Next Article in Journal
Genetic Aspects of Problematic and Risky Internet Use in Young Men—Analysis of ANKK1, DRD2 and NTRK3 Gene Polymorphism
Next Article in Special Issue
Developmental Changes in Genome Replication Progression in Pluripotent versus Differentiated Human Cells
Previous Article in Journal
Exploring Aegilops caudata: A Comprehensive Study of the CslF6 Gene and β-Glucan
Previous Article in Special Issue
Role of RPA Phosphorylation in the ATR-Dependent G2 Cell Cycle Checkpoint
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Intriguing Mystery of RPA Phosphorylation in DNA Double-Strand Break Repair

by
Valerie J. Fousek-Schuller
1 and
Gloria E. O. Borgstahl
2,*
1
Immunology, Pathology, and Infectious Diseases, UNMC, Omaha, NE 68198-6805, USA
2
Eppley Institute for Research in Cancer & Allied Diseases, UNMC, Omaha, NE 68198-6805, USA
*
Author to whom correspondence should be addressed.
Genes 2024, 15(2), 167; https://doi.org/10.3390/genes15020167
Submission received: 20 December 2023 / Revised: 23 January 2024 / Accepted: 25 January 2024 / Published: 27 January 2024
(This article belongs to the Special Issue Mechanisms and Regulation of Human DNA Replication)

Abstract

:
Human Replication Protein A (RPA) was historically discovered as one of the six components needed to reconstitute simian virus 40 DNA replication from purified components. RPA is now known to be involved in all DNA metabolism pathways that involve single-stranded DNA (ssDNA). Heterotrimeric RPA comprises several domains connected by flexible linkers and is heavily regulated by post-translational modifications (PTMs). The structure of RPA has been challenging to obtain. Various structural methods have been applied, but a complete understanding of RPA’s flexible structure, its function, and how it is regulated by PTMs has yet to be obtained. This review will summarize recent literature concerning how RPA is phosphorylated in the cell cycle, the structural analysis of RPA, DNA and protein interactions involving RPA, and how PTMs regulate RPA activity and complex formation in double-strand break repair. There are many holes in our understanding of this research area. We will conclude with perspectives for future research on how RPA PTMs control double-strand break repair in the cell cycle.

1. Introduction

Replication protein A (RPA) is a multifaceted, heterotrimeric protein involved in various DNA metabolism activities, including cell cycle regulation, DNA damage signaling, and DNA repair mechanisms [1]. RPA is the main ssDNA-binding protein in humans. Historically, RPA was first discovered in the Kelly laboratory at Johns Hopkins University and identified as one of six purified components needed for in vitro simian virus 40 DNA replication. Then, RPA was found to play essential roles in all DNA metabolism pathways that involve single-stranded DNA (ssDNA) [2]. RPA research was pioneered and studied throughout Dr. Marc Wold’s entire career, who discovered the critical role of RPA in various pathways and the importance of RPA regulation by post-translational modifications (PTMs) [3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33]. RPA research has dramatically expanded and is of interest to many scientists. The RPA heterotrimer has three subunits named after their corresponding molecular weights: RPA70, RPA32, and RPA14 [34].
Over the years, research has shown that RPA has multiple types and patterns of PTMs, including acetylation, ubiquitination, SUMOylation, ADP ribosylation, and phosphorylation [28,35,36,37,38]. In this review, we focus on the role of phosphorylation. Many studies dating back to 1990 have focused on the phosphorylation of RPA, specifically the N-terminal domain of RPA32, because of the impact phosphorylation has on cell cycle regulation [39]. Also, the phosphorylation of RPA32 is easy to detect due to a noticeable shift to a higher molecular weight of the RPA32 subunit on SDS-PAGE. RPA32 has a long, unstructured N-terminal tail that is phosphorylated, and these sites have several phosphospecific antibodies. The phosphorylation of other subunits has also been studied through western blot and isoelectric focusing analysis of phosphoisoforms [40,41]. However, reagents for these sites could be more developed.
Since RPA phosphorylation before and after DNA damage has been shown, it is important to determine which sites on RPA are most important for the function of RPA to repair DNA damage, such as double-stranded breaks (DSBs) [40]. Previous data revealed that phosphatidyl inositol 3-kinase-like serine/threonine protein kinase (PIKK) sites and cyclin-dependent kinase (CDK) sites are critical for DNA repair [42,43,44]. Some examples of PIKK sites that will be of interest throughout this review are AT-mutated (ataxia telangiectasia, ATM) and AT and Rad3-related (ATR). Interestingly, RPA contains numerous PIKK sites, which may affect RPA function.
Over the decades, numerous reviews have been conducted about RPA’s structure and function [16,23,29,35,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62]. This review will discuss novel research and more recent findings on RPA. Specifically, we will cover recent discoveries in the structure/function of RPA in DNA metabolism, interactions, and PTMs involving RPA, specific details on RPA phosphorylation in the DNA damage response (DDR) in the homologous recombination pathway, and finally conclude with a summary of necessary areas for future RPA research.

2. Structure of RPA

Heterotrimeric RPA has seven structured domains. RPA70 contains domains A, B, C, and F, with disordered linkers connecting the ordered domains and a zinc atom in domain C (Figure 1A). RPA32 has structured domains D and wHLH (winged helix-loop-helix), which are also connected by disordered linkers. Finally, RPA14 is entirely composed of one ordered domain named E. In Figure 1, each domain is drawn as a colored box, which indicates a known structure, whereas the black and white rectangles are disordered regions connecting the ordered, structured domains. Domains A, B, C, and D are where ssDNA binding occurs, and domains F, A, and wHLH contain sequences for protein interactions. The heterotrimeric core of RPA is formed by domains C, D, and E—one domain from each subunit.
This unique RPA structure divides six oligosaccharide binding folds (OB-folds) among the heterotrimeric protein subunits. OB-folds are typically five-stranded closed β barrels that interact with ligands such as oligosaccharides, proteins, catalytic substrates, and more. These six OB-folds in RPA are also frequently found in ssDNA-binding proteins (Figure 1B) [53]. RPA70 contains four OB-folds that comprise the DNA binding domains within domains A, B, C, and F. RPA14 domain E and RPA32 domain D are also OB-folds [63,64,65]. The OB-folds are conserved in structure, which have more structural homology than sequence homology (Figure 1C) [53].
Figure 1. Structure of RPA. (A) Domain diagram of RPA heterotrimer with domain functions and phosphorylation sites. Structured domains are colored and labeled A–F and wHLH. Unstructured domains are white rectangles with a black outline. Phosphorylation sites are labeled and shown with a yellow circle indicated by P. Protein interaction domains are underlined in navy blue. DNA binding domains are underlined in maroon. Subunit interactions that form the heterotrimeric core are underlined in dark yellow. (B) OB-fold diagram [66]. Arrows indicate the β-strands and an oval indicates the α-helices. The blue β-strands correspond to those that comprise the OB-fold. The L12 loop lies between β1′ and β2, and the L45 loop lies between β4′ and β5′. (C) Sequence and secondary structure alignment of domains A–E based on structure (updated from [53]). Orange secondary structure elements represent domains A and B, green elements represent domain C, and blue elements represent domains D and E. Lowercase z indicates the Cysteine residues in domain C that bind zinc. Residues in RPA70-A, RPA70-B, RPA70-C, and RPA32-D that bind to ssDNA are underlined. RPA32-D contains putative DNA binding residues that were determined through structural alignment of human and Ustilago maydis RPA. (D) AlphaFold predicted the structure of RPA with ssDNA. Domains are colored according to the domain diagram in A. The zinc in RPA70-C is shown as a gray ball. PIKK/CDK in G2 phase phosphorylation sites are black and labeled. The phosphorylation sites: in RPA70 are S38, T180, S207, and T483; in RPA32 are S4, S8, S11, S12, S13, T21, S23, S29, S33, S52, S72, S174. ssDNA, shown in maroon, was superimposed from 4GNX [67].
Figure 1. Structure of RPA. (A) Domain diagram of RPA heterotrimer with domain functions and phosphorylation sites. Structured domains are colored and labeled A–F and wHLH. Unstructured domains are white rectangles with a black outline. Phosphorylation sites are labeled and shown with a yellow circle indicated by P. Protein interaction domains are underlined in navy blue. DNA binding domains are underlined in maroon. Subunit interactions that form the heterotrimeric core are underlined in dark yellow. (B) OB-fold diagram [66]. Arrows indicate the β-strands and an oval indicates the α-helices. The blue β-strands correspond to those that comprise the OB-fold. The L12 loop lies between β1′ and β2, and the L45 loop lies between β4′ and β5′. (C) Sequence and secondary structure alignment of domains A–E based on structure (updated from [53]). Orange secondary structure elements represent domains A and B, green elements represent domain C, and blue elements represent domains D and E. Lowercase z indicates the Cysteine residues in domain C that bind zinc. Residues in RPA70-A, RPA70-B, RPA70-C, and RPA32-D that bind to ssDNA are underlined. RPA32-D contains putative DNA binding residues that were determined through structural alignment of human and Ustilago maydis RPA. (D) AlphaFold predicted the structure of RPA with ssDNA. Domains are colored according to the domain diagram in A. The zinc in RPA70-C is shown as a gray ball. PIKK/CDK in G2 phase phosphorylation sites are black and labeled. The phosphorylation sites: in RPA70 are S38, T180, S207, and T483; in RPA32 are S4, S8, S11, S12, S13, T21, S23, S29, S33, S52, S72, S174. ssDNA, shown in maroon, was superimposed from 4GNX [67].
Genes 15 00167 g001
Human RPA has been biochemically studied for over 30 years, with structure determination by X-ray crystallography and nuclear magnetic resonance (NMR) spectroscopy [66,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82]. Unfortunately, full-length RPA has not been crystallized as a heterotrimeric protein because of the many disordered regions throughout the protein complex. By dividing RPA into purified domains, it was possible to crystallize each domain individually or with specific binding partners, such as the heterotrimeric core involving domains C, D, and E [71,75,83]. A Ustilago maydis structure was solved composed of ssDNA and domains A, B, C, D, and E [67]. The wHLH domain is also very loosely packed, so NMR was used to determine the wHLH structure with a bound peptide [69,82]. In 2014, Feldkamp and coworkers obtained a crystal structure of the wHLH domain without a peptide [82]. In 2015, Brosey and colleagues studied structural dynamics by NMR to evaluate RPA70-A/B domain binding to ssDNA. They obtained NMR structures for ssDNA bound to RPA70-A/B and compared their results to other structural analyses [84].
Since the flexible disorder is a significant factor in RPA structure, small angle X-ray scattering (SAXS) is another method used. This method allows for low-resolution structures from proteins in solution. In 2010, Pretto and others used SAXS specifically on RPA70 domains A and B and the N-terminal region to see how these domains bind to ssDNA. They concluded that RPA70 domains A and B are more compact in the presence of ssDNA. The N-terminal region of RPA70 did not change in structure in the presence of DNA, allowing them to conclude that there is no interaction between the two [84,85,86].
With these structurally known domains of RPA, AlphaFold can be handy to try to obtain a prediction of the entire RPA structure using experimentally known structures in the Protein Data Bank (Table 1). We used AlphaFold to predict various conformations of the complete protein, including the disordered regions and structured domains (Figure 1D). The modeled AlphaFold structure was used to visualize candidate PIKK/CDK Ser and Thr phosphorylation sites involved in DNA damage repair, specifically DSBs. These sites are shown in black (Figure 1D) and labeled with the corresponding amino acid and residue number. There are four phosphorylation sites on RPA70, twelve on RPA32, and none on RPA14. ssDNA was superimposed using the 4GNX structure, which interacts with RPA70 domains A, B, C, and RPA32 domain D [67]. The specific residues that bind zinc and ssDNA are indicated in Figure 1C.

3. The Function of RPA in DNA Double-Strand Break Repair

RPA is critical in coordinating the cell cycle with DNA replication and any DNA repair pathway involving ssDNA. The cell cycle is crucial in completing a series of events that allow cells to grow and divide and includes the following stages: interphase (I), G1 phase, S phase, G2 phase, and mitosis (M). Regulation is key to successfully replicating DNA and normal cell division within these cell cycle phases. Several laboratories have studied RPA in cell cycle regulation. RPA was determined not to be phosphorylated in the G1 phase, but RPA34 in humans and S. cerevisiae is phosphorylated in cells that enter the S phase and then are dephosphorylated in the M phase [45,87,88]. Various researchers have focused on RPA’s role in recruiting proteins involved in DNA repair. RPA is involved in most DNA repair pathways, including base excision repair (BER), nucleotide excision repair (NER), mismatch repair (MMR), and HR [36,89,90,91,92]. It is also important to note that HR occurs only during S and G2 phases [40,44].
RPA plays a critical role in HR to repair DSBs. In HR, sections of homologous sequences, typically on a sister chromatid or nearby repeated sequence, will fill in the gap left by the DSB. On each side of this DSB, a 5′ strand is resected, leaving a 3′ overhang on the DNA. This resection is done by the MRN complex (composed of RAD50, mitotic recombination 11 (MRE11), and Nijmegen breakage syndrome 1 (NBS1)), along with other proteins, including CtBP-interacting protein (CtIP), exonuclease 1 (EXO1), DNA replication helicase/nuclease DNA2, and blood syndrome RecQ like helicase (BLM). Once there is a 3′ ssDNA overhang, RPA will bind to this ssDNA and follow the canonical HR pathway that will recruit a complex that includes breast cancer type 1 susceptibility protein (BRCA1), breast cancer type 2 susceptibility protein (BRCA2), and partner and localizer of BRCA2 (PALB1). This BRCA2 complex will recruit radiation sensitivity gene 51 (RAD51) and undergo ssDNA transfer from RPA to BRCA2 complex to RAD51, and RAD51 will perform a homology search on the sister chromatid, eventually repairing the DSB. Secondly, an alternative pathway for HR includes radiation sensitivity gene 52 (RAD52). RAD52 has similar functionality to the BRCA2 complex by undergoing a ssDNA handoff from RPA and allowing RAD51 nucleation of the ssDNA again to perform a homology search on the sister chromatid, eventually repairing the DSB [40,53,93,94,95].

4. ssDNA and Protein Interactions involving RPA

RPA involves an assorted array of interactions with ssDNA and proteins. The function of RPA binding to ssDNA has been studied in numerous ways, for example, through atomic force microscopy (AFM), circular dichroism (CD), optical tweezers, single-molecule fluorescence resonance energy transfer (smFRET), and many more [61,96,97]. The interaction between ssDNA and RPA is a rapid but extremely stable process that is difficult to evaluate due to the binding strength. A study in 2000 concluded that the zinc-finger motif is necessary to form a stable RPA:ssDNA complex, which depends on redox reactions [98]. Others focused on the binding and unfolding of non-canonical ssDNA structures by RPA in DNA replication. The heterotrimeric core of RPA selectively binds to a G-quadruplex forming sequence. G-quadruplexes create a unique problem for RPA to unfold, and various ligands could hinder DNA replication by not allowing RPA to unfold these G-quadruplexes correctly [99].
A more recent paper described how the ssDNA binding complex can change its binding mode to allow for nucleation of RAD51. The spacing between RPA:ssDNA complexes is essential for leaving bare nucleotides in between the complexes to allow for RAD51 nucleation. The exchange of RPA for RAD51 is known to be mediated by the RAD52 middle region [100]. Overall, the RPA:ssDNA complex is essential for life and continues to be investigated.
With this insight into the RPA:ssDNA complex, multiple laboratories have investigated how RPA binds to varying lengths of nucleotides (nt), for example, 8-nt, 20-nt, or 30-nt. Kang and coworkers in 2023 investigated a protein involved in Kallmann syndrome: N-methyl-D-aspartate receptor synaptonuclear signaling and neuronal migration factor (NSMF). They determined that NSMF will colocalize and physically interact with RPA during DNA damage, allowing RPA to bind in a 30-nt binding mode. This prepares RPA to become phosphorylated by ATR, an upstream kinase involved in HR [101]. They concluded that the 30-nt binding mode of RPA enhances the phosphorylation of RPA32 by ATR, and further phosphorylation will stabilize RPA binding to ssDNA.
RPA has many protein–protein interactions involving DSB repair proteins and others (Table 2). DSB repair binding partners include: BRCA2, RAD52, RAD51, ATR/ATRIP, DNA-PKcs, DSS1, MRE11-RAD50-NBS1, p53, and PP2A [18,21,36,38,69,72,74,90,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166,167,168,169,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196,197,198,199,200,201,202,203,204,205,206,207,208,209,210,211,212]. The interaction of RPA with RAD52 in the alternative HR DSB repair pathway and single-strand annealing is fascinating. Without the RAD52:RPA interaction, the ssDNA handoff from RPA to RAD51 in alternative HR would not be possible, and it is also critical that RPA is phosphorylated to allow this DNA handoff [93]. Previous studies identified residues 224-271 on RPA32 and 169-326 on RPA70 that include binding sites for RAD52. It was also found that RAD52 residues 218-303 bind RPA70, along with RPA32 [21]. Although the interaction sites have been defined, structural data has yet to be available for the RPA:RAD52 complex.
BRCA2 is an essential protein involved in the canonical HR pathway. BRCA2 also completes the DNA handoff, like RAD52, by interacting with RPA. BRCA2 mutations in women can cause familial, early-onset breast cancer, so understanding these mutations and their potential impact on the interaction with RPA will become vital in understanding DNA repair, specifically in cancer cells. A study in 2003 used a cancer-predisposing BRCA2 mutation (Y42C) to investigate its interaction with RPA. They found that the Y42C mutation inhibited the interaction between RPA:BRCA2, showing that the Y42C mutation has biological importance within the human body [213]. The BRCA2 interaction with RAD51 has been studied extensively over the years, as BRCA2 promotes RAD51 nucleation on the ssDNA. Still, there has yet to be data on the interaction between BRCA2 and RPA to specifically understand how and where this interaction occurs [214].

5. PTMs That Regulate RPA in DNA Metabolism

PTMs are chemical modifications involved in the functional regulation of cellular proteins. RPA relies on PTMs for its proper function in the cell cycle and DNA repair, and the main PTM we will focus on in this review is phosphorylation. Phosphorylation of RPA has been studied for decades, with most studies focused on the RPA32 N-terminal region. Understanding the importance of phosphorylation on other RPA subunits in RPA’s regulation is still unresolved.
Phosphorylation occurs throughout the cell cycle, as previously described. A paper by Yates and coworkers determined two cell cycle checkpoint kinases in yeast, Mec1 and Ddc2 (ATR and ATRIP, respectively, in humans), that are essential for replication stress response and DDR. These two kinases were shown to recruit ssDNA binding to RPA through Ddc2 by phosphorylation. A yeast Ddc2-RPA structure through X-ray crystallography indicated that this interaction is necessary to mediate RPA phosphorylation during DDR [215].
An early study in 1990 by Din and colleagues determined the phosphorylation of RPA32 and RPA70 in human and yeast cells by phosphoamino acid analysis using P32 labeling [39]. Then, numerous labs focused on the phosphorylation of RPA because of its critical regulation in the DNA repair pathway. In 2003, Binz and others discovered that the phosphorylation of RPA32 modulates RPA-dsDNA interactions and subsequent destabilization [22]. They concluded that an intersubunit interaction between phosphorylated RPA32NT and RPA70 N-terminal domain (RPA70N) was possible. Recently, NMR and docking studies were conducted to investigate the interaction between the RPA70 N-terminal domain (RPA70N) and a phosphomimetic N-terminal peptide of RPA32 with candidate serine/threonine sites mutated to aspartic acid. The study showed the possibility that RPA32 N-terminal phosphorylation could allow for transient interaction with RPA70N, which could alter or enhance other interaction sites for proteins or DNA [216]. These conclusions indicated a possible role of RPA32 N-terminal phosphorylation binding with RPA70N to form a less disordered structure if these interactions occur in the context of the heterotrimer.
Not only is RPA involved in RPA-DNA interactions, but RPA phosphorylation also influences subcellular localization, especially when there is mitotic phosphorylation at S23 and S29 in RPA32 [217]. When the cell has a DDR, RPA becomes hyperphosphorylated [50,56]. For example, in response to DSBs, RPA is a substrate of ATM kinase that phosphorylates RPA32 at Thr21 and most likely others [43]. In 2014, a study was performed on RPA hyperphosphorylation involving the HR pathway in cell cycle phases where HR DSB repair occurs. In these studies, a human squamous cell carcinoma cell line (UM-SCC-38) was synchronized in the S and G2 phases [218,219,220]. Phosphorylation sites in S and G2 phases with and without DSBs were analyzed with western blots using all available antibodies for phosphorylation sites on RPA32. In the G2 phase, chromatin-bound immunoprecipitation phosphorylation occurred at S4/8, S12, T21, S23, and S33 of RPA32 after DSBs.
A global view of RPA heterotrimer phosphorylation was explained using a high-quality RPA70 antibody and isoelectric focusing (IEF) to separate RPA isoforms based on the number of phosphorylations present. Capillary IEF immunoassay data of the phosphorylated RPA heterotrimeric isoforms with and without DSB treatment, synchronized in the G2 phase of the cell cycle, were collected (Figure 2) [40]. In this study, data showed that RPA is always a phosphoprotein in the G2 phase by having up to seven phosphates (Figure 2A, blue line). Hyperphosphorylated isoforms generated after DSBs contained up to 13 or 14 phosphates. IEF identified some of these sites with phosphospecific antibodies (Figure 2B–E). The hyperphosphorylated forms with 13–14 phosphates included pS4/8, pT21, and pS33. Several of the sites had yet to be characterized before. DSBs showed increased phosphorylation by PIKK and CDK kinases in S and G2 phases, providing new information on candidate PIKK and CDK sites for future research.
This research provided pertinent information about RPA heterotrimeric phosphorylation in response to DSBs in the S and G2 phases of the human cell cycle. Several of these phosphorylation sites are PIKK-specific sites in the DNA binding and protein interaction domains of RPA. A summary figure describes possible phosphorylation sites before and after DSBs (Figure 3). This study provided data about phosphatase and kinase action in remodeling RPA isoforms in response to DNA damage [40].

6. Future Directions for RPA Research

As previously described, PTMs, such as phosphorylation, play a crucial role in RPA regulation. Future RPA research should focus on the many isoforms of phosphorylated RPA and the complicated phosphorylation patterns of RPA to determine which ones are necessary for DNA repair, cell cycle regulation, or other metabolism pathways. Phosphorylation is critical in RPA regulation and impacts some DNA repair pathways.
RPA phosphorylation could impact protein–protein interactions in DNA repair, DNA DSB repair, replication, gene expression, and more (Table 2). These interactions could be explored through various binding studies, including thermal shift assays, biolayer interferometry (BLI), and surface plasmon resonance (SPR). CD spectroscopy could show if phosphorylation changes secondary structure [221]. In vivo and in vitro activity assays would be beneficial to determine where these phosphorylation sites impact RPA function and interactions with proteins. Phosphomimetics, for example, serine/threonine to glutamic acid mutations, could be used in these problematic studies. Understanding how phosphorylation affects the RPA–protein interactions related to DNA DSB repair would give insight into how to target the interactions only involved in cancerous cells.
Understanding the role of RPA phosphorylation could lead to the creation of drugs that specifically target these forms of protein and complexes. In cancer cells deficient in the BRCA2 complex, the alternative HR pathway involving RAD52 must step up to continue HR, but most importantly, it must continue to repair the broken DNA. Deng and coworkers demonstrated that RPA phosphorylation significantly stabilized the RPA:RAD52 complex and regulates strand transfer [93]. So, the phosphoRPA:RAD52 protein complex would be a potential target for a therapeutic inhibitor of complex formation. With this interaction inhibited in HR-deficient cancer cells, such as high-grade serous carcinoma ovarian cancer, these cells will not be able to survive. The most significant outcome of these potential pharmaceuticals will be the lack of impact on the non-cancerous cells in the body with functional canonical BRCA2 pathway.
Visualization of proteins is key for adequately understanding how they function, and the disordered regions of RPA make this process difficult. It is known that RPA alone and the RPA:RAD52 complex (with and without ssDNA) cannot be crystallized due to this intrinsic disorder. Techniques such as cryo-electron microscopy (cryo-EM) will be a helpful tool to structurally analyze RPA in its entirety [83,222,223]. Utilizing cryo-EM to study RPA protein–protein interactions (Table 2) would allow for investigation and fill a massive gap in the knowledge of these interactions.

7. Conclusions

RPA is a heterotrimeric protein involved in cell cycle regulation, repair, and countless other pathways involved in DNA metabolism. Regulation of this multifarious protein is essential in DNA metabolism. The structure of RPA has been studied for several decades, with only a handful of experiments on the entire protein because of the disordered regions. AlphaFold has predicted RPA as a complete heterotrimeric protein, but the disordered regions vary in each prediction. The model has allowed us to visualize where the phosphorylated PIKK/CDK sites are throughout the protein. This lets us understand which interactions in DNA binding, protein–protein interactions, or formation of the heterotrimeric core might be altered.
Regulation of RPA is a key component involved in various processes, such as PTMs. PTMs like acetylation, ubiquitination, SUMOylation, ADP ribosylation, and phosphorylation are fundamental to RPA regulation. Throughout this review, the main PTM focused on phosphorylation because of its complexity and misunderstanding throughout the years. The number of phosphorylation combinations is mind-boggling. Phosphorylation of RPA is complex and an important part of controlling multiple interactions, maintaining cell cycle regulation, and repairing DNA. Therefore, RPA is an intriguing mystery that needs in-depth characterization of phosphorylation in DNA double-strand break repair to understand RPA functions fully.

Author Contributions

Writing—original draft preparation and figures, V.J.F.-S.; writing—review and editing, G.E.O.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the CMDRP DOD OCRP (W81XWH-20-1-0816), Fred and Pamela Buffett NCI Cancer Center Support Grant (P30CA036727), and VFS was funded by a Department of Education GAANN fellowship (P200A210032).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank Lucas Struble for helpful discussion and guidance on this RPA project. We also thank Jeff Lovelace and Jahaun Azadmanesh for helping us with the AlphaFold model. Figures were created with BioRender.com.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bastin-Shanower, S.A.; Brill, S.J. Functional analysis of the four DNA binding domains of replication protein A. The role of RPA2 in ssDNA binding. J. Biol. Chem. 2001, 276, 36446–36453. [Google Scholar] [CrossRef]
  2. Wold, M.S.; Weinberg, D.H.; Virshup, D.M.; Li, J.J.; Kelly, T.J. Identification of cellular proteins required for simian virus 40 DNA replication. J. Biol. Chem. 1989, 264, 2801–2809. [Google Scholar] [CrossRef]
  3. Wold, M.S.; McMacken, R. Regulation of expression of the Escherichia coli dnaG gene and amplification of the dnaG primase. Proc. Natl. Acad. Sci. USA 1982, 79, 4907–4911. [Google Scholar] [CrossRef]
  4. Wold, M.S.; Mallory, J.B.; Roberts, J.D.; LeBowitz, J.H.; McMacken, R. Initiation of bacteriophage lambda DNA replication in vitro with purified lambda replication proteins. Proc. Natl. Acad. Sci. USA 1982, 79, 6176–6180. [Google Scholar] [CrossRef] [PubMed]
  5. Wold, M.S.; Li, J.J.; Kelly, T.J. Initiation of simian virus 40 DNA replication in vitro: Large-tumor-antigen- and origin-dependent unwinding of the template. Proc. Natl. Acad. Sci. USA 1987, 84, 3643–3647. [Google Scholar] [CrossRef] [PubMed]
  6. Kelly, T.J.; Rosenfeld, P.J.; Wides, R.J.; O’Neill, E.A.; Li, J.J.; Wold, M.S. Replication of adenovirus and SV40 chromosomes in vitro. Philos. Trans. R. Soc. Lond. B Biol. Sci. 1987, 317, 429–438. [Google Scholar] [CrossRef]
  7. Wold, M.S.; Kelly, T. Purification and characterization of replication protein A, a cellular protein required for in vitro replication of simian virus 40 DNA. Proc. Natl. Acad. Sci. USA 1988, 85, 2523–2527. [Google Scholar] [CrossRef]
  8. Erdile, L.F.; Wold, M.S.; Kelly, T.J. The primary structure of the 32-kDa subunit of human replication protein A. J. Biol. Chem. 1990, 265, 3177–3182. [Google Scholar] [CrossRef] [PubMed]
  9. Kim, C.; Snyder, R.O.; Wold, M.S. Binding properties of replication protein A from human and yeast cells. Mol. Cell Biol. 1992, 12, 3050–3059. [Google Scholar] [CrossRef]
  10. Henricksen, L.A.; Umbricht, C.B.; Wold, M.S. Recombinant replication protein A: Expression, complex formation, and functional characterization. J. Biol. Chem. 1994, 269, 11121–11132. [Google Scholar] [CrossRef]
  11. Henricksen, L.A.; Wold, M.S. Replication protein A mutants lacking phosphorylation sites for p34cdc2 kinase support DNA replication. J. Biol. Chem. 1994, 269, 24203–24208. [Google Scholar] [CrossRef]
  12. Gomes, X.V.; Wold, M.S. Structural analysis of human replication protein A. Mapping functional domains of the 70-kDa subunit. J. Biol. Chem. 1995, 270, 4534–4543. [Google Scholar] [CrossRef]
  13. He, Z.; Henricksen, L.A.; Wold, M.S.; Ingles, C.J. RPA involvement in the damage-recognition and incision steps of nucleotide excision repair. Nature 1995, 374, 566–569. [Google Scholar] [CrossRef]
  14. Gomes, X.V.; Henricksen, L.A.; Wold, M.S. Proteolytic mapping of human replication protein A: Evidence for multiple structural domains and a conformational change upon interaction with single-stranded DNA. Biochemistry 1996, 35, 5586–5595. [Google Scholar] [CrossRef]
  15. Henricksen, L.A.; Carter, T.; Dutta, A.; Wold, M.S. Phosphorylation of human replication protein A by the DNA-dependent protein kinase is involved in the modulation of DNA replication. Nucleic Acids Res. 1996, 24, 3107–3112. [Google Scholar] [CrossRef]
  16. Wold, M.S. Replication protein A: A heterotrimeric, single-stranded DNA-binding protein required for eukaryotic DNA metabolism. Annu. Rev. Biochem. 1997, 66, 61–92. [Google Scholar] [CrossRef]
  17. Sibenaller, Z.A.; Sorensen, B.R.; Wold, M.S. The 32- and 14-kilodalton subunits of replication protein A are responsible for species-specific interactions with single-stranded DNA. Biochemistry 1998, 37, 12496–12506. [Google Scholar] [CrossRef] [PubMed]
  18. Shao, R.G.; Cao, C.X.; Zhang, H.; Kohn, K.W.; Wold, M.S.; Pommier, Y. Replication-mediated DNA damage by camptothecin induces phosphorylation of RPA by DNA-dependent protein kinase and dissociates RPA:DNA-PK complexes. EMBO J. 1999, 18, 1397–1406. [Google Scholar] [CrossRef] [PubMed]
  19. Walther, A.P.; Gomes, X.V.; Lao, Y.; Lee, C.G.; Wold, M.S. Replication protein A interactions with DNA. 1. Functions of the DNA-binding and zinc-finger domains of the 70-kDa subunit. Biochemistry 1999, 38, 3963–3973. [Google Scholar] [CrossRef] [PubMed]
  20. Buchko, G.W.; Daughdrill, G.W.; de Lorimier, R.; Rao, B.K.; Isern, N.G.; Lingbeck, J.M.; Taylor, J.S.; Wold, M.S.; Gochin, M.; Spicer, L.D.; et al. Interactions of human nucleotide excision repair protein XPA with DNA and RPA70 Delta C327: Chemical shift mapping and 15N NMR relaxation studies. Biochemistry 1999, 38, 15116–15128. [Google Scholar] [CrossRef] [PubMed]
  21. Jackson, D.; Dhar, K.; Wahl, J.K.; Wold, M.S.; Borgstahl, G.E. Analysis of the human replication protein A:Rad52 complex: Evidence for crosstalk between RPA32, RPA70, Rad52 and DNA. J. Mol. Biol. 2002, 321, 133–148. [Google Scholar] [CrossRef]
  22. Binz, S.K.; Lao, Y.; Lowry, D.F.; Wold, M.S. The phosphorylation domain of the 32-kDa subunit of replication protein A (RPA) modulates RPA-DNA interactions. Evidence for an intersubunit interaction. J. Biol. Chem. 2003, 278, 35584–35591. [Google Scholar] [CrossRef] [PubMed]
  23. Binz, S.K.; Sheehan, A.M.; Wold, M.S. Replication protein A phosphorylation and the cellular response to DNA damage. DNA Repair 2004, 3, 1015–1024. [Google Scholar] [CrossRef] [PubMed]
  24. Mason, A.C.; Haring, S.J.; Pryor, J.M.; Staloch, C.A.; Gan, T.F.; Wold, M.S. An alternative form of replication protein a prevents viral replication in vitro. J. Biol. Chem. 2009, 284, 5324–5331. [Google Scholar] [CrossRef] [PubMed]
  25. Kemp, M.G.; Mason, A.C.; Carreira, A.; Reardon, J.T.; Haring, S.J.; Borgstahl, G.E.; Kowalczykowski, S.C.; Sancar, A.; Wold, M.S. An alternative form of replication protein a expressed in normal human tissues supports DNA repair. J. Biol. Chem. 2010, 285, 4788–4797. [Google Scholar] [CrossRef] [PubMed]
  26. Hass, C.S.; Gakhar, L.; Wold, M.S. Functional characterization of a cancer causing mutation in human replication protein A. Mol. Cancer Res. 2010, 8, 1017–1026. [Google Scholar] [CrossRef] [PubMed]
  27. Hass, C.S.; Lam, K.; Wold, M.S. Repair-specific functions of replication protein A. J. Biol. Chem. 2012, 287, 3908–3918. [Google Scholar] [CrossRef] [PubMed]
  28. Hass, C.S.; Chen, R.; Wold, M.S. Detection of posttranslational modifications of replication protein A. Methods Mol. Biol. 2012, 922, 193–204. [Google Scholar] [CrossRef]
  29. Chen, R.; Wold, M.S. Replication protein A: Single-stranded DNA’s first responder: Dynamic DNA-interactions allow replication protein A to direct single-strand DNA intermediates into different pathways for synthesis or repair. Bioessays 2014, 36, 1156–1161. [Google Scholar] [CrossRef]
  30. Chen, R.; Subramanyam, S.; Elcock, A.H.; Spies, M.; Wold, M.S. Dynamic binding of replication protein a is required for DNA repair. Nucleic Acids Res. 2016, 44, 5758–5772. [Google Scholar] [CrossRef]
  31. Bain, F.E.; Fischer, L.A.; Chen, R.; Wold, M.S. Single-Molecule Analysis of Replication Protein A-DNA Interactions. Methods Enzymol. 2018, 600, 439–461. [Google Scholar] [CrossRef] [PubMed]
  32. Pokhrel, N.; Caldwell, C.C.; Corless, E.I.; Tillison, E.A.; Tibbs, J.; Jocic, N.; Tabei, S.M.A.; Wold, M.S.; Spies, M.; Antony, E. Dynamics and selective remodeling of the DNA-binding domains of RPA. Nat. Struct. Mol. Biol. 2019, 26, 129–136. [Google Scholar] [CrossRef] [PubMed]
  33. Gall-Duncan, T.; Luo, J.; Jurkovic, C.M.; Fischer, L.A.; Fujita, K.; Deshmukh, A.L.; Harding, R.J.; Tran, S.; Mehkary, M.; Li, V.; et al. Antagonistic roles of canonical and Alternative-RPA in disease-associated tandem CAG repeat instability. Cell 2023, 186, 4898–4919.e25. [Google Scholar] [CrossRef] [PubMed]
  34. Wyka, I.M.; Dhar, K.; Binz, S.K.; Wold, M.S. Replication protein A interactions with DNA: Differential binding of the core domains and analysis of the DNA interaction surface. Biochemistry 2003, 42, 12909–12918. [Google Scholar] [CrossRef]
  35. Caldwell, C.C.; Spies, M. Dynamic elements of replication protein A at the crossroads of DNA replication, recombination, and repair. Crit. Rev. Biochem. Mol. Biol. 2020, 55, 482–507. [Google Scholar] [CrossRef]
  36. He, H.; Wang, J.; Liu, T. UV-Induced RPA1 Acetylation Promotes Nucleotide Excision Repair. Cell Rep. 2017, 20, 2010–2025. [Google Scholar] [CrossRef]
  37. Xia, D.; Zhu, X.; Wang, Y.; Gong, P.; Su, H.S.; Xu, X. Implications of ubiquitination and the maintenance of replication fork stability in cancer therapy. Biosci. Rep. 2023, 43, BSR20222591. [Google Scholar] [CrossRef]
  38. Dou, H.; Huang, C.; Singh, M.; Carpenter, P.B.; Yeh, E.T. Regulation of DNA repair through deSUMOylation and SUMOylation of replication protein A complex. Mol. Cell 2010, 39, 333–345. [Google Scholar] [CrossRef]
  39. Din, S.; Brill, S.J.; Fairman, M.P.; Stillman, B. Cell-cycle-regulated phosphorylation of DNA replication factor A from human and yeast cells. Genes. Dev. 1990, 4, 968–977. [Google Scholar] [CrossRef]
  40. Borgstahl, G.E.; Brader, K.; Mosel, A.; Liu, S.; Kremmer, E.; Goettsch, K.A.; Kolar, C.; Nasheuer, H.P.; Oakley, G.G. Interplay of DNA damage and cell cycle signaling at the level of human replication protein A. DNA Repair 2014, 21, 12–23. [Google Scholar] [CrossRef] [PubMed]
  41. Brush, G.S.; Kelly, T.J. Phosphorylation of the replication protein A large subunit in the Saccharomyces cerevisiae checkpoint response. Nucleic Acids Res. 2000, 28, 3725–3732. [Google Scholar] [CrossRef]
  42. Trovesi, C.; Manfrini, N.; Falcettoni, M.; Longhese, M.P. Regulation of the DNA damage response by cyclin-dependent kinases. J. Mol. Biol. 2013, 425, 4756–4766. [Google Scholar] [CrossRef]
  43. Block, W.D.; Yu, Y.; Lees-Miller, S.P. Phosphatidyl inositol 3-kinase-like serine/threonine protein kinases (PIKKs) are required for DNA damage-induced phosphorylation of the 32 kDa subunit of replication protein A at threonine 21. Nucleic Acids Res. 2004, 32, 997–1005. [Google Scholar] [CrossRef]
  44. Priya, B.; Ravi, S.; Kirubakaran, S. Targeting ATM and ATR for cancer therapeutics: Inhibitors in clinic. Drug Discov. Today 2023, 28, 103662. [Google Scholar] [CrossRef]
  45. Stillman, B.; Bell, S.P.; Dutta, A.; Marahrens, Y. DNA replication and the cell cycle. Ciba Found. Symp. 1992, 170, 147–156; discussion 156–160. [Google Scholar] [CrossRef] [PubMed]
  46. Sancar, A. DNA excision repair. Annu. Rev. Biochem. 1996, 65, 43–81. [Google Scholar] [CrossRef] [PubMed]
  47. Lohrer, H.D. Regulation of the cell cycle following DNA damage in normal and Ataxia telangiectasia cells. Experientia 1996, 52, 316–328. [Google Scholar] [CrossRef] [PubMed]
  48. Iftode, C.; Daniely, Y.; Borowiec, J.A. Replication protein A (RPA): The eukaryotic SSB. Crit. Rev. Biochem. Mol. Biol. 1999, 34, 141–180. [Google Scholar] [CrossRef] [PubMed]
  49. Bochkarev, A.; Bochkareva, E. From RPA to BRCA2: Lessons from single-stranded DNA binding by the OB-fold. Curr. Opin. Struct. Biol. 2004, 14, 36–42. [Google Scholar] [CrossRef]
  50. Zou, Y.; Liu, Y.; Wu, X.; Shell, S.M. Functions of human replication protein A (RPA): From DNA replication to DNA damage and stress responses. J. Cell Physiol. 2006, 208, 267–273. [Google Scholar] [CrossRef]
  51. Fanning, E.; Klimovich, V.; Nager, A.R. A dynamic model for replication protein A (RPA) function in DNA processing pathways. Nucleic Acids Res. 2006, 34, 4126–4137. [Google Scholar] [CrossRef]
  52. Oakley, G.G.; Patrick, S.M. Replication protein A: Directing traffic at the intersection of replication and repair. Front. Biosci. (Landmark Ed.) 2010, 15, 883–900. [Google Scholar] [CrossRef]
  53. Prakash, A.; Borgstahl, G.E. The structure and function of replication protein A in DNA replication. Subcell. Biochem. 2012, 62, 171–196. [Google Scholar] [CrossRef]
  54. Marechal, A.; Zou, L. RPA-coated single-stranded DNA as a platform for post-translational modifications in the DNA damage response. Cell Res. 2015, 25, 9–23. [Google Scholar] [CrossRef] [PubMed]
  55. Liu, T.; Huang, J. Replication protein A and more: Single-stranded DNA-binding proteins in eukaryotic cells. Acta Biochim Biophys Sin 2016, 48, 665–670. [Google Scholar] [CrossRef] [PubMed]
  56. Byrne, B.M.; Oakley, G.G. Replication protein A, the laxative that keeps DNA regular: The importance of RPA phosphorylation in maintaining genome stability. Semin. Cell Dev. Biol. 2019, 86, 112–120. [Google Scholar] [CrossRef] [PubMed]
  57. Anindya, R. Single-stranded DNA damage: Protecting the single-stranded DNA from chemical attack. DNA Repair 2020, 87, 102804. [Google Scholar] [CrossRef] [PubMed]
  58. Rechkunova, N.I.; Lavrik, O.I. Photoreactive DNA as a Tool to Study Replication Protein A Functioning in DNA Replication and Repair. Photochem. Photobiol. 2020, 96, 440–449. [Google Scholar] [CrossRef] [PubMed]
  59. Chowdhury, S.; Chowdhury, A.B.; Kumar, M.; Chakraborty, S. Revisiting regulatory roles of replication protein A in plant DNA metabolism. Planta 2021, 253, 130. [Google Scholar] [CrossRef] [PubMed]
  60. Barbour, A.T.; Wuttke, D.S. RPA-like single-stranded DNA-binding protein complexes including CST serve as specialized processivity factors for polymerases. Curr. Opin. Struct. Biol. 2023, 81, 102611. [Google Scholar] [CrossRef]
  61. Kusi-Appauh, N.; Ralph, S.F.; van Oijen, A.M.; Spenkelink, L.M. Understanding G-Quadruplex Biology and Stability Using Single-Molecule Techniques. J. Phys. Chem. B 2023, 127, 5521–5540. [Google Scholar] [CrossRef] [PubMed]
  62. Nasheuer, H.P.; Meaney, A.M.; Hulshoff, T.; Thiele, I.; Onwubiko, N.O. Replication Protein A, the Main Eukaryotic Single-Stranded DNA Binding Protein, a Focal Point in Cellular DNA Metabolism. Int. J. Mol. Sci. 2024, 25, 588. [Google Scholar] [CrossRef] [PubMed]
  63. Bochkareva, E.; Belegu, V.; Korolev, S.; Bochkarev, A. Structure of the major single-stranded DNA-binding domain of replication protein A suggests a dynamic mechanism for DNA binding. EMBO J. 2001, 20, 612–618. [Google Scholar] [CrossRef] [PubMed]
  64. Lin, Y.; Lin, L.J.; Sriratana, P.; Coleman, K.; Ha, T.; Spies, M.; Cann, I.K. Engineering of functional replication protein a homologs based on insights into the evolution of oligonucleotide/oligosaccharide-binding folds. J. Bacteriol. 2008, 190, 5766–5780. [Google Scholar] [CrossRef] [PubMed]
  65. Fernandes, C.A.H.; Morea, E.G.O.; Cano, M.I.N. RPA-1 from Leishmania sp.: Recombinant Protein Expression and Purification, Molecular Modeling, and Molecular Dynamics Simulations Protocols. Methods Mol. Biol. 2021, 2281, 169–191. [Google Scholar] [CrossRef] [PubMed]
  66. Bochkarev, A.; Pfuetzner, R.A.; Edwards, A.M.; Frappier, L. Structure of the single-stranded-DNA-binding domain of replication protein A bound to DNA. Nature 1997, 385, 176–181. [Google Scholar] [CrossRef] [PubMed]
  67. Fan, J.; Pavletich, N.P. Structure and conformational change of a replication protein A heterotrimer bound to ssDNA. Genes. Dev. 2012, 26, 2337–2347. [Google Scholar] [CrossRef]
  68. Jacobs, D.M.; Lipton, A.S.; Isern, N.G.; Daughdrill, G.W.; Lowry, D.F.; Gomes, X.; Wold, M.S. Human replication protein A: Global fold of the N-terminal RPA-70 domain reveals a basic cleft and flexible C-terminal linker. J. Biomol. NMR 1999, 14, 321–331. [Google Scholar] [CrossRef]
  69. Mer, G.; Bochkarev, A.; Gupta, R.; Bochkareva, E.; Frappier, L.; Ingles, C.J.; Edwards, A.M.; Chazin, W.J. Structural basis for the recognition of DNA repair proteins UNG2, XPA, and RAD52 by replication factor RPA. Cell 2000, 103, 449–456. [Google Scholar] [CrossRef]
  70. Zhou, C.; Pourmal, S.; Pavletich, N.P. Dna2 nuclease-helicase structure, mechanism and regulation by Rpa. Elife 2015, 4, e09832. [Google Scholar] [CrossRef]
  71. Bochkareva, E.; Korolev, S.; Lees-Miller, S.P.; Bochkarev, A. Structure of the RPA trimerization core and its role in the multistep DNA-binding mechanism of RPA. EMBO J. 2002, 21, 1855–1863. [Google Scholar] [CrossRef]
  72. Bochkareva, E.; Kaustov, L.; Ayed, A.; Yi, G.S.; Lu, Y.; Pineda-Lucena, A.; Liao, J.C.; Okorokov, A.L.; Milner, J.; Arrowsmith, C.H.; et al. Single-stranded DNA mimicry in the p53 transactivation domain interaction with replication protein A. Proc. Natl. Acad. Sci. USA 2005, 102, 15412–15417. [Google Scholar] [CrossRef]
  73. Frank, A.O.; Feldkamp, M.D.; Kennedy, J.P.; Waterson, A.G.; Pelz, N.F.; Patrone, J.D.; Vangamudi, B.; Camper, D.V.; Rossanese, O.W.; Chazin, W.J.; et al. Discovery of a potent inhibitor of replication protein a protein-protein interactions using a fragment-linking approach. J. Med. Chem. 2013, 56, 9242–9250. [Google Scholar] [CrossRef]
  74. Guilliam, T.A.; Brissett, N.C.; Ehlinger, A.; Keen, B.A.; Kolesar, P.; Taylor, E.M.; Bailey, L.J.; Lindsay, H.D.; Chazin, W.J.; Doherty, A.J. Molecular basis for PrimPol recruitment to replication forks by RPA. Nat. Commun. 2017, 8, 15222. [Google Scholar] [CrossRef]
  75. Deng, X.; Habel, J.E.; Kabaleeswaran, V.; Snell, E.H.; Wold, M.S.; Borgstahl, G.E. Structure of the full-length human RPA14/32 complex gives insights into the mechanism of DNA binding and complex formation. J. Mol. Biol. 2007, 374, 865–876. [Google Scholar] [CrossRef]
  76. Wu, Y.; Fu, W.; Zang, N.; Zhou, C. Structural characterization of human RPA70N association with DNA damage response proteins. Elife 2023, 12, e81639. [Google Scholar] [CrossRef]
  77. Patrone, J.D.; Pelz, N.F.; Bates, B.S.; Souza-Fagundes, E.M.; Vangamudi, B.; Camper, D.V.; Kuznetsov, A.G.; Browning, C.F.; Feldkamp, M.D.; Frank, A.O.; et al. Identification and Optimization of Anthranilic Acid Based Inhibitors of Replication Protein A. ChemMedChem 2016, 11, 893–899. [Google Scholar] [CrossRef] [PubMed]
  78. Waterson, A.G.; Kennedy, J.P.; Patrone, J.D.; Pelz, N.F.; Feldkamp, M.D.; Frank, A.O.; Vangamudi, B.; Souza-Fagundes, E.M.; Rossanese, O.W.; Chazin, W.J.; et al. Diphenylpyrazoles as replication protein a inhibitors. ACS Med. Chem. Lett. 2015, 6, 140–145. [Google Scholar] [CrossRef] [PubMed]
  79. Feldkamp, M.D.; Frank, A.O.; Kennedy, J.P.; Patrone, J.D.; Vangamudi, B.; Waterson, A.G.; Fesik, S.W.; Chazin, W.J. Surface reengineering of RPA70N enables cocrystallization with an inhibitor of the replication protein A interaction motif of ATR interacting protein. Biochemistry 2013, 52, 6515–6524. [Google Scholar] [CrossRef] [PubMed]
  80. Patrone, J.D.; Kennedy, J.P.; Frank, A.O.; Feldkamp, M.D.; Vangamudi, B.; Pelz, N.F.; Rossanese, O.W.; Waterson, A.G.; Chazin, W.J.; Fesik, S.W. Discovery of Protein-Protein Interaction Inhibitors of Replication Protein A. ACS Med. Chem. Lett. 2013, 4, 601–605. [Google Scholar] [CrossRef] [PubMed]
  81. Bochkarev, A.; Bochkareva, E.; Frappier, L.; Edwards, A.M. The crystal structure of the complex of replication protein A subunits RPA32 and RPA14 reveals a mechanism for single-stranded DNA binding. EMBO J. 1999, 18, 4498–4504. [Google Scholar] [CrossRef] [PubMed]
  82. Feldkamp, M.D.; Mason, A.C.; Eichman, B.F.; Chazin, W.J. Structural analysis of replication protein A recruitment of the DNA damage response protein SMARCAL1. Biochemistry 2014, 53, 3052–3061. [Google Scholar] [CrossRef] [PubMed]
  83. Madru, C.; Martinez-Carranza, M.; Laurent, S.; Alberti, A.C.; Chevreuil, M.; Raynal, B.; Haouz, A.; Le Meur, R.A.; Delarue, M.; Henneke, G.; et al. DNA-binding mechanism and evolution of replication protein A. Nat. Commun. 2023, 14, 2326. [Google Scholar] [CrossRef] [PubMed]
  84. Brosey, C.A.; Soss, S.E.; Brooks, S.; Yan, C.; Ivanov, I.; Dorai, K.; Chazin, W.J. Functional dynamics in replication protein A DNA binding and protein recruitment domains. Structure 2015, 23, 1028–1038. [Google Scholar] [CrossRef] [PubMed]
  85. Pretto, D.I.; Tsutakawa, S.; Brosey, C.A.; Castillo, A.; Chagot, M.E.; Smith, J.A.; Tainer, J.A.; Chazin, W.J. Structural dynamics and single-stranded DNA binding activity of the three N-terminal domains of the large subunit of replication protein A from small angle X-ray scattering. Biochemistry 2010, 49, 2880–2889. [Google Scholar] [CrossRef] [PubMed]
  86. Powers, K.T.; Gildenberg, M.S.; Washington, M.T. Modeling Conformationally Flexible Proteins With X-ray Scattering and Molecular Simulations. Comput. Struct. Biotechnol. J. 2019, 17, 570–578. [Google Scholar] [CrossRef] [PubMed]
  87. Dutta, A.; Din, S.; Brill, S.J.; Stillman, B. Phosphorylation of replication protein A: A role for cdc2 kinase in G1/S regulation. Cold Spring Harb. Symp. Quant. Biol. 1991, 56, 315–324. [Google Scholar] [CrossRef]
  88. Murti, K.G.; He, D.C.; Brinkley, B.R.; Scott, R.; Lee, S.H. Dynamics of human replication protein A subunit distribution and partitioning in the cell cycle. Exp. Cell Res. 1996, 223, 279–289. [Google Scholar] [CrossRef]
  89. Thoma, B.S.; Vasquez, K.M. Critical DNA damage recognition functions of XPC-hHR23B and XPA-RPA in nucleotide excision repair. Mol. Carcinog. 2003, 38, 1–13. [Google Scholar] [CrossRef]
  90. Zhao, M.; Geng, R.; Guo, X.; Yuan, R.; Zhou, X.; Zhong, Y.; Huo, Y.; Zhou, M.; Shen, Q.; Li, Y.; et al. PCAF/GCN5-Mediated Acetylation of RPA1 Promotes Nucleotide Excision Repair. Cell Rep. 2017, 20, 1997–2009. [Google Scholar] [CrossRef]
  91. Hayran, A.B.; Liabakk, N.B.; Aas, P.A.; Kusnierczyk, A.; Vagbo, C.B.; Sarno, A.; Iveland, T.S.; Chawla, K.; Zahn, A.; Di Noia, J.M.; et al. RPA guides UNG to uracil in ssDNA to facilitate antibody class switching and repair of mutagenic uracil at the replication fork. Nucleic Acids Res. 2024, 52, 784–800. [Google Scholar] [CrossRef]
  92. Gavande, N.S.; VanderVere-Carozza, P.S.; Hinshaw, H.D.; Jalal, S.I.; Sears, C.R.; Pawelczak, K.S.; Turchi, J.J. DNA repair targeted therapy: The past or future of cancer treatment? Pharmacol. Ther. 2016, 160, 65–83. [Google Scholar] [CrossRef]
  93. Deng, X.; Prakash, A.; Dhar, K.; Baia, G.S.; Kolar, C.; Oakley, G.G.; Borgstahl, G.E. Human replication protein A-Rad52-single-stranded DNA complex: Stoichiometry and evidence for strand transfer regulation by phosphorylation. Biochemistry 2009, 48, 6633–6643. [Google Scholar] [CrossRef]
  94. Wright, W.D.; Shah, S.S.; Heyer, W.D. Homologous recombination and the repair of DNA double-strand breaks. J. Biol. Chem. 2018, 293, 10524–10535. [Google Scholar] [CrossRef] [PubMed]
  95. Creeden, J.F.; Nanavaty, N.S.; Einloth, K.R.; Gillman, C.E.; Stanbery, L.; Hamouda, D.M.; Dworkin, L.; Nemunaitis, J. Homologous recombination proficiency in ovarian and breast cancer patients. BMC Cancer 2021, 21, 1154. [Google Scholar] [CrossRef] [PubMed]
  96. Prakash, A.; Natarajan, A.; Marky, L.A.; Ouellette, M.M.; Borgstahl, G.E. Identification of the DNA-Binding Domains of Human Replication Protein A That Recognize G-Quadruplex DNA. J. Nucleic Acids 2011, 2011, 896947. [Google Scholar] [CrossRef]
  97. Lysetska, M.; Knoll, A.; Boehringer, D.; Hey, T.; Krauss, G.; Krausch, G. UV light-damaged DNA and its interaction with human replication protein A: An atomic force microscopy study. Nucleic Acids Res. 2002, 30, 2686–2691. [Google Scholar] [CrossRef] [PubMed]
  98. You, J.S.; Wang, M.; Lee, S.H. Functional characterization of zinc-finger motif in redox regulation of RPA-ssDNA interaction. Biochemistry 2000, 39, 12953–12958. [Google Scholar] [CrossRef]
  99. Prakash, A.; Kieken, F.; Marky, L.A.; Borgstahl, G.E. Stabilization of a G-Quadruplex from Unfolding by Replication Protein A Using Potassium and the Porphyrin TMPyP4. J. Nucleic Acids 2011, 2011, 529828. [Google Scholar] [CrossRef]
  100. Ding, J.; Li, X.; Shen, J.; Zhao, Y.; Zhong, S.; Lai, L.; Niu, H.; Qi, Z. ssDNA accessibility of Rad51 is regulated by orchestrating multiple RPA dynamics. Nat. Commun. 2023, 14, 3864. [Google Scholar] [CrossRef]
  101. Kang, Y.; Han, Y.G.; Khim, K.W.; Choi, W.G.; Ju, M.K.; Park, K.; Shin, K.J.; Chae, Y.C.; Choi, J.H.; Kim, H.; et al. Alteration of replication protein A binding mode on single-stranded DNA by NSMF potentiates RPA phosphorylation by ATR kinase. Nucleic Acids Res. 2023, 51, 7936–7950. [Google Scholar] [CrossRef]
  102. Braun, K.A.; Lao, Y.; He, Z.; Ingles, C.J.; Wold, M.S. Role of protein-protein interactions in the function of replication protein A (RPA): RPA modulates the activity of DNA polymerase Alpha by multiple mechanisms. Biochemistry 1997, 36, 8443–8454. [Google Scholar] [CrossRef]
  103. Golub, E.I.; Gupta, R.C.; Haaf, T.; Wold, M.S.; Radding, C.M. Interaction of human rad51 recombination protein with single-stranded DNA binding protein, RPA. Nucleic Acids Res. 1998, 26, 5388–5393. [Google Scholar] [CrossRef]
  104. Riva, F.; Zuco, V.; Vink, A.A.; Supino, R.; Prosperi, E. UV-induced DNA incision and proliferating cell nuclear antigen recruitment to repair sites occur independently of p53-replication protein A interaction in p53 wild type and mutant ovarian carcinoma cells. Carcinogenesis 2001, 22, 1971–1978. [Google Scholar] [CrossRef]
  105. Wang, M.; Mahrenholz, A.; Lee, S.H. RPA stabilizes the XPA-damaged DNA complex through protein-protein interaction. Biochemistry 2000, 39, 6433–6439. [Google Scholar] [CrossRef]
  106. Yoo, E.; Kim, B.U.; Lee, S.Y.; Cho, C.H.; Chung, J.H.; Lee, C.H. 53BP1 is associated with replication protein A and is required for RPA2 hyperphosphorylation following DNA damage. Oncogene 2005, 24, 5423–5430. [Google Scholar] [CrossRef]
  107. Chaudhuri, J.; Khuong, C.; Alt, F.W. Replication protein A interacts with AID to promote deamination of somatic hypermutation targets. Nature 2004, 430, 992–998. [Google Scholar] [CrossRef] [PubMed]
  108. Kalan, S.; Matveyenko, A.; Loayza, D. LIM Protein Ajuba Participates in the Repression of the ATR-Mediated DNA Damage Response. Front. Genet. 2013, 4, 95. [Google Scholar] [CrossRef]
  109. Fowler, S.; Maguin, P.; Kalan, S.; Loayza, D. LIM Protein Ajuba associates with the RPA complex through direct cell cycle-dependent interaction with the RPA70 subunit. Sci. Rep. 2018, 8, 9536. [Google Scholar] [CrossRef] [PubMed]
  110. Zou, L.; Elledge, S.J. Sensing DNA damage through ATRIP recognition of RPA-ssDNA complexes. Science 2003, 300, 1542–1548. [Google Scholar] [CrossRef] [PubMed]
  111. Ball, H.L.; Myers, J.S.; Cortez, D. ATRIP binding to replication protein A-single-stranded DNA promotes ATR-ATRIP localization but is dispensable for Chk1 phosphorylation. Mol. Biol. Cell 2005, 16, 2372–2381. [Google Scholar] [CrossRef]
  112. Namiki, Y.; Zou, L. ATRIP associates with replication protein A-coated ssDNA through multiple interactions. Proc. Natl. Acad. Sci. USA 2006, 103, 580–585. [Google Scholar] [CrossRef] [PubMed]
  113. Liu, Y.; Vaithiyalingam, S.; Shi, Q.; Chazin, W.J.; Zinkel, S.S. BID binds to replication protein A and stimulates ATR function following replicative stress. Mol. Cell Biol. 2011, 31, 4298–4309. [Google Scholar] [CrossRef] [PubMed]
  114. Brosh, R.M., Jr.; Li, J.L.; Kenny, M.K.; Karow, J.K.; Cooper, M.P.; Kureekattil, R.P.; Hickson, I.D.; Bohr, V.A. Replication protein A physically interacts with the Bloom’s syndrome protein and stimulates its helicase activity. J. Biol. Chem. 2000, 275, 23500–23508. [Google Scholar] [CrossRef] [PubMed]
  115. Doherty, K.M.; Sommers, J.A.; Gray, M.D.; Lee, J.W.; von Kobbe, C.; Thoma, N.H.; Kureekattil, R.P.; Kenny, M.K.; Brosh, R.M., Jr. Physical and functional mapping of the replication protein a interaction domain of the werner and bloom syndrome helicases. J. Biol. Chem. 2005, 280, 29494–29505. [Google Scholar] [CrossRef] [PubMed]
  116. Blackwell, L.J.; Borowiec, J.A.; Mastrangelo, I.A. Single-stranded-DNA binding alters human replication protein A structure and facilitates interaction with DNA-dependent protein kinase. Mol. Cell Biol. 1996, 16, 4798–4807. [Google Scholar] [CrossRef] [PubMed]
  117. Feng, S.; Zhao, Y.; Xu, Y.; Ning, S.; Huo, W.; Hou, M.; Gao, G.; Ji, J.; Guo, R.; Xu, D. Ewing Tumor-associated Antigen 1 Interacts with Replication Protein A to Promote Restart of Stalled Replication Forks. J. Biol. Chem. 2016, 291, 21956–21962. [Google Scholar] [CrossRef]
  118. Haahr, P.; Hoffmann, S.; Tollenaere, M.A.; Ho, T.; Toledo, L.I.; Mann, M.; Bekker-Jensen, S.; Raschle, M.; Mailand, N. Activation of the ATR kinase by the RPA-binding protein ETAA1. Nat. Cell Biol. 2016, 18, 1196–1207. [Google Scholar] [CrossRef]
  119. Bass, T.E.; Luzwick, J.W.; Kavanaugh, G.; Carroll, C.; Dungrawala, H.; Glick, G.G.; Feldkamp, M.D.; Putney, R.; Chazin, W.J.; Cortez, D. ETAA1 acts at stalled replication forks to maintain genome integrity. Nat. Cell Biol. 2016, 18, 1185–1195. [Google Scholar] [CrossRef]
  120. Lee, Y.C.; Zhou, Q.; Chen, J.; Yuan, J. RPA-Binding Protein ETAA1 Is an ATR Activator Involved in DNA Replication Stress Response. Curr. Biol. 2016, 26, 3257–3268. [Google Scholar] [CrossRef]
  121. Zhao, W.; Vaithiyalingam, S.; San Filippo, J.; Maranon, D.G.; Jimenez-Sainz, J.; Fontenay, G.V.; Kwon, Y.; Leung, S.G.; Lu, L.; Jensen, R.B.; et al. Promotion of BRCA2-Dependent Homologous Recombination by DSS1 via RPA Targeting and DNA Mimicry. Mol. Cell 2015, 59, 176–187. [Google Scholar] [CrossRef] [PubMed]
  122. Fujimoto, M.; Takaki, E.; Takii, R.; Tan, K.; Prakasam, R.; Hayashida, N.; Iemura, S.; Natsume, T.; Nakai, A. RPA assists HSF1 access to nucleosomal DNA by recruiting histone chaperone FACT. Mol. Cell 2012, 48, 182–194. [Google Scholar] [CrossRef] [PubMed]
  123. Gupta, R.; Sharma, S.; Sommers, J.A.; Kenny, M.K.; Cantor, S.B.; Brosh, R.M., Jr. FANCJ (BACH1) helicase forms DNA damage inducible foci with replication protein A and interacts physically and functionally with the single-stranded DNA-binding protein. Blood 2007, 110, 2390–2398. [Google Scholar] [CrossRef] [PubMed]
  124. Lai, Y.; Zhu, M.; Wu, W.; Rokutanda, N.; Togashi, Y.; Liang, W.; Ohta, T. HERC2 regulates RPA2 by mediating ATR-induced Ser33 phosphorylation and ubiquitin-dependent degradation. Sci. Rep. 2019, 9, 14257. [Google Scholar] [CrossRef] [PubMed]
  125. Wu, W.; Rokutanda, N.; Takeuchi, J.; Lai, Y.; Maruyama, R.; Togashi, Y.; Nishikawa, H.; Arai, N.; Miyoshi, Y.; Suzuki, N.; et al. HERC2 Facilitates BLM and WRN Helicase Complex Interaction with RPA to Suppress G-Quadruplex DNA. Cancer Res. 2018, 78, 6371–6385. [Google Scholar] [CrossRef] [PubMed]
  126. Zhang, H.; Gan, H.; Wang, Z.; Lee, J.H.; Zhou, H.; Ordog, T.; Wold, M.S.; Ljungman, M.; Zhang, Z. RPA Interacts with HIRA and Regulates H3.3 Deposition at Gene Regulatory Elements in Mammalian Cells. Mol. Cell 2017, 65, 272–284. [Google Scholar] [CrossRef] [PubMed]
  127. Liu, S.; Xu, Z.; Leng, H.; Zheng, P.; Yang, J.; Chen, K.; Feng, J.; Li, Q. RPA binds histone H3-H4 and functions in DNA replication-coupled nucleosome assembly. Science 2017, 355, 415–420. [Google Scholar] [CrossRef]
  128. Dueva, R.; Iliakis, G. Replication protein A: A multifunctional protein with roles in DNA replication, repair and beyond. NAR Cancer 2020, 2, zcaa022. [Google Scholar] [CrossRef]
  129. Sukhodolets, K.E.; Hickman, A.B.; Agarwal, S.K.; Sukhodolets, M.V.; Obungu, V.H.; Novotny, E.A.; Crabtree, J.S.; Chandrasekharappa, S.C.; Collins, F.S.; Spiegel, A.M.; et al. The 32-kilodalton subunit of replication protein A interacts with menin, the product of the MEN1 tumor suppressor gene. Mol. Cell Biol. 2003, 23, 493–509. [Google Scholar] [CrossRef]
  130. Chen, C.C.; Juan, C.W.; Chen, K.Y.; Chang, Y.C.; Lee, J.C.; Chang, M.C. Upregulation of RPA2 promotes NF-kappaB activation in breast cancer by relieving the antagonistic function of menin on NF-kappaB-regulated transcription. Carcinogenesis 2017, 38, 196–206. [Google Scholar] [CrossRef]
  131. Robison, J.G.; Elliott, J.; Dixon, K.; Oakley, G.G. Replication protein A and the Mre11.Rad50.Nbs1 complex co-localize and interact at sites of stalled replication forks. J. Biol. Chem. 2004, 279, 34802–34810. [Google Scholar] [CrossRef]
  132. Oakley, G.G.; Tillison, K.; Opiyo, S.A.; Glanzer, J.G.; Horn, J.M.; Patrick, S.M. Physical interaction between replication protein A (RPA) and MRN: Involvement of RPA2 phosphorylation and the N-terminus of RPA1. Biochemistry 2009, 48, 7473–7481. [Google Scholar] [CrossRef]
  133. Daniely, Y.; Borowiec, J.A. Formation of a complex between nucleolin and replication protein A after cell stress prevents initiation of DNA replication. J. Cell Biol. 2000, 149, 799–810. [Google Scholar] [CrossRef]
  134. Kim, K.; Dimitrova, D.D.; Carta, K.M.; Saxena, A.; Daras, M.; Borowiec, J.A. Novel checkpoint response to genotoxic stress mediated by nucleolin-replication protein a complex formation. Mol. Cell Biol. 2005, 25, 2463–2474. [Google Scholar] [CrossRef]
  135. Wang, Y.; Guan, J.; Wang, H.; Wang, Y.; Leeper, D.; Iliakis, G. Regulation of DNA replication after heat shock by replication protein a-nucleolin interactions. J. Biol. Chem. 2001, 276, 20579–20588. [Google Scholar] [CrossRef] [PubMed]
  136. Li, R.; Botchan, M.R. The acidic transcriptional activation domains of VP16 and p53 bind the cellular replication protein A and stimulate in vitro BPV-1 DNA replication. Cell 1993, 73, 1207–1221. [Google Scholar] [CrossRef] [PubMed]
  137. He, Z.; Brinton, B.T.; Greenblatt, J.; Hassell, J.A.; Ingles, C.J. The transactivator proteins VP16 and GAL4 bind replication factor A. Cell 1993, 73, 1223–1232. [Google Scholar] [CrossRef] [PubMed]
  138. Dutta, A.; Ruppert, J.M.; Aster, J.C.; Winchester, E. Inhibition of DNA replication factor RPA by p53. Nature 1993, 365, 79–82. [Google Scholar] [CrossRef]
  139. Romanova, L.Y.; Willers, H.; Blagosklonny, M.V.; Powell, S.N. The interaction of p53 with replication protein A mediates suppression of homologous recombination. Oncogene 2004, 23, 9025–9033. [Google Scholar] [CrossRef] [PubMed]
  140. Serrano, M.A.; Li, Z.; Dangeti, M.; Musich, P.R.; Patrick, S.; Roginskaya, M.; Cartwright, B.; Zou, Y. DNA-PK, ATM and ATR collaboratively regulate p53-RPA interaction to facilitate homologous recombination DNA repair. Oncogene 2013, 32, 2452–2462. [Google Scholar] [CrossRef] [PubMed]
  141. Murphy, A.K.; Fitzgerald, M.; Ro, T.; Kim, J.H.; Rabinowitsch, A.I.; Chowdhury, D.; Schildkraut, C.L.; Borowiec, J.A. Phosphorylated RPA recruits PALB2 to stalled DNA replication forks to facilitate fork recovery. J. Cell Biol. 2014, 206, 493–507. [Google Scholar] [CrossRef]
  142. Wan, L.; Lou, J.; Xia, Y.; Su, B.; Liu, T.; Cui, J.; Sun, Y.; Lou, H.; Huang, J. hPrimpol1/CCDC111 is a human DNA primase-polymerase required for the maintenance of genome integrity. EMBO Rep. 2013, 14, 1104–1112. [Google Scholar] [CrossRef]
  143. Guilliam, T.A.; Jozwiakowski, S.K.; Ehlinger, A.; Barnes, R.P.; Rudd, S.G.; Bailey, L.J.; Skehel, J.M.; Eckert, K.A.; Chazin, W.J.; Doherty, A.J. Human PrimPol is a highly error-prone polymerase regulated by single-stranded DNA binding proteins. Nucleic Acids Res. 2015, 43, 1056–1068. [Google Scholar] [CrossRef]
  144. Marechal, A.; Li, J.M.; Ji, X.Y.; Wu, C.S.; Yazinski, S.A.; Nguyen, H.D.; Liu, S.; Jimenez, A.E.; Jin, J.; Zou, L. PRP19 transforms into a sensor of RPA-ssDNA after DNA damage and drives ATR activation via a ubiquitin-mediated circuitry. Mol. Cell 2014, 53, 235–246. [Google Scholar] [CrossRef] [PubMed]
  145. Wang, G.; Li, Y.; Wang, P.; Liang, H.; Cui, M.; Zhu, M.; Guo, L.; Su, Q.; Sun, Y.; McNutt, M.A.; et al. PTEN regulates RPA1 and protects DNA replication forks. Cell Res. 2015, 25, 1189–1204. [Google Scholar] [CrossRef] [PubMed]
  146. Wu, X.; Shell, S.M.; Zou, Y. Interaction and colocalization of Rad9/Rad1/Hus1 checkpoint complex with replication protein A in human cells. Oncogene 2005, 24, 4728–4735. [Google Scholar] [CrossRef] [PubMed]
  147. Zou, L.; Liu, D.; Elledge, S.J. Replication protein A-mediated recruitment and activation of Rad17 complexes. Proc. Natl. Acad. Sci. USA 2003, 100, 13827–13832. [Google Scholar] [CrossRef]
  148. Hedglin, M.; Aitha, M.; Pedley, A.; Benkovic, S.J. Replication protein A dynamically regulates monoubiquitination of proliferating cell nuclear antigen. J. Biol. Chem. 2019, 294, 5157–5168. [Google Scholar] [CrossRef]
  149. Wu, X.; Yang, Z.; Liu, Y.; Zou, Y. Preferential localization of hyperphosphorylated replication protein A to double-strand break repair and checkpoint complexes upon DNA damage. Biochem. J. 2005, 391, 473–480. [Google Scholar] [CrossRef]
  150. Ma, C.J.; Kwon, Y.; Sung, P.; Greene, E.C. Human RAD52 interactions with replication protein A and the RAD51 presynaptic complex. J. Biol. Chem. 2017, 292, 11702–11713. [Google Scholar] [CrossRef]
  151. Stauffer, M.E.; Chazin, W.J. Physical interaction between replication protein A and Rad51 promotes exchange on single-stranded DNA. J. Biol. Chem. 2004, 279, 25638–25645. [Google Scholar] [CrossRef]
  152. Park, M.S.; Ludwig, D.L.; Stigger, E.; Lee, S.H. Physical interaction between human RAD52 and RPA is required for homologous recombination in mammalian cells. J. Biol. Chem. 1996, 271, 18996–19000. [Google Scholar] [CrossRef]
  153. Shinohara, A.; Shinohara, M.; Ohta, T.; Matsuda, S.; Ogawa, T. Rad52 forms ring structures and co-operates with RPA in single-strand DNA annealing. Genes. Cells 1998, 3, 145–156. [Google Scholar] [CrossRef]
  154. Sugiyama, T.; Kowalczykowski, S.C. Rad52 protein associates with replication protein A (RPA)-single-stranded DNA to accelerate Rad51-mediated displacement of RPA and presynaptic complex formation. J. Biol. Chem. 2002, 277, 31663–31672. [Google Scholar] [CrossRef]
  155. Plate, I.; Hallwyl, S.C.; Shi, I.; Krejci, L.; Muller, C.; Albertsen, L.; Sung, P.; Mortensen, U.H. Interaction with RPA is necessary for Rad52 repair center formation and for its mediator activity. J. Biol. Chem. 2008, 283, 29077–29085. [Google Scholar] [CrossRef]
  156. Seong, C.; Sehorn, M.G.; Plate, I.; Shi, I.; Song, B.; Chi, P.; Mortensen, U.; Sung, P.; Krejci, L. Molecular anatomy of the recombination mediator function of Saccharomyces cerevisiae Rad52. J. Biol. Chem. 2008, 283, 12166–12174. [Google Scholar] [CrossRef] [PubMed]
  157. Liu, S.; Chu, J.; Yucer, N.; Leng, M.; Wang, S.Y.; Chen, B.P.; Hittelman, W.N.; Wang, Y. RING finger and WD repeat domain 3 (RFWD3) associates with replication protein A (RPA) and facilitates RPA-mediated DNA damage response. J. Biol. Chem. 2011, 286, 22314–22322. [Google Scholar] [CrossRef] [PubMed]
  158. Gong, Z.; Chen, J. E3 ligase RFWD3 participates in replication checkpoint control. J. Biol. Chem. 2011, 286, 22308–22313. [Google Scholar] [CrossRef] [PubMed]
  159. Nguyen, H.D.; Yadav, T.; Giri, S.; Saez, B.; Graubert, T.A.; Zou, L. Functions of Replication Protein A as a Sensor of R Loops and a Regulator of RNaseH1. Mol. Cell 2017, 65, 832–847.e834. [Google Scholar] [CrossRef] [PubMed]
  160. Yusufzai, T.; Kong, X.; Yokomori, K.; Kadonaga, J.T. The annealing helicase HARP is recruited to DNA repair sites via an interaction with RPA. Genes. Dev. 2009, 23, 2400–2404. [Google Scholar] [CrossRef] [PubMed]
  161. Ciccia, A.; Bredemeyer, A.L.; Sowa, M.E.; Terret, M.E.; Jallepalli, P.V.; Harper, J.W.; Elledge, S.J. The SIOD disorder protein SMARCAL1 is an RPA-interacting protein involved in replication fork restart. Genes. Dev. 2009, 23, 2415–2425. [Google Scholar] [CrossRef] [PubMed]
  162. Yuan, J.; Ghosal, G.; Chen, J. The annealing helicase HARP protects stalled replication forks. Genes. Dev. 2009, 23, 2394–2399. [Google Scholar] [CrossRef] [PubMed]
  163. Postow, L.; Woo, E.M.; Chait, B.T.; Funabiki, H. Identification of SMARCAL1 as a component of the DNA damage response. J. Biol. Chem. 2009, 284, 35951–35961. [Google Scholar] [CrossRef] [PubMed]
  164. Kemp, M.G.; Akan, Z.; Yilmaz, S.; Grillo, M.; Smith-Roe, S.L.; Kang, T.H.; Cordeiro-Stone, M.; Kaufmann, W.K.; Abraham, R.T.; Sancar, A.; et al. Tipin-replication protein A interaction mediates Chk1 phosphorylation by ATR in response to genotoxic stress. J. Biol. Chem. 2010, 285, 16562–16571. [Google Scholar] [CrossRef] [PubMed]
  165. Unsal-Kacmaz, K.; Chastain, P.D.; Qu, P.P.; Minoo, P.; Cordeiro-Stone, M.; Sancar, A.; Kaufmann, W.K. The human Tim/Tipin complex coordinates an Intra-S checkpoint response to UV that slows replication fork displacement. Mol. Cell Biol. 2007, 27, 3131–3142. [Google Scholar] [CrossRef] [PubMed]
  166. Nagelhus, T.A.; Haug, T.; Singh, K.K.; Keshav, K.F.; Skorpen, F.; Otterlei, M.; Bharati, S.; Lindmo, T.; Benichou, S.; Benarous, R.; et al. A sequence in the N-terminal region of human uracil-DNA glycosylase with homology to XPA interacts with the C-terminal part of the 34-kDa subunit of replication protein A. J. Biol. Chem. 1997, 272, 6561–6566. [Google Scholar] [CrossRef] [PubMed]
  167. Machwe, A.; Lozada, E.; Wold, M.S.; Li, G.M.; Orren, D.K. Molecular cooperation between the Werner syndrome protein and replication protein A in relation to replication fork blockage. J. Biol. Chem. 2011, 286, 3497–3508. [Google Scholar] [CrossRef]
  168. Hyun, M.; Park, S.; Kim, E.; Kim, D.H.; Lee, S.J.; Koo, H.S.; Seo, Y.S.; Ahn, B. Physical and functional interactions of Caenorhabditis elegans WRN-1 helicase with RPA-1. Biochemistry 2012, 51, 1336–1345. [Google Scholar] [CrossRef]
  169. Brosh, R.M., Jr.; Orren, D.K.; Nehlin, J.O.; Ravn, P.H.; Kenny, M.K.; Machwe, A.; Bohr, V.A. Functional and physical interaction between WRN helicase and human replication protein A. J. Biol. Chem. 1999, 274, 18341–18350. [Google Scholar] [CrossRef]
  170. Shen, J.C.; Lao, Y.; Kamath-Loeb, A.; Wold, M.S.; Loeb, L.A. The N-terminal domain of the large subunit of human replication protein A binds to Werner syndrome protein and stimulates helicase activity. Mech. Ageing Dev. 2003, 124, 921–930. [Google Scholar] [CrossRef]
  171. Li, L.; Lu, X.; Peterson, C.A.; Legerski, R.J. An interaction between the DNA repair factor XPA and replication protein A appears essential for nucleotide excision repair. Mol. Cell Biol. 1995, 15, 5396–5402. [Google Scholar] [CrossRef]
  172. Saijo, M.; Kuraoka, I.; Masutani, C.; Hanaoka, F.; Tanaka, K. Sequential binding of DNA repair proteins RPA and ERCC1 to XPA in vitro. Nucleic Acids Res. 1996, 24, 4719–4724. [Google Scholar] [CrossRef] [PubMed]
  173. Matsuda, T.; Saijo, M.; Kuraoka, I.; Kobayashi, T.; Nakatsu, Y.; Nagai, A.; Enjoji, T.; Masutani, C.; Sugasawa, K.; Hanaoka, F.; et al. DNA repair protein XPA binds replication protein A (RPA). J. Biol. Chem. 1995, 270, 4152–4157. [Google Scholar] [CrossRef]
  174. de Laat, W.L.; Appeldoorn, E.; Sugasawa, K.; Weterings, E.; Jaspers, N.G.; Hoeijmakers, J.H. DNA-binding polarity of human replication protein A positions nucleases in nucleotide excision repair. Genes. Dev. 1998, 12, 2598–2609. [Google Scholar] [CrossRef]
  175. Matsunaga, T.; Park, C.H.; Bessho, T.; Mu, D.; Sancar, A. Replication protein A confers structure-specific endonuclease activities to the XPF-ERCC1 and XPG subunits of human DNA repair excision nuclease. J. Biol. Chem. 1996, 271, 11047–11050. [Google Scholar] [CrossRef] [PubMed]
  176. Bessho, T.; Sancar, A.; Thompson, L.H.; Thelen, M.P. Reconstitution of human excision nuclease with recombinant XPF-ERCC1 complex. J. Biol. Chem. 1997, 272, 3833–3837. [Google Scholar] [CrossRef]
  177. Fisher, L.A.; Bessho, M.; Wakasugi, M.; Matsunaga, T.; Bessho, T. Role of interaction of XPF with RPA in nucleotide excision repair. J. Mol. Biol. 2011, 413, 337–346. [Google Scholar] [CrossRef] [PubMed]
  178. Ellison, V.; Stillman, B. Biochemical characterization of DNA damage checkpoint complexes: Clamp loader and clamp complexes with specificity for 5’ recessed DNA. PLoS Biol. 2003, 1, E33. [Google Scholar] [CrossRef] [PubMed]
  179. Guler, G.D.; Liu, H.; Vaithiyalingam, S.; Arnett, D.R.; Kremmer, E.; Chazin, W.J.; Fanning, E. Human DNA helicase B (HDHB) binds to replication protein A and facilitates cellular recovery from replication stress. J. Biol. Chem. 2012, 287, 6469–6481. [Google Scholar] [CrossRef]
  180. Tkac, J.; Xu, G.; Adhikary, H.; Young, J.T.F.; Gallo, D.; Escribano-Diaz, C.; Krietsch, J.; Orthwein, A.; Munro, M.; Sol, W.; et al. HELB Is a Feedback Inhibitor of DNA End Resection. Mol. Cell 2016, 61, 405–418. [Google Scholar] [CrossRef]
  181. Sparks, J.L.; Kumar, R.; Singh, M.; Wold, M.S.; Pandita, T.K.; Burgers, P.M. Human exonuclease 5 is a novel sliding exonuclease required for genome stability. J. Biol. Chem. 2012, 287, 42773–42783. [Google Scholar] [CrossRef] [PubMed]
  182. Loo, Y.M.; Melendy, T. Recruitment of replication protein A by the papillomavirus E1 protein and modulation by single-stranded DNA. J. Virol. 2004, 78, 1605–1615. [Google Scholar] [CrossRef]
  183. Han, Y.; Loo, Y.M.; Militello, K.T.; Melendy, T. Interactions of the papovavirus DNA replication initiator proteins, bovine papillomavirus type 1 E1 and simian virus 40 large T antigen, with human replication protein A. J. Virol. 1999, 73, 4899–4907. [Google Scholar] [CrossRef]
  184. Weisshart, K.; Taneja, P.; Fanning, E. The replication protein A binding site in simian virus 40 (SV40) T antigen and its role in the initial steps of SV40 DNA replication. J. Virol. 1998, 72, 9771–9781. [Google Scholar] [CrossRef] [PubMed]
  185. Arunkumar, A.I.; Klimovich, V.; Jiang, X.; Ott, R.D.; Mizoue, L.; Fanning, E.; Chazin, W.J. Insights into hRPA32 C-terminal domain--mediated assembly of the simian virus 40 replisome. Nat. Struct. Mol. Biol. 2005, 12, 332–339. [Google Scholar] [CrossRef]
  186. Park, C.J.; Lee, J.H.; Choi, B.S. Solution structure of the DNA-binding domain of RPA from Saccharomyces cerevisiae and its interaction with single-stranded DNA and SV40 T antigen. Nucleic Acids Res. 2005, 33, 4172–4181. [Google Scholar] [CrossRef] [PubMed]
  187. Christensen, J.; Tattersall, P. Parvovirus initiator protein NS1 and RPA coordinate replication fork progression in a reconstituted DNA replication system. J. Virol. 2002, 76, 6518–6531. [Google Scholar] [CrossRef]
  188. Yuzhakov, A.; Kelman, Z.; Hurwitz, J.; O’Donnell, M. Multiple competition reactions for RPA order the assembly of the DNA polymerase delta holoenzyme. EMBO J. 1999, 18, 6189–6199. [Google Scholar] [CrossRef]
  189. Kim, H.S.; Brill, S.J. Rfc4 interacts with Rpa1 and is required for both DNA replication and DNA damage checkpoints in Saccharomyces cerevisiae. Mol. Cell Biol. 2001, 21, 3725–3737. [Google Scholar] [CrossRef]
  190. Otterlei, M.; Warbrick, E.; Nagelhus, T.A.; Haug, T.; Slupphaug, G.; Akbari, M.; Aas, P.A.; Steinsbekk, K.; Bakke, O.; Krokan, H.E. Post-replicative base excision repair in replication foci. EMBO J. 1999, 18, 3834–3844. [Google Scholar] [CrossRef]
  191. Loor, G.; Zhang, S.J.; Zhang, P.; Toomey, N.L.; Lee, M.Y. Identification of DNA replication and cell cycle proteins that interact with PCNA. Nucleic Acids Res. 1997, 25, 5041–5046. [Google Scholar] [CrossRef] [PubMed]
  192. Dianov, G.L.; Jensen, B.R.; Kenny, M.K.; Bohr, V.A. Replication protein A stimulates proliferating cell nuclear antigen-dependent repair of abasic sites in DNA by human cell extracts. Biochemistry 1999, 38, 11021–11025. [Google Scholar] [CrossRef] [PubMed]
  193. Farina, A.; Shin, J.H.; Kim, D.H.; Bermudez, V.P.; Kelman, Z.; Seo, Y.S.; Hurwitz, J. Studies with the human cohesin establishment factor, ChlR1. Association of ChlR1 with Ctf18-RFC and Fen1. J. Biol. Chem. 2008, 283, 20925–20936. [Google Scholar] [CrossRef] [PubMed]
  194. Jeong, Y.T.; Rossi, M.; Cermak, L.; Saraf, A.; Florens, L.; Washburn, M.P.; Sung, P.; Schildkraut, C.L.; Pagano, M. FBH1 promotes DNA double-strand breakage and apoptosis in response to DNA replication stress. J. Cell Biol. 2013, 200, 141–149. [Google Scholar] [CrossRef] [PubMed]
  195. Elia, A.E.; Wang, D.C.; Willis, N.A.; Boardman, A.P.; Hajdu, I.; Adeyemi, R.O.; Lowry, E.; Gygi, S.P.; Scully, R.; Elledge, S.J. RFWD3-Dependent Ubiquitination of RPA Regulates Repair at Stalled Replication Forks. Mol. Cell 2015, 60, 280–293. [Google Scholar] [CrossRef] [PubMed]
  196. MacKay, C.; Toth, R.; Rouse, J. Biochemical characterisation of the SWI/SNF family member HLTF. Biochem. Biophys. Res. Commun. 2009, 390, 187–191. [Google Scholar] [CrossRef] [PubMed]
  197. Sommers, J.A.; Banerjee, T.; Hinds, T.; Wan, B.; Wold, M.S.; Lei, M.; Brosh, R.M., Jr. Novel function of the Fanconi anemia group J or RECQ1 helicase to disrupt protein-DNA complexes in a replication protein A-stimulated manner. J. Biol. Chem. 2014, 289, 19928–19941. [Google Scholar] [CrossRef]
  198. Cui, S.; Arosio, D.; Doherty, K.M.; Brosh, R.M., Jr.; Falaschi, A.; Vindigni, A. Analysis of the unwinding activity of the dimeric RECQ1 helicase in the presence of human replication protein A. Nucleic Acids Res. 2004, 32, 2158–2170. [Google Scholar] [CrossRef]
  199. Cui, S.; Klima, R.; Ochem, A.; Arosio, D.; Falaschi, A.; Vindigni, A. Characterization of the DNA-unwinding activity of human RECQ1, a helicase specifically stimulated by human replication protein A. J. Biol. Chem. 2003, 278, 1424–1432. [Google Scholar] [CrossRef]
  200. Garcia, P.L.; Liu, Y.; Jiricny, J.; West, S.C.; Janscak, P. Human RECQ5beta, a protein with DNA helicase and strand-annealing activities in a single polypeptide. EMBO J. 2004, 23, 2882–2891. [Google Scholar] [CrossRef]
  201. Hu, Y.; Raynard, S.; Sehorn, M.G.; Lu, X.; Bussen, W.; Zheng, L.; Stark, J.M.; Barnes, E.L.; Chi, P.; Janscak, P.; et al. RECQL5/Recql5 helicase regulates homologous recombination and suppresses tumor formation via disruption of Rad51 presynaptic filaments. Genes. Dev. 2007, 21, 3073–3084. [Google Scholar] [CrossRef]
  202. Galanty, Y.; Belotserkovskaya, R.; Coates, J.; Jackson, S.P. RNF4, a SUMO-targeted ubiquitin E3 ligase, promotes DNA double-strand break repair. Genes. Dev. 2012, 26, 1179–1195. [Google Scholar] [CrossRef]
  203. Dornreiter, I.; Erdile, L.F.; Gilbert, I.U.; von Winkler, D.; Kelly, T.J.; Fanning, E. Interaction of DNA polymerase Alpha-primase with cellular replication protein A and SV40 T antigen. EMBO J. 1992, 11, 769–776. [Google Scholar] [CrossRef]
  204. Feng, J.; Wakeman, T.; Yong, S.; Wu, X.; Kornbluth, S.; Wang, X.F. Protein phosphatase 2A-dependent dephosphorylation of replication protein A is required for the repair of DNA breaks induced by replication stress. Mol. Cell Biol. 2009, 29, 5696–5709. [Google Scholar] [CrossRef]
  205. Majka, J.; Binz, S.K.; Wold, M.S.; Burgers, P.M. Replication protein A directs loading of the DNA damage checkpoint clamp to 5’-DNA junctions. J. Biol. Chem. 2006, 281, 27855–27861. [Google Scholar] [CrossRef]
  206. Davis, A.P.; Symington, L.S. The Rad52-Rad59 complex interacts with Rad51 and replication protein A. DNA Repair 2003, 2, 1127–1134. [Google Scholar] [CrossRef]
  207. Grimme, J.M.; Honda, M.; Wright, R.; Okuno, Y.; Rothenberg, E.; Mazin, A.V.; Ha, T.; Spies, M. Human Rad52 binds and wraps single-stranded DNA and mediates annealing via two hRad52-ssDNA complexes. Nucleic Acids Res. 2010, 38, 2917–2930. [Google Scholar] [CrossRef] [PubMed]
  208. Bansbach, C.E.; Betous, R.; Lovejoy, C.A.; Glick, G.G.; Cortez, D. The annealing helicase SMARCAL1 maintains genome integrity at stalled replication forks. Genes. Dev. 2009, 23, 2405–2414. [Google Scholar] [CrossRef] [PubMed]
  209. Yang, Z.G.; Liu, Y.; Mao, L.Y.; Zhang, J.T.; Zou, Y. Dimerization of human XPA and formation of XPA2-RPA protein complex. Biochemistry 2002, 41, 13012–13020. [Google Scholar] [CrossRef] [PubMed]
  210. Stigger, E.; Drissi, R.; Lee, S.H. Functional analysis of human replication protein A in nucleotide excision repair. J. Biol. Chem. 1998, 273, 9337–9343. [Google Scholar] [CrossRef] [PubMed]
  211. Daughdrill, G.W.; Buchko, G.W.; Botuyan, M.V.; Arrowsmith, C.; Wold, M.S.; Kennedy, M.A.; Lowry, D.F. Chemical shift changes provide evidence for overlapping single-stranded DNA- and XPA-binding sites on the 70 kDa subunit of human replication protein A. Nucleic Acids Res. 2003, 31, 4176–4183. [Google Scholar] [CrossRef] [PubMed]
  212. Aboussekhra, A.; Biggerstaff, M.; Shivji, M.K.; Vilpo, J.A.; Moncollin, V.; Podust, V.N.; Protic, M.; Hubscher, U.; Egly, J.M.; Wood, R.D. Mammalian DNA nucleotide excision repair reconstituted with purified protein components. Cell 1995, 80, 859–868. [Google Scholar] [CrossRef] [PubMed]
  213. Wong, J.M.; Ionescu, D.; Ingles, C.J. Interaction between BRCA2 and replication protein A is compromised by a cancer-predisposing mutation in BRCA2. Oncogene 2003, 22, 28–33. [Google Scholar] [CrossRef] [PubMed]
  214. Liu, J.; Doty, T.; Gibson, B.; Heyer, W.D. Human BRCA2 protein promotes RAD51 filament formation on RPA-covered single-stranded DNA. Nat. Struct. Mol. Biol. 2010, 17, 1260–1262. [Google Scholar] [CrossRef] [PubMed]
  215. Yates, L.A.; Tannous, E.A.; Morgan, R.M.; Burgers, P.M.; Zhang, X. A DNA damage-induced phosphorylation circuit enhances Mec1(ATR) Ddc2(ATRIP) recruitment to Replication Protein A. Proc. Natl. Acad. Sci. USA 2023, 120, e2300150120. [Google Scholar] [CrossRef]
  216. Lee, S.; Heo, J.; Park, C.J. Determinants of replication protein A subunit interactions revealed using a phosphomimetic peptide. J. Biol. Chem. 2020, 295, 18449–18458. [Google Scholar] [CrossRef]
  217. Stephan, H.; Concannon, C.; Kremmer, E.; Carty, M.P.; Nasheuer, H.P. Ionizing radiation-dependent and independent phosphorylation of the 32-kDa subunit of replication protein A during mitosis. Nucleic Acids Res. 2009, 37, 6028–6041. [Google Scholar] [CrossRef]
  218. Nuss, J.E.; Patrick, S.M.; Oakley, G.G.; Alter, G.M.; Robison, J.G.; Dixon, K.; Turchi, J.J. DNA damage induced hyperphosphorylation of replication protein A. 1. Identification of novel sites of phosphorylation in response to DNA damage. Biochemistry 2005, 44, 8428–8437. [Google Scholar] [CrossRef]
  219. Patrick, S.M.; Oakley, G.G.; Dixon, K.; Turchi, J.J. DNA damage induced hyperphosphorylation of replication protein A. 2. Characterization of DNA binding activity, protein interactions, and activity in DNA replication and repair. Biochemistry 2005, 44, 8438–8448. [Google Scholar] [CrossRef]
  220. Zernik-Kobak, M.; Vasunia, K.; Connelly, M.; Anderson, C.W.; Dixon, K. Sites of UV-induced phosphorylation of the p34 subunit of replication protein A from HeLa cells. J. Biol. Chem. 1997, 272, 23896–23904. [Google Scholar] [CrossRef]
  221. Gregoire, S.; Irwin, J.; Kwon, I. Techniques for Monitoring Protein Misfolding and Aggregation in Vitro and in Living Cells. Korean J. Chem. Eng. 2012, 29, 693–702. [Google Scholar] [CrossRef] [PubMed]
  222. Jones, M.L.; Aria, V.; Baris, Y.; Yeeles, J.T.P. How Pol Alpha-primase is targeted to replisomes to prime eukaryotic DNA replication. Mol. Cell 2023, 83, 2911–2924.e2916. [Google Scholar] [CrossRef] [PubMed]
  223. Zheng, F.; Georgescu, R.; Yao, N.Y.; Li, H.; O’Donnell, M.E. Cryo-EM structures reveal that RFC recognizes both the 3′- and 5′-DNA ends to load PCNA onto gaps for DNA repair. Elife 2022, 11, e77469. [Google Scholar] [CrossRef] [PubMed]
Figure 2. RPA isoforms from cells synced in the G2 phase either with DNA damage (blue) or without DNA damage (pink), separated with isoelectric focusing using a pH gradient from 5-6 and probed with phospho-specific antibodies. RPA heterotrimer probed with (A) anti-RPA70-CT, (B) phosphoRPA32(S4/8), (C) phosphoRPA32(T21), (D) phosphoRPA32(S23), and (E) phosphoRPA32(S33) antibodies. The red number above the corresponding peaks indicates the number of phosphorylations on each isoform. Figure adapted from [40].
Figure 2. RPA isoforms from cells synced in the G2 phase either with DNA damage (blue) or without DNA damage (pink), separated with isoelectric focusing using a pH gradient from 5-6 and probed with phospho-specific antibodies. RPA heterotrimer probed with (A) anti-RPA70-CT, (B) phosphoRPA32(S4/8), (C) phosphoRPA32(T21), (D) phosphoRPA32(S23), and (E) phosphoRPA32(S33) antibodies. The red number above the corresponding peaks indicates the number of phosphorylations on each isoform. Figure adapted from [40].
Genes 15 00167 g002
Figure 3. Summary diagram of RPA phosphorylation before and after DSBs. RPA is always phosphorylated (some sites are unknown and indicated with a ?), but after DDR is activated, RPA becomes hyperphosphorylated by the PIKK family of kinases and CDK cell cycle kinases, which phosphorylate RPA. Phosphorylation sites after DNA damage are shown in Figure 1D. Adapted from [40].
Figure 3. Summary diagram of RPA phosphorylation before and after DSBs. RPA is always phosphorylated (some sites are unknown and indicated with a ?), but after DDR is activated, RPA becomes hyperphosphorylated by the PIKK family of kinases and CDK cell cycle kinases, which phosphorylate RPA. Phosphorylation sites after DNA damage are shown in Figure 1D. Adapted from [40].
Genes 15 00167 g003
Table 1. Human RPA structures in the Protein Data Bank (PDB).
Table 1. Human RPA structures in the Protein Data Bank (PDB).
Subunits/DomainExperiment InformationPDB IDResolution (Å)Citation
X-ray; MRE11; RPA70(1-121)8K001.4[76]
RPA70-FX-ray; RAD9; RPA70(1-120)8JZY1.5[76]
X-ray; ETAA1; RPA70(1-120)8JZV1.5[76]
X-ray; ATRIP; RPA70(1-123)7XV41.6[76]
X-ray; HelB; RPA70(1-120)7XV11.8[76]
X-ray; BLMp1; RPA70(1-120)7XV01.5[76]
X-ray; RMI1; RPA70(1-120)7XUV1.6[76]
X-ray; WRN; RPA70(1-120)7XUT1.6[76]
X-ray; BLMp2; RPA70(1-120)7XUW1.8[76]
X-ray; drug; RPA70(1-123)5E7N1.21[77]
X-ray; Dna2 peptide; RPA70(1-123)5EAY1.55[70]
X-ray; PrimPol(514-525); RPA70(1-123)5N852[74]
X-ray; PrimPol(480-560); RPA70(1-123)5N8A1.28[74]
X-ray; drug; RPA70(1-123)4R4T1.28 [78]
X-ray; drug; RPA70(1-123)4R4Q1.35[78]
X-ray; drug; RPA70(1-123)4R4O1.33[78]
X-ray; drug; RPA70(1-123)4R4I1.4[78]
X-ray; drug; RPA70(1-123)4R4C1.4[78]
X-ray; drug; RPA70(1-123)4O0A1.2[73]
X-ray; drug; RPA70(1-123)4LWC1.61[73]
X-ray; drug; RPA70(1-123)4LW11.631[73]
X-ray; drug; RPA70(1-123)4LUZ1.9[73]
X-ray; drug; RPA70(1-123)4LUV1.4[73]
X-ray; drug; RPA70(1-123)4LUO1.54[73]
X-ray; drug; RPA70(1-123)4IPH1.94[79]
X-ray; drug; RPA70(1-123)4IPG1.58[79]
X-ray; drug; RPA70(1-123)4IPD1.51[79]
X-ray; drug; RPA70(1-123)4IPC1.22[79]
X-ray; drug; RPA70(1-123)4IJL1.7[80]
X-ray; drug; RPA70(1-123)4IJH1.498[80]
X-ray; p53N(33-60)/RPA70(1-123)2B3G1.6[72]
X-ray; RPA70(1-123)2B291.6[72]
NMR; RPA70(1-114)1EWI-[68]
RPA70-A/BX-ray; RPA70(181-422)1FGU2.5[63]
X-ray; ssDNA; RPA70(181-422)1JMC2.4[66]
NMR; XPA-MBD/RPA70(168-326)1D4U-[20]
RPA32-wHLHNMR; UNG2(73-88); RPA32(171-270)1DPU-[69]
X-ray; RPA32(197-270)4OU01.4[82]
RPA32/RPA14X-ray; full length2Z6K3[75]
X-ray; full length2PQA2.5[75]
X-ray; full length2PI22[75]
X-ray; RPA32(43-172)/RPA14(1-121)1QUQ2.5[81]
RPA70/RPA32/RPA14X-ray; Zn2+;
RPA70(435-616)/RPA32(43-171)/RPA14(1-121)
1L1O2.8[71]
Table 2. Human proteins that interact with RPA.
Table 2. Human proteins that interact with RPA.
Interacting ProteinInteraction Site on RPADNA Metabolism PathwayCitation
AIDRPA32Immunoglobulin diversification[107]
* AjubaRPA70DNA damage response (DDR)[108,109]
** ATR/ATRIPRPA70-FCheckpoint signaling, DNA repair[110,111,112]
BIDRPA70-FReplication stress response[113]
* BLMRPA70-A/BDNA unwinding, resection[114,115,125]
** BRCA2?Homologous Recombination (HR)[105]
DDX11?Chromosome segregation[193]
** DNA-PKcs?DNA repair[18,116]
** DSS1RPA70HR[121]
* ETAA1RPA70-F/RPA32ATR activation, repair at stalled
replication forks
[117,118,119,120]
* EXO5RPA70-FIntrastrand crosslink repair[181]
* FANCJRPA70DNA repair, genome stability[123]
* FBH1RPA32DNA unwinding, resection[193]
HELBRPA70-FReplication stress response[179,180]
HERC2RPA70Replication[124,125]
HIRARPA70-CChromatin remodeling[126]
Histones H3 and H4RPA70-FChromatin remodeling[127]
* HLTFRPA70Genome stability[195,196]
HSF1RPA70Gene expression[122]
* MeninRPA32Genome stability[129,130]
** MRE11-RAD50-NBS1RPA70-FHR[131,132]
NucleolinRPA14Replication (stress)[133,134,135]
NSMFRPA32DDR[101]
** p53RPA70-FHR[136,137,138,139,140]
* p53BP1RPA70/RPA32DDR[106]
* PALB2RPA32Recovery of stalled replication forks[141]
Papillomavirus E1RPA70-AReplication[182,183]
Parvovirus NS1RPA70/RPA32DNA unwinding, resection[187]
PCNARPA70Replication[191]
PolδRPA70Replication[188]
Pol-PrimRPA70-F/A/BReplication restart, DNA damage
tolerance
[102,203]
** PP2ARPA32DDR[204]
* PRP19/BCAS2RPA70-F/CDNA repair[144]
* PTENRPA70Genome stability[145]
* RAD9/RAD1/HUS1 (9-1-1)RPA70/RPA32DDR[146]
RAD17RPA70-FDDR, replication stress response[147,178,205]
** RAD51RPA70-ARecombination[103,150,151,206]
** RAD52RPA70-A/B & RPA32-wHLHDNA repair[21,69,152,153,154,155,156,207]
RECQL1RPA70DNA unwinding[198,199]
RECQ5β?DNA unwinding[200,201]
RFCRPA70DNA unwinding[188,189]
* RFWD3RPA32-wHLHDNA repair[157,158,195]
* RNaseH?Transcription, DNA repair[159]
* RNF4RPA70DNA DSB repair by HR[202]
SENP6RPA70Unperturbed DNA replication[38]
SMARCAL1/HARPRPA32-wHLHReplication fork restart[160,161,162,208]
SV40 Large T antigenRPA70-A/B & RPA32-wHLHReplication[182,183,184,185,186]
* TipinRPA32-wHLHDDR[164,165]
* UDGRPA32-wHLHBase excision repair[69,166,190]
* UNG2RPA32-wHLHBase excision repair[69,166]
* WRNRPA70-A/BDNA unwinding, resection[115,170]
* XPARPA70-A & RPA32-wHLHNucleotide excision repair (NER)[69,166,171,209,210,211]
* XPF-ERCC1RPA70NER[174,175,176]
* XPG?NER[174,176,212]
* Involved in DNA repair; ** Involved in DNA DSB repair.
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

Fousek-Schuller, V.J.; Borgstahl, G.E.O. The Intriguing Mystery of RPA Phosphorylation in DNA Double-Strand Break Repair. Genes 2024, 15, 167. https://doi.org/10.3390/genes15020167

AMA Style

Fousek-Schuller VJ, Borgstahl GEO. The Intriguing Mystery of RPA Phosphorylation in DNA Double-Strand Break Repair. Genes. 2024; 15(2):167. https://doi.org/10.3390/genes15020167

Chicago/Turabian Style

Fousek-Schuller, Valerie J., and Gloria E. O. Borgstahl. 2024. "The Intriguing Mystery of RPA Phosphorylation in DNA Double-Strand Break Repair" Genes 15, no. 2: 167. https://doi.org/10.3390/genes15020167

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