-Affinity Ligands for Proteins: SAR by NMR. Science 1996, 274, 1531–1534. [Google Scholar] [CrossRef]
  11. Gossert, A.D.; Jahnke, W. NMR in Drug Discovery: A Practical Guide to Identification and Validation of Ligands Interacting with Biological Macromolecules. Prog. Nucl. Magn. Reson. Spectrosc. 2016, 97, 82–125. [Google Scholar] [CrossRef] [PubMed]
  12. Navratilova, I.; Hopkins, A.L. Fragment Screening by Surface Plasmon Resonance. ACS Med. Chem. Lett. 2010, 1, 44–48. [Google Scholar] [CrossRef]
  13. Ciulli, A. Biophysical Screening for the Discovery of Small-Molecule Ligands. In Protein-Ligand Interactions: Methods and Applications; Mark, A.W., Daviter, T., Eds.; Humana Press: Totowa, NJ, USA, 2013; pp. 357–388. [Google Scholar]
  14. Goodsell, D.S.; Zardecki, C.; Di Costanzo, L.; Duarte, J.M.; Hudson, B.P.; Persikova, I.; Segura, J.; Shao, C.; Voigt, M.; Westbrook, J.D.; et al. RCSB Protein Data Bank: Enabling Biomedical Research and Drug Discovery. Protein Sci. A Publ. Protein Soc. 2020, 29, 52–65. [Google Scholar] [CrossRef] [PubMed]
  15. Saur, M.; Hartshorn, M.J.; Dong, J.; Reeks, J.; Bunkoczi, G.; Jhoti, H.; Williams, P.A. Fragment-Based Drug Discovery Using Cryo-EM. Drug Discov. Today 2020, 25, 485–490. [Google Scholar] [CrossRef] [PubMed]
  16. Saha, P.; Mandal, T.; Talukdar, A.D.; Kumar, D.; Kumar, S.; Tripathi, P.P.; Wang, Q.-E.; Srivastava, A.K. DNA Polymerase Eta: A Potential Pharmacological Target for Cancer Therapy. J. Cell. Physiol. 2020, 236, 4106–4120. [Google Scholar] [CrossRef]
  17. Yang, W. An Overview of Y-Family DNA Polymerases and a Case Study of Human DNA Polymerase η. Biochemistry 2014, 53, 2793–2803. [Google Scholar] [CrossRef]
  18. Eckert, K.A. Nontraditional Roles of DNA Polymerase Eta Support Genome Duplication and Stability. Genes 2023, 14, 175. [Google Scholar] [CrossRef]
  19. Zhang, J.; Sun, W.; Ren, C.; Kong, X.; Yan, W.; Chen, X. A PolH Transcript with a Short 3′UTR Enhances PolH Expression and Mediates Cisplatin Resistance. Cancer Res. 2019, 79, 3714–3724. [Google Scholar] [CrossRef]
  20. Zafar, M.K.; Maddukuri, L.; Ketkar, A.; Penthala, N.R.; Reed, M.R.; Eddy, S.; Crooks, P.A.; Eoff, R.L. A Small-Molecule Inhibitor of Human DNA Polymerase η Potentiates the Effects of Cisplatin in Tumor Cells. Biochemistry 2018, 57, 1262–1273. [Google Scholar] [CrossRef]
  21. Srivastava, A.K.; Han, C.; Zhao, R.; Cui, T.; Dai, Y.; Mao, C.; Zhao, W.; Zhang, X.; Yu, J.; Wang, Q.-E. Enhanced Expression of DNA Polymerase Eta Contributes to Cisplatin Resistance of Ovarian Cancer Stem Cells. Proc. Natl. Acad. Sci. USA 2015, 112, 4411–4416. [Google Scholar] [CrossRef]
  22. Alt, A.; Lammens, K.; Chiocchini, C.; Lammens, A.; Pieck, J.C.; Kuch, D.; Hopfner, K.-P.; Carell, T. Bypass of DNA Lesions Generated during Anticancer Treatment with Cisplatin by DNA Polymerase Eta. Science 2007, 318, 967–970. [Google Scholar] [CrossRef] [PubMed]
  23. Albertella, M.R.; Green, C.M.; Lehmann, A.R.; O’Connor, M.J. A Role for Polymerase Eta in the Cellular Tolerance to Cisplatin-Induced Damage. Cancer Res. 2005, 65, 9799–9806. [Google Scholar] [CrossRef] [PubMed]
  24. Bassett, E.; King, N.M.; Bryant, M.F.; Hector, S.; Pendyala, L.; Chaney, S.G.; Cordeiro-Stone, M. The Role of DNA Polymerase Eta in Translesion Synthesis Past Platinum-DNA Adducts in Human Fibroblasts. Cancer Res. 2004, 64, 6469–6475. [Google Scholar] [CrossRef] [PubMed]
  25. Vaisman, A.; Masutani, C.; Hanaoka, F.; Chaney, S.G. Efficient Translesion Replication Past Oxaliplatin and Cisplatin GpG Adducts by Human DNA Polymerase Eta. Biochemistry 2000, 39, 4575–4580. [Google Scholar] [CrossRef] [PubMed]
  26. Inomata, Y.; Abe, T.; Tsuda, M.; Takeda, S.; Hirota, K. Division of Labor of Y-Family Polymerases in Translesion-DNA Synthesis for Distinct Types of DNA Damage. PLoS ONE 2021, 16, e0252587. [Google Scholar] [CrossRef]
  27. Chen, C.-Y.; Kawasumi, M.; Lan, T.-Y.; Poon, C.-L.; Lin, Y.-S.; Wu, P.-J.; Chen, Y.-C.; Chen, B.-H.; Wu, C.-H.; Lo, J.-F.; et al. Adaptation to Endoplasmic Reticulum Stress Enhances Resistance of Oral Cancer Cells to Cisplatin by Up-Regulating Polymerase η and Increasing DNA Repair Efficiency. Int. J. Mol. Sci. 2020, 22, 355. [Google Scholar] [CrossRef]
  28. Yoon, J.-H.; Choudhury, J.R.; Prakash, L.; Prakash, S. Translesion Synthesis DNA Polymerases η, ι, and ν Promote Mutagenic Replication through the Anticancer Nucleoside Cytarabine. J. Biol. Chem. 2019, 294, 19048–19054. [Google Scholar] [CrossRef]
  29. Wilkinson, N.A.; Mnuskin, K.S.; Ashton, N.W.; Woodgate, R. Ubiquitin and Ubiquitin-Like Proteins Are Essential Regulators of DNA Damage Bypass. Cancers 2020, 12, 2848. [Google Scholar] [CrossRef]
  30. Yang, W.; Gao, Y. Translesion and Repair DNA Polymerases: Diverse Structure and Mechanism. Annu. Rev. Biochem. 2018, 87, 239–261. [Google Scholar] [CrossRef]
  31. Yao, R.; Shi, L.; Wu, C.; Qiao, W.; Liu, L.; Wu, J. Lsm12 Mediates Deubiquitination of DNA Polymerase η To Help Saccharomyces cerevisiae Resist Oxidative Stress. Appl. Environ. Microbiol. 2019, 85, e01988-18. [Google Scholar] [CrossRef]
  32. Weng, P.J.; Gao, Y.; Gregory, M.T.; Wang, P.; Wang, Y.; Yang, W. Bypassing a 8,5′-Cyclo-2’-Deoxyadenosine Lesion by Human DNA Polymerase η at Atomic Resolution. Proc. Natl. Acad. Sci. USA 2018, 115, 10660–10665. [Google Scholar] [CrossRef] [PubMed]
  33. Kuraoka, I.; Robins, P.; Masutani, C.; Hanaoka, F.; Gasparutto, D.; Cadet, J.; Wood, R.D.; Lindahl, T. Oxygen Free Radical Damage to DNA. Translesion Synthesis by Human DNA Polymerase Eta and Resistance to Exonuclease Action at Cyclopurine Deoxynucleoside Residues. J. Biol. Chem. 2001, 276, 49283–49288. [Google Scholar] [CrossRef] [PubMed]
  34. Latancia, M.T.; Moreno, N.C.; Leandro, G.S.; Ribeiro, V.C.; de Souza, I.; Vieira, W.K.M.; Bastos, A.U.; Hoch, N.C.; Rocha, C.R.R.; Menck, C.F.M. DNA Polymerase Eta Protects Human Cells against DNA Damage Induced by the Tumor Chemotherapeutic Temozolomide. Mutat. Res. Genet. Toxicol. Environ. Mutagen. 2022, 878, 503498. [Google Scholar] [CrossRef]
  35. Chakraborty, A.; Tapryal, N.; Islam, A.; Sarker, A.H.; Manohar, K.; Mitra, J.; Hegde, M.L.; Hazra, T. Human DNA Polymerase η Promotes RNA-Templated Error-Free Repair of DNA Double-Strand Breaks. J. Biol. Chem. 2023, 299, 102991. [Google Scholar] [CrossRef] [PubMed]
  36. Teng, K.-Y.; Qiu, M.-Z.; Li, Z.-H.; Luo, H.-Y.; Zeng, Z.-L.; Luo, R.-Z.; Zhang, H.-Z.; Wang, Z.-Q.; Li, Y.-H.; Xu, R.-H. DNA Polymerase η Protein Expression Predicts Treatment Response and Survival of Metastatic Gastric Adenocarcinoma Patients Treated with Oxaliplatin-Based Chemotherapy. J. Transl. Med. 2010, 8, 126. [Google Scholar] [CrossRef]
  37. Zhou, W.; Chen, Y.-W.; Liu, X.; Chu, P.; Loria, S.; Wang, Y.; Yen, Y.; Chou, K.-M. Expression of DNA Translesion Synthesis Polymerase η in Head and Neck Squamous Cell Cancer Predicts Resistance to Gemcitabine and Cisplatin-Based Chemotherapy. PLoS ONE 2013, 8, e83978. [Google Scholar] [CrossRef]
  38. Barnes, R.P.; Tsao, W.-C.; Moldovan, G.-L.; Eckert, K.A. DNA Polymerase Eta Prevents Tumor Cell-Cycle Arrest and Cell Death during Recovery from Replication Stress. Cancer Res. 2018, 78, 6549–6560. [Google Scholar] [CrossRef]
  39. Li, X.-Q.; Ren, J.; Chen, P.; Chen, Y.-J.; Wu, M.; Wu, Y.; Chen, K.; Li, J. Co-Inhibition of Pol η and ATR Sensitizes Cisplatin-Resistant Non-Small Cell Lung Cancer Cells to Cisplatin by Impeding DNA Damage Repair. Acta Pharmacol. Sin. 2018, 39, 1359–1372. [Google Scholar] [CrossRef]
  40. Gorecki, L.; Andrs, M.; Korabecny, J. Clinical Candidates Targeting the ATR-CHK1-WEE1 Axis in Cancer. Cancers 2021, 13, 795. [Google Scholar] [CrossRef]
  41. Barnieh, F.M.; Loadman, P.M.; Falconer, R.A. Progress towards a Clinically-Successful ATR Inhibitor for Cancer Therapy. Curr. Res. Pharmacol. Drug Discov. 2021, 2, 100017. [Google Scholar] [CrossRef]
  42. Ngoi, N.Y.L.; Peng, G.; Yap, T.A. A Tale of Two Checkpoints: ATR Inhibition and PD-(L)1 Blockade. Annu. Rev. Med. 2022, 73, 231–250. [Google Scholar] [CrossRef] [PubMed]
  43. Menoyo, A.; Alazzouzi, H.; Espín, E.; Armengol, M.; Yamamoto, H.; Schwartz, S., Jr. Somatic Mutations in the DNA Damage-Response Genes ATR and CHK1 in Sporadic Stomach Tumors with Microsatellite Instability. Cancer Res. 2001, 61, 7727–7730. [Google Scholar] [PubMed]
  44. Tarantino, D.; Walker, C.; Weekes, D.; Pemberton, H.; Davidson, K.; Torga, G.; Frankum, J.; Mendes-Pereira, A.M.; Prince, C.; Ferro, R.; et al. Functional Screening Reveals HORMAD1-Driven Gene Dependencies Associated with Translesion Synthesis and Replication Stress Tolerance. Oncogene 2022, 41, 3969–3977. [Google Scholar] [CrossRef]
  45. Coggins, G.E.; Maddukuri, L.; Penthala, N.R.; Hartman, J.H.; Eddy, S.; Ketkar, A.; Crooks, P.A.; Eoff, R.L. N-Aroyl Indole Thiobarbituric Acids as Inhibitors of DNA Repair and Replication Stress Response Polymerases. ACS Chem. Biol. 2013, 8, 1722–1729. [Google Scholar] [CrossRef] [PubMed]
  46. Munafò, F.; Nigro, M.; Brindani, N.; Manigrasso, J.; Geronimo, I.; Ottonello, G.; Armirotti, A.; De Vivo, M. Computer-Aided Identification, Synthesis, and Biological Evaluation of DNA Polymerase η Inhibitors for the Treatment of Cancer. Eur. J. Med. Chem. 2023, 248, 115044. [Google Scholar] [CrossRef]
  47. Kalhor, S.; Fattahi, A. Design of Amino Acid- and Carbohydrate-Based Anticancer Drugs to Inhibit Polymerase η. Sci. Rep. 2022, 12, 18461. [Google Scholar] [CrossRef] [PubMed]
  48. McNeill, D.R.; Whitaker, A.M.; Stark, W.J.; Illuzzi, J.L.; McKinnon, P.J.; Freudenthal, B.D.; Wilson, D.M., 3rd. Functions of the Major Abasic Endonuclease (APE1) in Cell Viability and Genotoxin Resistance. Mutagenesis 2020, 35, 27–38. [Google Scholar] [CrossRef]
  49. Caston, R.A.; Gampala, S.; Armstrong, L.; Messmann, R.A.; Fishel, M.L.; Kelley, M.R. The Multifunctional APE1 DNA Repair–redox Signaling Protein as a Drug Target in Human Disease. Drug Discov. Today 2021, 26, 218–228. [Google Scholar] [CrossRef]
  50. Li, M.; Wilson, D.M., 3rd. Human Apurinic/apyrimidinic Endonuclease 1. Antioxid. Redox Signal. 2014, 20, 678–707. [Google Scholar] [CrossRef]
  51. Jacobs, A.L.; Schär, P. DNA Glycosylases: In DNA Repair and beyond. Chromosoma 2012, 121, 1–20. [Google Scholar] [CrossRef]
  52. Grundy, G.J.; Parsons, J.L. Base Excision Repair and Its Implications to Cancer Therapy. Essays Biochem. 2020, 64, 831–843. [Google Scholar] [PubMed]
  53. Mijit, M.; Caston, R.; Gampala, S.; Fishel, M.L.; Fehrenbacher, J.; Kelley, M.R. APE1/Ref-1—One Target with Multiple Indications: Emerging Aspects and New Directions. J. Cell. Signal. 2021, 2, 151–161. [Google Scholar] [PubMed]
  54. Lindahl, T.; Nyberg, B. Rate of Depurination of Native Deoxyribonucleic Acid. Biochemistry 1972, 11, 3610–3618. [Google Scholar] [CrossRef]
  55. Boiteux, S.; Guillet, M. Abasic Sites in DNA: Repair and Biological Consequences in Saccharomyces cerevisiae. DNA Repair 2004, 3, 1–12. [Google Scholar] [CrossRef] [PubMed]
  56. Wilson, D.M., 3rd; Barsky, D. The Major Human Abasic Endonuclease: Formation, Consequences and Repair of Abasic Lesions in DNA. Mutat. Res. 2001, 485, 283–307. [Google Scholar] [CrossRef] [PubMed]
  57. Fung, H.; Demple, B. A Vital Role for Ape1/Ref1 Protein in Repairing Spontaneous DNA Damage in Human Cells. Mol. Cell 2005, 17, 463–470. [Google Scholar] [CrossRef]
  58. Izumi, T.; Brown, D.B.; Naidu, C.V.; Bhakat, K.K.; Macinnes, M.A.; Saito, H.; Chen, D.J.; Mitra, S. Two Essential but Distinct Functions of the Mammalian Abasic Endonuclease. Proc. Natl. Acad. Sci. USA 2005, 102, 5739–5743. [Google Scholar] [CrossRef]
  59. Kim, D.V.; Kulishova, L.M.; Torgasheva, N.A.; Melentyev, V.S.; Dianov, G.L.; Medvedev, S.P.; Zakian, S.M.; Zharkov, D.O. Mild Phenotype of Knockouts of the Major Apurinic/apyrimidinic Endonuclease APEX1 in a Non-Cancer Human Cell Line. PLoS ONE 2021, 16, e0257473. [Google Scholar] [CrossRef]
  60. Walker, L.J.; Craig, R.B.; Harris, A.L.; Hickson, I.D. A Role for the Human DNA Repair Enzyme HAP1 in Cellular Protection against DNA Damaging Agents and Hypoxic Stress. Nucleic Acids Res. 1994, 22, 4884–4889. [Google Scholar] [CrossRef]
  61. Silber, J.R.; Bobola, M.S.; Blank, A.; Schoeler, K.D.; Haroldson, P.D.; Huynh, M.B.; Kolstoe, D.D. The Apurinic/apyrimidinic Endonuclease Activity of Ape1/Ref-1 Contributes to Human Glioma Cell Resistance to Alkylating Agents and Is Elevated by Oxidative Stress. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2002, 8, 3008–3018. [Google Scholar]
  62. Wang, D.; Luo, M.; Kelley, M.R. Human Apurinic Endonuclease 1 (APE1) Expression and Prognostic Significance in Osteosarcoma: Enhanced Sensitivity of Osteosarcoma to DNA Damaging Agents Using Silencing RNA APE1 Expression Inhibition. Mol. Cancer Ther. 2004, 3, 679–686. [Google Scholar] [CrossRef] [PubMed]
  63. Liu, L.; Yan, L.; Donze, J.R.; Gerson, S.L. Blockage of Abasic Site Repair Enhances Antitumor Efficacy of 1,3-Bis-(2-Chloroethyl)-1-Nitrosourea in Colon Tumor Xenografts. Mol. Cancer Ther. 2003, 2, 1061–1066. [Google Scholar] [PubMed]
  64. Fishel, M.L.; He, Y.; Reed, A.M.; Chin-Sinex, H.; Hutchins, G.D.; Mendonca, M.S.; Kelley, M.R. Knockdown of the DNA Repair and Redox Signaling Protein Ape1/Ref-1 Blocks Ovarian Cancer Cell and Tumor Growth. DNA Repair 2008, 7, 177–186. [Google Scholar] [CrossRef] [PubMed]
  65. Bobola, M.S.; Blank, A.; Berger, M.S.; Stevens, B.A.; Silber, J.R. Apurinic/apyrimidinic Endonuclease Activity Is Elevated in Human Adult Gliomas. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2001, 7, 3510–3518. [Google Scholar]
  66. Bobola, M.S.; Finn, L.S.; Ellenbogen, R.G.; Geyer, J.R.; Berger, M.S.; Braga, J.M.; Meade, E.H.; Gross, M.E.; Silber, J.R. Apurinic/apyrimidinic Endonuclease Activity Is Associated with Response to Radiation and Chemotherapy in Medulloblastoma and Primitive Neuroectodermal Tumors. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2005, 11, 7405–7414. [Google Scholar] [CrossRef]
  67. Bobola, M.S.; Silber, J.R.; Ellenbogen, R.G.; Geyer, J.R.; Blank, A.; Goff, R.D. O6-Methylguanine-DNA Methyltransferase, O6-Benzylguanine, and Resistance to Clinical Alkylators in Pediatric Primary Brain Tumor Cell Lines. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2005, 11, 2747–2755. [Google Scholar] [CrossRef]
  68. Yoshimoto, K.; Mizoguchi, M.; Hata, N.; Murata, H.; Hatae, R.; Amano, T.; Nakamizo, A.; Sasaki, T. Complex DNA Repair Pathways as Possible Therapeutic Targets to Overcome Temozolomide Resistance in Glioblastoma. Front. Oncol. 2012, 2, 186. [Google Scholar] [CrossRef]
  69. Montaldi, A.P.; Godoy, P.R.D.V.; Sakamoto-Hojo, E.T. APE1/REF-1 down-Regulation Enhances the Cytotoxic Effects of Temozolomide in a Resistant Glioblastoma Cell Line. Mutat. Res. Genet. Toxicol. Environ. Mutagen. 2015, 793, 19–29. [Google Scholar] [CrossRef]
  70. Wang, H.; Cai, S.; Bailey, B.J.; Saadatzadeh, M.R.; Ding, J.; Tonsing-Carter, E.; Georgiadis, T.M.; Gunter, T.Z.; Long, E.C.; Minto, R.E.; et al. Combination Therapy in a Xenograft Model of Glioblastoma: Enhancement of the Antitumor Activity of Temozolomide by an MDM2 Antagonist. J. Neurosurg. 2017, 126, 446–459. [Google Scholar] [CrossRef]
  71. Wheeler, D.A.; Takebe, N.; Hinoue, T.; Hoadley, K.A.; Cardenas, M.F.; Hamilton, A.M.; Laird, P.W.; Wang, L.; Johnson, A.; Dewal, N.; et al. Molecular Features of Cancers Exhibiting Exceptional Responses to Treatment. Cancer Cell 2021, 39, 38–53.e7. [Google Scholar] [CrossRef]
  72. Gojo, I.; Beumer, J.H.; Pratz, K.W.; McDevitt, M.A.; Baer, M.R.; Blackford, A.L.; Smith, B.D.; Gore, S.D.; Carraway, H.E.; Showel, M.M.; et al. A Phase 1 Study of the PARP Inhibitor Veliparib in Combination with Temozolomide in Acute Myeloid Leukemia. Clin. Cancer Res. 2017, 23, 697–706. [Google Scholar] [CrossRef] [PubMed]
  73. Sultana, R.; McNeill, D.R.; Abbotts, R.; Mohammed, M.Z.; Zdzienicka, M.Z.; Qutob, H.; Seedhouse, C.; Laughton, C.A.; Fischer, P.M.; Patel, P.M.; et al. Synthetic Lethal Targeting of DNA Double-Strand Break Repair Deficient Cells by Human Apurinic/apyrimidinic Endonuclease Inhibitors. Int. J. Cancer J. Int. Du Cancer 2012, 131, 2433–2444. [Google Scholar]
  74. Abbotts, R.; Jewell, R.; Nsengimana, J.; Maloney, D.J.; Simeonov, A.; Seedhouse, C.; Elliott, F.; Laye, J.; Walker, C.; Jadhav, A.; et al. Targeting Human Apurinic/apyrimidinic Endonuclease 1 (APE1) in Phosphatase and Tensin Homolog (PTEN) Deficient Melanoma Cells for Personalized Therapy. Oncotarget 2014, 5, 3273–3286. [Google Scholar] [PubMed]
  75. Wang, C.; Tang, M.; Chen, Z.; Nie, L.; Li, S.; Xiong, Y.; Szymonowicz, K.A.; Park, J.-M.; Zhang, H.; Feng, X.; et al. Genetic Vulnerabilities upon Inhibition of DNA Damage Response. Nucleic Acids Res. 2021, 49, 8214–8231. [Google Scholar]
  76. McMahon, A.; Zhao, J.; Yan, S. APE2: Catalytic Function and Synthetic Lethality Draw Attention as a Cancer Therapy Target. NAR Cancer 2023, 5, zcad006. [Google Scholar] [PubMed]
  77. Fleury, H.; MacEachern, M.K.; Stiefel, C.M.; Anand, R.; Sempeck, C.; Nebenfuehr, B.; Maurer-Alcalá, K.; Dillon, K.J.; Hickson, I.; Knights, C.; et al. The APE2 Nuclease Is Essential for DNA Double-Strand Break Repair by Microhomology-Mediated End Joining. Mol. Cell 2023, 83, 1429–1445.e8. [Google Scholar]
  78. Mengwasser, K.E.; Adeyemi, R.O.; Leng, Y.; Choi, M.Y.; Clairmont, C.; D’Andrea, A.D.; Elledge, S.J. Genetic Screens Reveal FEN1 and APEX2 as BRCA2 Synthetic Lethal Targets. Mol. Cell 2019, 73, 885–899.e6. [Google Scholar]
  79. Madhusudan, S.; Smart, F.; Shrimpton, P.; Parsons, J.L.; Gardiner, L.; Houlbrook, S.; Talbot, D.C.; Hammonds, T.; Freemont, P.A.; Sternberg, M.J.E.; et al. Isolation of a Small Molecule Inhibitor of DNA Base Excision Repair. Nucleic Acids Res. 2005, 33, 4711–4724. [Google Scholar]
  80. Aiello, F.; Shabaik, Y.; Esqueda, A.; Sanchez, T.W.; Grande, F.; Garofalo, A.; Neamati, N. Design and Synthesis of 3-Carbamoylbenzoic Acid Derivatives as Inhibitors of Human Apurinic/Apyrimidinic Endonuclease 1 (APE1). ChemMedChem 2012, 7, 1825–1839. [Google Scholar] [CrossRef]
  81. Zawahir, Z.; Dayam, R.; Deng, J.; Pereira, C.; Neamati, N. Pharmacophore Guided Discovery of Small-Molecule Human Apurinic/apyrimidinic Endonuclease 1 Inhibitors. J. Med. Chem. 2009, 52, 20–32. [Google Scholar]
  82. Seiple, L.A.; Cardellina, J.H., 2nd; Akee, R.; Stivers, J.T. Potent Inhibition of Human Apurinic/apyrimidinic Endonuclease 1 by Arylstibonic Acids. Mol. Pharmacol. 2008, 73, 669–677. [Google Scholar] [PubMed]
  83. Dorjsuren, D.; Kim, D.; Vyjayanti, V.N.; Maloney, D.J.; Jadhav, A.; Wilson, D.M., 3rd; Simeonov, A. Diverse Small Molecule Inhibitors of Human Apurinic/apyrimidinic Endonuclease APE1 Identified from a Screen of a Large Public Collection. PLoS ONE 2012, 7, e47974. [Google Scholar]
  84. Rai, G.; Vyjayanti, V.N.; Dorjsuren, D.; Simeonov, A.; Jadhav, A.; Wilson, D.M.; Maloney, D.J. Small Molecule Inhibitors of the Human Apurinic/Apyrimidinic Endonuclease 1 (APE1). National Center for Biotechnology Information (US); 2013. Available online: https://www.ncbi.nlm.nih.gov/books/NBK133448/ (accessed on 28 August 2023).
  85. Rai, G.; Vyjayanti, V.N.; Dorjsuren, D.; Simeonov, A.; Jadhav, A.; Wilson, D.M.; Maloney, D.J. Synthesis, Biological Evaluation, and Structure–Activity Relationships of a Novel Class of Apurinic/Apyrimidinic Endonuclease 1 Inhibitors. J. Med. Chem. 2012, 55, 3101–3112. [Google Scholar] [PubMed]
  86. Simeonov, A.; Kulkarni, A.; Dorjsuren, D.; Jadhav, A.; Shen, M.; McNeill, D.R.; Austin, C.P.; Wilson, D.M., 3rd. Identification and Characterization of Inhibitors of Human Apurinic/apyrimidinic Endonuclease APE1. PLoS ONE 2009, 4, e5740. [Google Scholar]
  87. Wilson, D.M., 3rd; Simeonov, A. Small Molecule Inhibitors of DNA Repair Nuclease Activities of APE1. Cell. Mol. Life Sci. CMLS 2010, 67, 3621–3631. [Google Scholar]
  88. Zhang, Z.; Wu, Z.; Shi, X.; Guo, D.; Cheng, Y.; Gao, J.; Liu, L.; Liu, W.; Liang, L.; Peng, L.; et al. Research Progress in Human AP Endonuclease 1: Structure, Catalytic Mechanism, and Inhibitors. Curr. Protein Pept. Sci. 2022, 23, 77–88. [Google Scholar]
  89. Laev, S.S.; Salakhutdinov, N.F.; Lavrik, O.I. Inhibitors of Nuclease and Redox Activity of Apurinic/apyrimidinic Endonuclease 1/redox Effector Factor 1 (APE1/Ref-1). Bioorganic Med. Chem. 2017, 25, 2531–2544. [Google Scholar]
  90. Pidugu, L.S.; Servius, H.W.; Sevdalis, S.E.; Cook, M.E.; Varney, K.M.; Pozharski, E.; Drohat, A.C. Characterizing Inhibitors of Human AP Endonuclease 1. PLoS ONE 2023, 18, e0280526. [Google Scholar]
  91. Mol, C.D.; Izumi, T.; Mitra, S.; Tainer, J.A. DNA-Bound Structures and Mutants Reveal Abasic DNA Binding by APE1 and DNA Repair Coordination. Nature 2000, 403, 451–456. [Google Scholar]
  92. Tsutakawa, S.E.; Shin, D.S.; Mol, C.D.; Izumi, T.; Arvai, A.S.; Mantha, A.K.; Szczesny, B.; Ivanov, I.N.; Hosfield, D.J.; Maiti, B.; et al. Conserved Structural Chemistry for Incision Activity in Structurally Non-Homologous Apurinic/Apyrimidinic Endonuclease APE1 and Endonuclease IV DNA Repair Enzymes. J. Biol. Chem. 2013, 288, 8445–8455. [Google Scholar] [CrossRef]
  93. Wilson, D.M.; Duncton, M.A.J.; Chang, C.; Luo, C.L.; Georgiadis, T.M.; Pellicena, P.; Deacon, A.M.; Gao, Y.; Das, D. Early Drug Discovery and Development of Novel Cancer Therapeutics Targeting DNA Polymerase Eta (POLH). Front. Oncol. 2021, 11, 778925. [Google Scholar] [CrossRef] [PubMed]
  94. He, H.; Chen, Q.; Georgiadis, M.M. High-Resolution Crystal Structures Reveal Plasticity in the Metal Binding Site of Apurinic/apyrimidinic Endonuclease I. Biochemistry 2014, 53, 6520–6529. [Google Scholar] [CrossRef] [PubMed]
  95. McPhillips, T.M.; McPhillips, S.E.; Chiu, H.-J.; Cohen, A.E.; Deacon, A.M.; Ellis, P.J.; Garman, E.; Gonzalez, A.; Sauter, N.K.; Phizackerley, R.P.; et al. Blu-Ice and the Distributed Control System: Software for Data Acquisition and Instrument Control at Macromolecular Crystallography Beamlines. J. Synchrotron Radiat. 2002, 9 Pt 6, 401–406. [Google Scholar] [CrossRef]
  96. Kabsch, W. XDS. Acta Crystallogr. Sect. D Biol. Crystallogr. 2010, 66, 125–132. [Google Scholar] [CrossRef]
  97. Winn, M.D.; Ballard, C.C.; Cowtan, K.D.; Dodson, E.J.; Emsley, P.; Evans, P.R.; Keegan, R.M.; Krissinel, E.B.; Leslie, A.G.W.; McCoy, A.; et al. Overview of the CCP4 Suite and Current Developments. Acta Crystallogr. Sect. D Biol. Crystallogr. 2011, 67 Pt 4, 235–242. [Google Scholar] [CrossRef]
  98. Samara, N.L.; Gao, Y.; Wu, J.; Yang, W. Detection of Reaction Intermediates in Mg 2 -Dependent DNA Synthesis and RNA Degradation by Time-Resolved X-ray Crystallography. Methods Enzymol. 2017, 592, 283–327. [Google Scholar] [CrossRef]
  99. Park, M.-J.; Kim, M.-S.; Park, I.-C.; Kang, H.-S.; Yoo, H.; Park, S.H.; Rhee, C.H.; Hong, S.-I.; Lee, S.-H. PTEN Suppresses Hyaluronic Acid-Induced Matrix Metalloproteinase-9 Expression in U87MG Glioblastoma Cells through Focal Adhesion Kinase Dephosphorylation. Cancer Res. 2002, 62, 6318–6322. [Google Scholar]
  100. Ishii, N.; Maier, D.; Merlo, A.; Tada, M.; Sawamura, Y.; Diserens, A.C.; Van Meir, E.G. Frequent Co-Alterations of TP53, p16/CDKN2A, p14ARF, PTEN Tumor Suppressor Genes in Human Glioma Cell Lines. Brain Pathol. 1999, 9, 469–479. [Google Scholar] [CrossRef] [PubMed]
  101. Tang, J.B.; Svilar, D.; Trivedi, R.N.; Wang, X.-H.; Goellner, E.M.; Moore, B.; Hamilton, R.L.; Banze, L.A.; Brown, A.R.; Sobol, R.W. N-Methylpurine DNA Glycosylase and DNA Polymerase Beta Modulate BER Inhibitor Potentiation of Glioma Cells to Temozolomide. Neuro-Oncology 2011, 13, 471–486. [Google Scholar] [CrossRef]
  102. Sykora, P.; Witt, K.L.; Revanna, P.; Smith-Roe, S.L.; Dismukes, J.; Lloyd, D.G.; Engelward, B.P.; Sobol, R.W. Next Generation High Throughput DNA Damage Detection Platform for Genotoxic Compound Screening. Sci. Rep. 2018, 8, 2771. [Google Scholar] [CrossRef]
  103. Bollag, G.; Tsai, J.; Zhang, J.; Zhang, C.; Ibrahim, P.; Nolop, K.; Hirth, P. Vemurafenib: The First Drug Approved for BRAF-Mutant Cancer. Nat. Rev. Drug Discov. 2012, 11, 873–886. [Google Scholar] [CrossRef] [PubMed]
  104. Souers, A.J.; Leverson, J.D.; Boghaert, E.R.; Ackler, S.L.; Catron, N.D.; Chen, J.; Dayton, B.D.; Ding, H.; Enschede, S.H.; Fairbrother, W.J.; et al. ABT-199, a Potent and Selective BCL-2 Inhibitor, Achieves Antitumor Activity While Sparing Platelets. Nat. Med. 2013, 19, 202–208. [Google Scholar] [CrossRef]
  105. Perera, T.P.S.; Jovcheva, E.; Mevellec, L.; Vialard, J.; De Lange, D.; Verhulst, T.; Paulussen, C.; Van De Ven, K.; King, P.; Freyne, E.; et al. Discovery and Pharmacological Characterization of JNJ-42756493 (Erdafitinib), a Functionally Selective Small-Molecule FGFR Family InhibitorPreclinical Characterization of Erdafitinib. Mol. Cancer Ther. 2017, 16, 1010–1020. [Google Scholar] [CrossRef]
  106. Tap, W.D.; Wainberg, Z.A.; Anthony, S.P.; Ibrahim, P.N.; Zhang, C.; Healey, J.H.; Chmielowski, B.; Staddon, A.P.; Cohn, A.L.; Shapiro, G.I.; et al. Structure-Guided Blockade of CSF1R Kinase in Tenosynovial Giant-Cell Tumor. N. Engl. J. Med. 2015, 373, 428–437. [Google Scholar] [CrossRef]
  107. Grädler, U.; Schwarz, D.; Blaesse, M.; Leuthner, B.; Johnson, T.L.; Bernard, F.; Jiang, X.; Marx, A.; Gilardone, M.; Lemoine, H.; et al. Discovery of Novel Cyclophilin D Inhibitors Starting from Three Dimensional Fragments with Millimolar Potencies. Bioorganic Med. Chem. Lett. 2019, 29, 126717. [Google Scholar] [CrossRef]
  108. Sabbah, M.; Mendes, V.; Vistal, R.G.; Dias, D.M.G.; Záhorszká, M.; Mikušová, K.; Korduláková, J.; Coyne, A.G.; Blundell, T.L.; Abell, C. Fragment-Based Design of Mycobacterium Tuberculosis InhA Inhibitors. J. Med. Chem. 2020, 63, 4749–4761. [Google Scholar] [CrossRef]
  109. Simon, S.C.; Wang, F.; Thomas, L.R.; Phan, J.; Zhao, B.; Olejniczak, E.T.; Macdonald, J.D.; Shaw, J.G.; Schlund, C.; Payne, W.; et al. Discovery of WD Repeat-Containing Protein 5 (WDR5)-MYC Inhibitors Using Fragment-Based Methods and Structure-Based Design. J. Med. Chem. 2020, 63, 4315–4333. [Google Scholar] [CrossRef]
Figure 1. (left) POLH Hit 1 site showing the fragment-binding pocket with fragment (blue/gray) and protein (white). The space-filling spheres (red and white spheres) show potential extents of the pocket. (right) POLH Hit 1 site with DNA (stick model) visible. Space-filling spheres (red and white dots) show empty pockets.
Figure 1. (left) POLH Hit 1 site showing the fragment-binding pocket with fragment (blue/gray) and protein (white). The space-filling spheres (red and white spheres) show potential extents of the pocket. (right) POLH Hit 1 site with DNA (stick model) visible. Space-filling spheres (red and white dots) show empty pockets.
Ijms 24 16637 g001
Figure 2. (right) POLH Hit 2 site showing the fragment-binding pocket with fragment (blue/gray) and protein (white). Space-filling spheres (red and white spheres) show potential extents of the pocket. (left) PLH Hit 2 site with DNA (white ribbon).
Figure 2. (right) POLH Hit 2 site showing the fragment-binding pocket with fragment (blue/gray) and protein (white). Space-filling spheres (red and white spheres) show potential extents of the pocket. (left) PLH Hit 2 site with DNA (white ribbon).
Ijms 24 16637 g002
Figure 3. APE1 structure with ligand at the E site (gold spheres).
Figure 3. APE1 structure with ligand at the E site (gold spheres).
Ijms 24 16637 g003
Figure 4. Electron density map of eight APE1 hits at the A site showing the quality of the hits.
Figure 4. Electron density map of eight APE1 hits at the A site showing the quality of the hits.
Ijms 24 16637 g004
Figure 5. (A) IC50 of XPTx-0091 and Inhibitor III. APE1 activity was measured as a function of increasing concentrations of XPTx-0091 (blue) or the lead APE1 inhibitor from the literature, Inhibitor III (green). (B) Specificity of XPTx-0091 for APE1 over other endonucleases. The IC50 for XPTx-001 was determined for APE1 (blue triangles) and EndoIV (gray triangles) using a fluorescence-based endonuclease assay. XPTx-0091 is > 100-fold more selective for APE1 than EndoIV. In contrast, Inhibitor III, the current lead inhibitor from the literature, is only 3-fold more selective for APE1 than EndoIV (orange and gray squares).
Figure 5. (A) IC50 of XPTx-0091 and Inhibitor III. APE1 activity was measured as a function of increasing concentrations of XPTx-0091 (blue) or the lead APE1 inhibitor from the literature, Inhibitor III (green). (B) Specificity of XPTx-0091 for APE1 over other endonucleases. The IC50 for XPTx-001 was determined for APE1 (blue triangles) and EndoIV (gray triangles) using a fluorescence-based endonuclease assay. XPTx-0091 is > 100-fold more selective for APE1 than EndoIV. In contrast, Inhibitor III, the current lead inhibitor from the literature, is only 3-fold more selective for APE1 than EndoIV (orange and gray squares).
Ijms 24 16637 g005
Figure 6. XPTx-0091 increases MMS-induced loss of proliferation. Cells were cultured at 37 °C, 5% CO2 in α-Eagle’s MEM supplemented with 10% heat-inactivated fetal bovine serum, glutamine, antibiotic/antimycotic, and gentamicin. The LN428/MPG cell line also required supplementation with G418 (600 mg/mL) (Tang et al., 2011) [101]. Cellular proliferation was determined using the direct cell counting function of the Celigo S imaging cytometer. Statistical analysis was also performed in Prism using one-way ANOVA followed by Tukey’s multiple comparisons test. Data are shown as mean +/− SEM; ** p < 0.01, *** p < 0.001, **** p < 0.0001.
Figure 6. XPTx-0091 increases MMS-induced loss of proliferation. Cells were cultured at 37 °C, 5% CO2 in α-Eagle’s MEM supplemented with 10% heat-inactivated fetal bovine serum, glutamine, antibiotic/antimycotic, and gentamicin. The LN428/MPG cell line also required supplementation with G418 (600 mg/mL) (Tang et al., 2011) [101]. Cellular proliferation was determined using the direct cell counting function of the Celigo S imaging cytometer. Statistical analysis was also performed in Prism using one-way ANOVA followed by Tukey’s multiple comparisons test. Data are shown as mean +/− SEM; ** p < 0.01, *** p < 0.001, **** p < 0.0001.
Ijms 24 16637 g006
Figure 7. XPTx-0091 increases MMS-induced DNA damage. (A) The alkaline CometChip assay was used to quantify the amount of DNA damage in XPTx-0091-treated LN428/MPG cells in the presence or absence of MMS. (B) Low-level MPG expressing LN428 cells were also tested using the CometChip assay. In brief, cells were gravity-loaded into 30 µM microwells of the CometChip apparatus for 30 min at 4 °C. After loading, the wells were washed multiple times with PBS and sealed with 0.8% low-melting-point agarose (Topvision, Thermo Fisher Scientific, Waltham, MA, USA, Cat# R0801) in PBS. The CometChip was then submerged in lysis solution with Triton X-100 detergent (BioTechne, Minneapolis, MN, USA, Cat# 4250-050-01) for 2 h at 4 °C and subsequently electrophoresed under alkaline (pH > 13) conditions (200 mM NaOH, 1 mM EDTA, 0.1% Triton X-100) at 22 V for 50 min at 4 °C. After electrophoresis, the CometChip was re-equilibrated to neutral pH using Tris buffer (0.4 M Tris·Cl, pH 7.4), and the DNA was then stained with 1x SYBR Gold (Thermo Fisher Scientific, Cat# S11494), diluted in Tris buffer* (20 mM Tris·Cl, pH 7.4) for 30 min, and de-stained for 1 h in Tris buffer*. Image acquisition was conducted on a Celigo S imaging cytometer (Nexcelom Bioscience, Lawrence, MA, USA) at a resolution of 1 micron/pixel with whole plate imaging to avoid imaging variability. Image analysis was conducted using the dedicated Comet Analysis Software (CAS) v 1.2 (https://casplab-comet-assay-software-project.soft112.com (accessed on 20 October 2023)) with the box size set to 220 × 180 pixels, representing a box size that would capture comets from heavily damaged cells without box overlap. The data acquired were exported to Excel (Microsoft, Redmond, WA, USA) and subsequently to GraphPad (Prism, V9) for statistical analysis. Data are shown as mean +/− SEM; *** p < 0.001.
Figure 7. XPTx-0091 increases MMS-induced DNA damage. (A) The alkaline CometChip assay was used to quantify the amount of DNA damage in XPTx-0091-treated LN428/MPG cells in the presence or absence of MMS. (B) Low-level MPG expressing LN428 cells were also tested using the CometChip assay. In brief, cells were gravity-loaded into 30 µM microwells of the CometChip apparatus for 30 min at 4 °C. After loading, the wells were washed multiple times with PBS and sealed with 0.8% low-melting-point agarose (Topvision, Thermo Fisher Scientific, Waltham, MA, USA, Cat# R0801) in PBS. The CometChip was then submerged in lysis solution with Triton X-100 detergent (BioTechne, Minneapolis, MN, USA, Cat# 4250-050-01) for 2 h at 4 °C and subsequently electrophoresed under alkaline (pH > 13) conditions (200 mM NaOH, 1 mM EDTA, 0.1% Triton X-100) at 22 V for 50 min at 4 °C. After electrophoresis, the CometChip was re-equilibrated to neutral pH using Tris buffer (0.4 M Tris·Cl, pH 7.4), and the DNA was then stained with 1x SYBR Gold (Thermo Fisher Scientific, Cat# S11494), diluted in Tris buffer* (20 mM Tris·Cl, pH 7.4) for 30 min, and de-stained for 1 h in Tris buffer*. Image acquisition was conducted on a Celigo S imaging cytometer (Nexcelom Bioscience, Lawrence, MA, USA) at a resolution of 1 micron/pixel with whole plate imaging to avoid imaging variability. Image analysis was conducted using the dedicated Comet Analysis Software (CAS) v 1.2 (https://casplab-comet-assay-software-project.soft112.com (accessed on 20 October 2023)) with the box size set to 220 × 180 pixels, representing a box size that would capture comets from heavily damaged cells without box overlap. The data acquired were exported to Excel (Microsoft, Redmond, WA, USA) and subsequently to GraphPad (Prism, V9) for statistical analysis. Data are shown as mean +/− SEM; *** p < 0.001.
Ijms 24 16637 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Das, D.; Duncton, M.A.J.; Georgiadis, T.M.; Pellicena, P.; Clark, J.; Sobol, R.W.; Georgiadis, M.M.; King-Underwood, J.; Jobes, D.V.; Chang, C.; et al. A New Drug Discovery Platform: Application to DNA Polymerase Eta and Apurinic/Apyrimidinic Endonuclease 1. Int. J. Mol. Sci. 2023, 24, 16637. https://doi.org/10.3390/ijms242316637

AMA Style

Das D, Duncton MAJ, Georgiadis TM, Pellicena P, Clark J, Sobol RW, Georgiadis MM, King-Underwood J, Jobes DV, Chang C, et al. A New Drug Discovery Platform: Application to DNA Polymerase Eta and Apurinic/Apyrimidinic Endonuclease 1. International Journal of Molecular Sciences. 2023; 24(23):16637. https://doi.org/10.3390/ijms242316637

Chicago/Turabian Style

Das, Debanu, Matthew A. J. Duncton, Taxiarchis M. Georgiadis, Patricia Pellicena, Jennifer Clark, Robert W. Sobol, Millie M. Georgiadis, John King-Underwood, David V. Jobes, Caleb Chang, and et al. 2023. "A New Drug Discovery Platform: Application to DNA Polymerase Eta and Apurinic/Apyrimidinic Endonuclease 1" International Journal of Molecular Sciences 24, no. 23: 16637. https://doi.org/10.3390/ijms242316637

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop