Next Article in Journal
A New, Convenient Way to Fully Substituted α,β-Unsaturated γ-Hydroxy Butyrolactams
Next Article in Special Issue
Increased Drop in Activity of Alkaline Phosphatase in Plasma from Patients with Endocarditis
Previous Article in Journal
Light-Dependent Expression and Promoter Methylation of the Genes Encoding Succinate Dehydrogenase, Fumarase, and NAD-Malate Dehydrogenase in Maize (Zea mays L.) Leaves
Previous Article in Special Issue
Can Allostery Be a Key Strategy for Targeting PTP1B in Drug Discovery? A Lesson from Trodusquemine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Maintaining Genome Integrity: Protein Kinases and Phosphatases Orchestrate the Balancing Act of DNA Double-Strand Breaks Repair in Cancer

1
Department of Pathology, University of Chicago, Chicago, IL 60637, USA
2
Department of Integrated Biomedical Science, Soonchunhyang Institute of Medi-Bio Science (SIMS), Soonchunhyang University, Cheonan 31151, Chungcheongnam-do, Republic of Korea
3
Analytical Instrumentation Center, Hunan University, Changsha 410082, China
4
College of Biology, Hunan University, Changsha 410082, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(12), 10212; https://doi.org/10.3390/ijms241210212
Submission received: 27 April 2023 / Revised: 13 June 2023 / Accepted: 14 June 2023 / Published: 16 June 2023
(This article belongs to the Special Issue The Role of Phosphatases in Human Health and Disease)

Abstract

:
DNA double-strand breaks (DSBs) are the most lethal DNA damages which lead to severe genome instability. Phosphorylation is one of the most important protein post-translation modifications involved in DSBs repair regulation. Kinases and phosphatases play coordinating roles in DSB repair by phosphorylating and dephosphorylating various proteins. Recent research has shed light on the importance of maintaining a balance between kinase and phosphatase activities in DSB repair. The interplay between kinases and phosphatases plays an important role in regulating DNA-repair processes, and alterations in their activity can lead to genomic instability and disease. Therefore, study on the function of kinases and phosphatases in DSBs repair is essential for understanding their roles in cancer development and therapeutics. In this review, we summarize the current knowledge of kinases and phosphatases in DSBs repair regulation and highlight the advancements in the development of cancer therapies targeting kinases or phosphatases in DSBs repair pathways. In conclusion, understanding the balance of kinase and phosphatase activities in DSBs repair provides opportunities for the development of novel cancer therapeutics.

1. Introduction

DNA exposed to radiation or chemical agents may cause damages. The resulting damages lead to a higher occurrence of mutations and genome instability, eventually leading to the development of various diseases, including cancer. However, mammalian cells have evolved several mechanisms to detect and repair those DNA damages. The DNA-repair signaling is activated immediately when DNA damages are detected, resulting in the activation of cell-cycle checkpoint and recruitment of DNA-repair proteins to damage sites [1].
Phosphorylation is an important protein post-translational modification that regulates DNA-repair signaling [2]. Enzymes can be divided into two categories, kinases and phosphatases, both of which are involved in maintaining the regulation of protein phosphorylation during DNA damage. These enzymes are responsible for the reversible phosphorylation of proteins. Kinases add a phosphate group to specific amino acids on proteins, while phosphatases remove the phosphate group. The interplay between kinases and phosphatases enables cells to rapidly and precisely control a broad range of cellular processes, including DNA repair.
Numerous kinases and phosphatases have been proved to have function on DNA repair. These enzymes are responsible for modulating DNA damage response (DDR), controlling the cell cycle, and acting in a coordinated manner to regulate DNA-repair processes [3]. Briefly, most kinases play roles in initiating and modulating DNA repair under different types of DNA damage, including DSBs, SSBs (singledouble-strand breaks), stalled replication forks, or UV-induced base damages. Those kinases function by phosphorylating a broad range of substrates involved in the DNA-repair pathway. ATM (Ataxia-Telangiectasia Mutated), ATR (ATM and RAD3-related), and DNA-PK (DNA-dependent protein kinase), the three most important kinases in DNA repair, work together or respectively to activate downstream kinases, such as CHK1 (checkpoint kinases 1) and CHK2 (checkpoint kinases 1), which, in turn, phosphorylate other downstream targets [4]. Phosphatases function in opposition to kinases by removing phosphate groups from DNA-repair-related proteins. These enzymes play an essential role in terminating kinase signaling, and are required for resetting the DNA-repair machinery. For example, PP1 (protein phosphatase 1) or WIP1 phosphatases terminate DNA damage response by dephosphorylating various ATM and Chk2 substrates [5].
The balance between kinase and phosphatase activities is important for proper DDR signaling transduction. Alterations in kinase or phosphatase activity lead to genome instability. For example, ATM mutation leads to reduced DNA-repair efficacy and accumulation of double-strand breaks in cells, resulting in higher mutation frequency and genome instability. Inactivating mutations in PP1, Wip1, or other phosphatases can lead to hyperactivation of DDR signaling and a predisposition to cancer [5].
In conclusion, kinases and phosphatases function coordinately in DNA repair and maintaining genome stability. Understanding the functions of kinases and phosphatases in DNA repair will enable the development of targeted therapies to improve the therapeutics for diseases such as cancer.

2. Kinases in the DNA DSBs Repair

In mammalian cells, protein phosphorylation kinases are crucial components in the DSB signaling cascade, from sensing DNA damage, to transducing signals, to repairing DSBs. The protein kinases are traditionally categorized as either apical or effector kinases [4,6]. Most of the protein kinases involved in DNA DSB repair are evolutionarily conserv ed, from yeast to mammalian systems. To date, facilitated by mass spectrometry (MS)-based proteomics, dozens of DNA-repair proteins have been identified to be phosphorylated by these kinases and a more systematic network of phosphorylation events triggered by kinases has been established [6,7,8]. Therefore, we have gathered a more comprehensive vision of the machinery of the DSBs repair.

2.1. Apical (Sensor) Kinases

The apical kinases in DSB repair, including DNA-PKcs (DNA-dependent protein kinase, catalytic subunit), ATM, and ATR, belong to the superfamily of phosphatidylinositol-3-kinase-related kinases (PIKKs) [4,9]. DNA-PKcs, ATM, and ATR are serine/threonine protein kinases, and, as PIKKs members, they share some structural similarity in the catalytic loop within the kinase domain [9]. DNA-PKcs, ATM, and ATR can be activated by DSBs and, once activated by DSBs, they lie at the heart of the signal-transduction process as they can phosphorylate a broadboard and overlapping spectrum of substrates transducing DNA damage signals (Figure 1) [9,10].
There are two pathways that dominate the repair of DSBs: non-homologous end-joining (NHEJ) and homologous recombination (HR) [11,12,13,14]. NHEJ repairs most DSBs in mammalian cells, while HR is preferred when NHEJ cannot be completed [15,16]. DNA-PK is a critical kinase in the NHEJ pathway [4]. ATM is essential for the initiation and completion of the DSB repair by HR pathway [17,18,19]. ATR is activated in response to persistent single-stranded DNA, which could be an intermediate in DSB repair pathways such as HR [20,21].
In response to DSBs and being tightly regulated by their co-factors (Ku80 for DNA-PKcs, NBS1 for ATM, and ATRIP for ATR) [22,23,24], the apical kinases can be activated by auto-phosphorylation [4]. Once activated, they can phosphorylate H2A histone family member X (H2AX). The pS139 of H2AX (γH2AX), a marker for DSBs, generates binding sites for mediators/adaptors proteins such as DNA damage checkpoint protein 1 (MDC1), breast cancer susceptibility protein 1 (BRCA1) and p53 binding protein 1 (53BP1) [10,25,26]. The apical kinases then transfer the phosphorylation to downstream effector kinases via mediators in the DSB repair pathways [4].
Specifically, DNA-PKcs can be auto-phosphorylated or occasionally phosphorylated by ATM and ATR [27,28]. S2056 and T2609 are two major phosphorylation sites of DNA-PKcs, and both are crucial for its activity in DNA repair [8]. Once activated, DNA-PKcs dissociates from the DSBs and phosphorylates a number of substrates, including Replication protein A (RPA), Werner syndrome ATP-dependent helicase (WRN), Artemis, and H2AX [8,29,30].
Upon DNA damage, ATM auto-phosphorylates on S1981, then further auto-phosphorylates on S367, S1893, S1981, and, potentially, other sites to activate the ATM kinase [31,32]. Proteins involved in the ATM signaling, such as CHK2, p53, BRCA1, NBS1, and H2AX, are all targets of activated ATM [33].
Unlike DNA-PKcs and ATM, ATR is activated by DSBs and also a broad spectrum of other DNA damages [20]. Autophosphorylation T1989 is critical for ATR activation [34]. Many of the proteins in ATR signaling network, such as RPA, TopBP1 (DNA topoisomerase II binding protein 1), Claspin, and CHK1, are substrates of ATR [20] (Figure 2).

2.2. Effector Kinases

In mammalian cells, activated apical kinases transfer stimulatory phosphorylation to the downstream checkpoint kinases CHK1 and CHK2. As shown in Figure 2, in response to DSBs, CHK2 is activated via phosphorylation on T68 by ATM and, subsequently, oligomerization and autophosphorylation at T383 and T387 [35,36,37], while, in response to single-stranded DNA which can be generated from DSBs, CHK1 is phosphorylated on S317 and S345 by ATR kinase to become activated [38]. Both CHK1 and CHK2 are serine/threonine-specific protein kinases. Activated CHK1 and CHK2 phosphorylate a variety of downstream substrates, such as p53 and CDC25A (cell division cycle 25A), to activate DNA-repair pathways, bring out cell-cycle arrest, or induce apoptosis when DNA damage is severe [39,40].

2.3. Additional Kinases

There are other kinases involved in the DSB repair downstream signaling, such as cyclin-dependent kinases (CDKs), WEE1, polo-like kinase 1 (PLK1), and casein kinase II (CK2); see Figure. 2. They can be phosphorylated by apical and effector kinases. They also belong to the serine/threonine family of protein kinases and play important roles in repairing DNA and regulating a range of other processes, such as the cell-cycle arrest and initiation of apoptosis [6].
In addition to the serine/threonine kinases mentioned above, a number of tyrosine kinases have been demonstrated to participate in the DSBs repair signaling pathways, as previously described [41]. In brief, tyrosine kinases, such as epidermal growth factor receptor (EGFR), insulin-like growth factor 1 receptor (IGF-1R), ACK1/ TNK2, and Src family of tyrosine kinases, can be activated by DNA damage and epigenetically regulate DNA damage signaling by phosphorylating the core histones and chromatin modifiers [41].

2.4. Kinase Mutations and Related Diseases

As DNA damage response can be activated at early stages of tumorigenesis, the protein kinases involved in the DNA-damage-signaling pathways may serve as a barrier to tumorigenesis and development [42,43]. Concordant with this, the key kinase genes in DSB repair, ATM and ATR, are frequently mutated in various human cancers [44]. The XRCC7 or PRKDC gene (which encodes DNA-PK) has also been reported to present a higher mutation rate in certain types of cancers, such as colorectal, gastric, and endometrial cancers [45]. The CHK2 gene has also been found to be associated with some types of cancers, such as breast, kidney, and thyroid cancers [46]. The CHK1 gene, however, is not regarded as a canonical tumor suppressor based on its conflicting evidence in cancer [47].
Actually, most of the disease related to aberrations in DSBs repair protein kinases are marked by growth and developmental problems, such as cancer (abnormal cell proliferation, see details above), Ataxia telangiectasia (with immune and neuron dysfunctions, and radiosensitivity and cancer predisposition features; associated with ATM, PRKDC, CHK1, or CHK2 mutations) [48,49] and Seckel syndrome (growth retardation; associated with ATR mutations) [50].
In addition, the kinase-regulated DNA repair is crucial for the generation of diverse foreign antigen. The failure to process DNA repair can lead to immunodeficiency disease [51], such as Ataxia telangiectasia (mentioned above) and severe combined immunodeficiency (associated with PRKDC mutations) [52].
Regarding the mutation hotspots and mutation types in cancer, most of the truncating mutations of ATM, ATR, CHEK1, and CHEK2 are considered to be, at least, likely pathogenic. While missense mutations in ATR and CHK1 are not common, ATM missense mutations have been reported across the whole ATM gene. Moreover, pathogenic/likely pathogenic missense mutations of ATM and CHK2 are mainly located within their kinase domains [53,54,55].

3. Phosphatases in DSBs Repair

Though the activation of DSBs-repair signaling is driven by the phosphorylation cascades of DNA-repair proteins, the steady state of phosphorylation is regulated by both kinases and phosphatases. Therefore, reversible phosphorylation of DNA-repair proteins is an important regulatory mechanism in DSBs repair [53,54,55].
Unlike kinases that share a common catalytic fold and mechanism, phosphatases present greater diversity of structures and catalytic mechanism [56]. Thus, based on the differences in sequence, structure, and catalytic mechanisms, protein phosphatases that are involved in DSBs repair can be divided into different groups: the classic serine/threonine phosphoprotein phosphatase (PPP) family (comprising PP1, PP2A, PP2B, PP4, PP5, PP6, and PP7), the serine/threonine protein phosphatase family that is Mg2+ or Mn2+-dependent (PPM) (including PP2C), and the dual-specificity phosphatase (DUSP) family (such as cyclin-dependent kinase (CDK)-antagonizing cell-division cycle 14 (CDC14)) [5,57].

3.1. PPP Family

Among the PPP family members, PP1, PP2A, PP2B, PP4, and PP6 are all considered as split enzymes, as their selective dephosphorylation function requires the assembly of two modules, the catalytic subunit and the regulatory (non-catalytic) subunit [56].
PP1 (protein phosphatase 1) is an abundant protein in cells and is responsible for dephosphorylation of about one third of all phosphor-proteins [5,56]. The PP1 phosphatase has been demonstrated as a main regulator in DNA repair. Repo-Man/PP1γ holoenzyme (with Repo-Man as the regulatory component) can dephosphorylate ATM at S1981 and reduce ATM kinase activity [58]. Together with PP1-binding factor PNUTS (phosphatase 1 nuclear targeting subunit), Repo-Man/PP1 can also regulate the phosphorylation status of CHK1, H2AX, 53BP1, RPA, and RAD51 [59,60].
The conserved region of the catalytic subunit of PP2A is similar to that of PP1, which may explain the overlapping targets (including ATM, CHK1, γ-H2AX, and RPA) of these two phosphatases in DNA repair [57]. The distinguished wide range of regulatory subunit of PP2A also provide the holoenzyme with the ability to recognize and mediate other proteins in DNA repair, such as DNA-PK and CHK2 [57,61].
PP2B is also called PP3 or calcineurin. Little is known of its involvement in the direct dephosphorylation of DNA-repair proteins. PP4 is 41% identical to PP1 and 65% identical to PP2A in human. PP4 is a ubiquitously expressed phosphatase and it regulates multiple cellular processes, including DNA repair independently from PP1 or PP2A. In humans, PP4 is capable of dephosphorylating γ-H2AX and RPA [57,62].
Different from other PPP family members, PP5 is a single subunit enzyme [63]. In mammalian cells, PP5 has been shown to interact with ATM and ATR in response to DNA damage. PP5′s interaction with ATM/ATR is necessary for ATM/ATR activation, and DNA-PK can be dephosphorylated by PP5 at T2609 to modulate the DNA-PK signaling [64]. Besides, there is also a reciprocal regulatory interplay between p53 and PP5, as PP5 can directly dephosphorylate p53 while p53 suppresses PP5′s transcription [65].
Though closely related to PP2A and PP4, PP6 has not been well studied. Some evidence showed that PP6 is recruited to DSBs sites by DNA-PK and PP6 may dephosphorylate γ-H2AX [66,67]. PP7 can target various phosphoproteins, such as myelin basic protein. Its function in DNA repair is not clearly known.

3.2. PPM Family

In contrast to the PPP family, PPM phosphatases are mostly monomeric enzymes [65]. PPM family members all share a highly similar catalytic core and they are resistant to classic inhibitors and toxins of the PPP family [54]. PP2C belongs to the PPM family and among the PP2C family members, WIP1/PPM1D (wild-type p53-inducible phosphatase 1) has been shown to actively participate in DNA repair. In human cells, WIP1 dephosphorylates CHK1, CHK2, and p53 to downregulate DNA-damage checkpoint signaling and enhance cell-cycle recovery [68,69,70]. It can dephosphorylate γ-H2AX to inhibit the activation of the DNA-damage checkpoint. Moreover, Wip1 can dephosphorylate Atm at S1981 to modulate Atm activation in mice [71,72].
Another PP2C phosphatase, PPM1G (also termed PP2Cγ), also has a role in the DNA repair. It can mediate the exchange of H2A-H2B, and PPM1G-deficient cells are sensitive to DNA damage. It is a target of ATM and dephosphorylates USP7 (ubiquitin specific peptidase 7) to regulate p53 [5,73].

3.3. DUSP Family

Within the DUSPs, CDC14, which is capable of dephosphorylating both phosphoserine/phosphothreonine and phosphotyrosine residues, has been involved in DNA repair. Several knockout experiments conducted in mammalian cells have implied the direct role of CDC14A and CDC14B in DSBs repair [57]. CDC14 tends to dephosphorylate substrates that are previously phosphorylated by CDK, suggesting its role counteracting CDK substrates in the DNA repair process [57,62].

3.4. Phosphatase Alterations and Related Diseases

Other than involving in the DNA repair, protein phosphatases have a broad substrate specificity and diverse cellular functions, including cell division, apoptosis, protein synthesis, glycogen metabolism, neuronal activities, etc. Therefore, aberrations of phosphatases are related to a variety of diseases, such as cancers, viral infection, and heart diseases [74,75,76,77].
Compared to kinases, DNA-damage-repair-related phosphatase genes are less frequently mutated, their differential expression and interaction alterations between the substrates and phosphatases can also result in cellular dysfunction and disorders [78,79].

4. Dynamic Balance between Kinases and Phosphatases in DSBs Repair and Cancer Therapeutics

Upon DNA DSBs, ATM, ATR, and DNA-PK initiate signaling transduction. These kinases then activate CHK1 and CHK2 kinases and other substrates to result in DNA repair or apoptosis [3]. DNA-repair signaling is relayed by protein phosphorylation cascades, and the protein phosphatases can negatively regulate these phosphorylation events [80]. Though most of the phosphatase negatively regulates the proteins in the DNA DSBs repair, certain phosphatases could enhance the activity of proteins in the pathway [80]. Table 1 is a summary of kinases, phosphatases, and their substrates in DSBs repair. The dynamic balance between kinases and phosphatases fine-tunes the phosphorylation cascades in the DNA-repair process, including regulating the DNA-damage checkpoint and the repair of DNA lesion [57].
Targeted therapy is a more precise and effective approach to treating cancer by delivering therapeutic agents specifically to cancer cells while minimizing harm to healthy ones. Researchers have developed inhibitors that target specific proteins involved in DSBs repair and cell-cycle regulation, such as ATM, ATR, DNA-PK, CHK1, CHK2, PP1, PP2A, and PP1D. By inhibiting these key kinases and phosphatases, the effectiveness of cancer therapy can be enhanced, rendering cancer cells more vulnerable to treatment-induced DNA damage [323,324]. Table 2 provides a detailed overview of specific inhibitors and their targets, further highlighting their potential in enhancing cancer therapy.

4.1. ATM Inhibitors

AZD0156 is an orally bioavailable ATM inhibitor that has been identified to interact with catalytic lysine (Lys2717), kinase hinge (Cys2770), and back pocket (Tyr2755) to bind with the ATP site of ATM. With a potent inhibition effect, AZD0156 can act as a radiosensitizer, inhibiting tumor growth, and promoting a killing effect on cancer cells. It can also be used in combination with some drugs, such as olaparib, fluorouracil, and irinotecan [325,326,327]. Previous research showed that AZD0156 could prevent the repair mechanism induced by olaparib-induced DNA damage, resulting in apoptotic cellular responses. However, as a single agent, AZD0156 has shown limited effects on colorectal cancer cells. The combination between AZD0156 and irinotecan can improve dsDNA breaks, leading to cell-cycle arrest [328].
Another ATM inhibitor, M3541, is an ATP-selective inhibitor that can impede DSB-repair mechanisms. With an IC50 less than 1 nM, it can increase the sensitivity of cancer cells and improve anti-tumor effects of ionizing radiation. The administration of M3541 orally in nude mice has also shown a decrease in tumor size [325,329,330]. The combination of an ATM inhibitor with a CHK2 inhibitor (NCT03571438) has been under investigation as a potential therapy leading to apoptosis [331]. On the other hand, there are some ATM inhibitors with IC50 higher than that of AZD0156 and M3541, such as KU-59403 (IC50 = 3 nM), KU60019 (IC50 = 6.3 nM), KU-55933 (IC50 = 13 nM), and AZ31 (IC50 = 46 nM) [327].

4.2. ATR Inhibitors

Berzosertib, also known as VE-822, M6620, or VX-970, is a selective ATR inhibitor with an IC50 of 19 nM and has entered in clinical trial NCT04052555. It is an intravenously administered upgraded version of VE-822. Berzosertib can prevent the phosphorylation of γ-H2AX, which blocks DNA repair process and triggers DNA damage accumulation. Additionally, the combination of berzosertib with gemcitabine (NCT02157792) was found to be well-tolerated by patients. A previous study conducted on 48 patients showed that four patients had a partial response (PR), and 29 patients experienced no change. [332,333].
Ceralasertib, or AZD6738, is a potent and selective oral ATR inhibitor. The mechanism of ceralasertib is to impede the CHK1 phosphorylation and promote the γ-H2AX phosphorylation during stress response in the S-phase of the cell cycle [334]. Ceralasertib can induce an ATM-dependent signaling pathway and inhibit the HR-repair pathway. A previous study showed that anti-tumor activity and a decrease in tumor size was observed in patients treated with ceralasertib in combination with carboplatin (NCT02264678) [335]. Another study also described a combination of ceralasertib with durvalumab (NCT03780608) in gastric-cancer patients. This kind of combination can result in increased anti-tumor activity and loss of essential genes in DDR [336].

4.3. DNA-PK Inhibitor

Peposertib, or M3841, has been known as a DNA-PK inhibitor both in monotherapy and in combination with another agent. Peposertib is an orally selective inhibitor at sub-nanomolar concentrations. This inhibitor can impede DNA-PK and then trigger p53-dependent anti-tumor activity [337]. The selectivity of peposertib has been examined among several protein members of the phosphoinositide-3 kinase (PI3K) and PI3K-related family, such as ATR, ATM, and DNA-PK. As a result, the inhibition of peposertib on DNA-PK is significantly higher than that on other proteins. A previous clinical trial (NCT02316197) demonstrated that peposertib could be tolerated by patients and has displayed acceptable efficacy with the recommended phase 2 dose (RP2D) of 300 mg twice daily (BID) [338].
AZD7648 also has been discovered as a potent and selective DNA-PK inhibitor. AZD7648 is an orally administered inhibitor that can bind with the catalytic subunit of DNA-PK. Based on the previous study (NCT03907969), adding AZD7648 can improve the effectiveness of other anti-cancer agents, such as radiation and olaparib (PARP inhibitor). Compared to monotherapy treatment with AZD7648, radiation, or olaparib, combination treatment with AZD7648 can improve the inhibition of cancer growth, as well as reduce macronuclei formation and chromosomal aberration. [339,340].

4.4. CHEK1 and CHEK2 Inhibitors

LY2606368, known as prexasertib, is studied as a selective and ATP-competitive CHK1 inhibitor that can be used as a monotherapy or in combination with other inhibitors. LY2606368 is a second generation of CHK1 inhibitor with IC50<1 nM, while the first generation, such as LY2603618, has higher IC50 (about 8 nM). Not only does prexasertib inhibit CHK1, but it also inhibits CHK2, although the inhibitory effect is lower for CHK2. A previous study (NCT02808650) has demonstrated that prexasertib can induce apoptosis as a single agent [341,342,343]. Prexasertib can also be combined with PD-L1 antibody LY3300054 (NCT03495323) to activate the cytotoxic T-cell in Cyclin E1 (CCNE1)-amplified high-grade serous ovarian-cancer (HGSOC) patients [344].
MK-8776 is also known as a specific CHK1 inhibitor. It can act as a single agent or in combination with other drugs, such as gemcitabine and cytarabine (NCT00779584). MK-8776 was previously known as SCH 900776, a pyrazolo[1,5-a]pyrimidine derivative that binds with the hinge region located in the kinase ATP-binding site [341,345]. Previous studies show that MK-8776 treatment can enhance cancer-cell sensitivity to hydroxyurea, gemcitabine, and radiation, leading to DNA damage accumulation [346,347].
On the other hand, CHK2 inhibitors are less effective in cancer treatment. PHI-101 (NCT04678102) was discovered as an orally administered and highly selective CHK2 inhibitor with anti-tumor activity in ovarian cancer. It impedes the DSBs repair pathway in cancer, especially in platinum-resistant recurrent ovarian cancer [348]. CCT241533, containing a 2-(quinazolin-2-yl) phenol scaffold, is another highly selective inhibitor of CHK2, and is also a moderate selective inhibitor of CHK1. CCT241533 can be used as a single agent or combined with chemotherapy drugs, such as topotecan, camptothecin, and etoposide for cancer therapy [349].

4.5. PP1 Inhibitors

Different types of PP1 inhibitors have been identified, targeting either the catalytic site or the regulatory subunits of PP1. Natural product PP1 inhibitors include okadaic acid, microcystin-LR, cantharidin, and tautomycin, which have been isolated from marine organisms, cyanobacteria, and higher plants (1–2). These compounds bind to the catalytic subunit of PP1 and prevent it from dephosphorylating its substrates. On the other hand, synthetic PP1 inhibitors have been designed and synthesized based on the structure of natural product PP1 inhibitors. These inhibitors can be further divided into several structural classes, such as indole derivatives, benzimidazole derivatives, and pyridine derivatives.
PP1 inhibitors have shown promise in preclinical studies as potential therapeutics for cancers. For instance, fostriecin has demonstrated antitumor activity in animal models of human cancer (3). However, the development of PP1 inhibitors as clinical therapeutics is still in its early stages, and more research is needed to optimize their efficacy and safety.

4.6. PP2A Inhibitors

LB100, also known as LB1, has been recorded as a norcantharidin derivative compound, while norcantharidin was known as a traditional PP2A inhibitor. LB100 can bind and competitively inhibit the PP2A protein, which can sensitize cells to chemotherapy and radiotherapy. A phase I clinical trial of LB100 (NCT01837667) reported that LB100 has anti-tumor activity alone and is safe and tolerated by adult patients. In addition, a phase II clinical trial of LB100 (NCT03027388) showed an anti-glioblastoma (GBM) effect from LB100 in combination with PRMT5 inhibition. LB100 has also been used in combination therapy with doxorubicin or temozolomide as a DNA-damage agent, leading to the inhibition of cancer-cell growth. [350,351,352,353].

4.7. PPM1D Inhibitors

GSK2830371 is a potent and selective allosteric WIP1/PPM1D inhibitor used in cancer therapy. Its binding with the flap subdomain of the protein can confer significant effects on cell death and sensitize cells to chemotherapy. GSK2830371 can also reduce the doxorubicin dose, alleviating side-effects. Furthermore, GSK2830371 can stimulate the inhibition of cell proliferation and lead to cell-cycle arrest. [354,355,356,357].
Table 2. Cancer therapeutic inhibitors and their targets, analyzed based on clinicaltrials.gov database (accessed on 21 April 2023).
Table 2. Cancer therapeutic inhibitors and their targets, analyzed based on clinicaltrials.gov database (accessed on 21 April 2023).
TargetNational
Clinical Trials Number
Clinical PhaseDiseaseTreatmentTitleResult
ATM NCT02588105Phase 1Advanced
solid tumors
Drug: AZD0156
Drug: Olaparib
Drug: Irinotecan
Drug: Fluorouracil
Drug: Folinic acid
Study to assess
the safety and
preliminary efficacy of AZD0156 at
increasing doses alone or in
combination with other anti-cancer treatment in
patients with
advanced cancer (AToM)
Increased
radiosensitivity [325], increased anti-tumor
activity of olapararib
[326], enhanced DSBs [328]
NCT03225105Phase 1Solid tumorsDrug:
M3541
Radiation:
Palliative
radiotherapy (RT)
M3541 in
combination with radiotherapy in
subjects with
solid tumors
Decreased the tumor size, radiosensitivity [329,330], apoptosis [331]
ATRNCT04052555Phase 1
Bilateral breast
carcinoma
HER2-negative breast carcinoma
Localized breast
carcinoma
Recurrent breast
carcinoma
Triple-negative breast carcinoma
Drug:
Berzosertib
Procedure:
biospecimen
collection
Other: quality of life assessment
Other:
questionnaire
administration
Radiation:
radiation therapy
Testing the addition of an anti-cancer drug, berzosertib, to the usual treatment (radiation therapy) for chemotherapy-resistant triple-negative and estrogen and/or progesterone receptor positive, HER2 negative breast cancerBlocked DNA repair
process, accumulated DNA damage [333]
NCT02157792Phase 1Advanced
solid tumor
Drug:
M6620
M6620 first in
human study
RP2D is berzosertib 210 mg/m2 for days 2 and 9 + gemcitabine 1000 mg/m2 for days 1 and 8 within 3 weeks [332]
NCT02264678Phase 1Advanced solid
Malig—H&N SCC, ATM Pro/Def NSCLC, gastric, breast and ovarian cancer
Drug:
Gemcitabine
Ascending doses of ceralasertib in
combination with chemotherapy and/or novel
anti-cancer agents
RP2D for combination with carboplatin
: ceralasertib 40 mg
for days 1 and 2
+ carboplatin every
3 weeks [335]
NCT03780608Phase 2Gastric
adenocarcinoma
Drug:
Cisplatin
This study is a phase II study
of AZD6738 in
combination with durvalumab in
patients with solid tumor (cohort A (N = 30): GC who have failed secondary chemotherapy treatments regimen; cohort B (B = 30): Melanoma patients who have failed
immunotherapy (IO))
Increased anti-tumor
activity [336]
DNA-PKNCT02316197Phase 1Advanced
solid tumors
Chronic
lymphocytic
leukemia
Drug: MSC2490484A (M3814)Clinical phase I study investigating MSC2490484A, an inhibitor of a DNA-dependent protein kinase, in advanced solid tumors or chronic lymphocytic leukemiaRP2D is 400 mg
of peposertib [338]
NCT03907969Phase 1
Phase 2
Advanced
malignancies
Drug: AZD7648
Drug: PLD
A clinical trial to evaluate AZD7648 alone and in
combination with other anti-cancer agents in patients with advanced
cancers
Inhibited cancer growth, increased
macronuclei formation, and chromosomal
aberration [339]
CHK1NCT03495323Phase 1CancerDrug: LY3300054
Drug: Prexasertib
A study of
prexasertib (LY2606368),
CHK1 inhibitor,
and LY3300054,
PD-L1 inhibitor, in patients with
advanced solid
tumors
Activated the T-cells in blood samples of CCNE1-amplified HGSOC patients [344]
NCT02808650Phase 1Childhood
solid neoplasm
Recurrent malignant solid neoplasm
Recurrent primary central nervous
system neoplasm
Refractory
malignant solid
neoplasm
Refractory primary central nervous
system neoplasm
Other: Laboratory biomarker analysis
Other:
Pharmacological study
Drug: Prexasertib
Prexasertib in
treating pediatric patients with
recurrent or r
efractory solid tumors
Apoptosis, 150 mg/m2 administered IV for day 1 and day 15 of 28 days [341,342]
NCT00779584Phase 1Hodgkin’s disease
Lymphoma,
Non-Hodgkin’s
neoplasms
Drug: MK-8776
Drug: Gemcitabine
A dose-escalation study of MK-8776 (SCH 900776) with and without
gemcitabine in
participants with solid tumors or lymphoma (MK-8776-002/P05248)
RP2D is 200 mg/m2for MK-8776 with gemcitabine (1000 mg/m2) for day 1 and 8 of 21 days [347]
CHK2NCT04678102Phase 1Platinum-resistant ovarian cancer
Platinum-refractory ovarian carcinoma
Platinum-resistant fallopian tube
carcinoma
Platinum-resistant primary peritoneal carcinoma
Drug: PHI-101
administration
CHK2 inhibitor for Recurrent EpitheliAl periToneal,
fallopIan, or
oVarian cancEr
(CREATIVE Phase IA Trial)
Inhibited DNA repair,
provided anti-tumor activity [349]
PP2ANCT01837667Phase 1Tumors
Neoplasms
Cancer
Drug: LB-100
for injection
Drug: Docetaxel
Phase I study
of LB-100 with
Docetaxel in
solid tumors
RP2D is 2.33 mg/m2
for 3 consecutive days every 3 weeks [351]
NCT03027388Phase 2Astrocytoma
Grades II, III, and IV
Glioblastoma
Multiforme
Giant cell
Glioblastoma
Glioma
Oligodendrogliomas
Drug: LB-100Protein phosphatase 2A inhibitor,
in recurrent
glioblastoma
Inhibited cancer cells growth [352]

5. Challenge and Perspective

Kinases and phosphatases have essential function on coordinating DNA-repair pathways. Dysregulation of these enzymes leads to enhanced genomic instability, which is a hallmark of cancer. Understanding the functions of kinases and phosphatases in DNA-repair regulation is essential for developing effective cancer therapeutics.
Kinases and phosphatases work in opposition to each other, and the balance between their activities is critical for proper DDR signaling and cancer-cell survival. Therefore, developing inhibitors to modulate kinase or phosphatase activity to disrupt DDR signaling is a promising strategy for cancer therapeutics. However, one of the main challenges in studying kinases and phosphatases in DNA repair is identifying specific targets for therapeutic intervention. There are numerous kinases and phosphatases involved in DNA repair. However, many of these enzymes have multiple functions and substrates. For example, ATM is an essential kinase involved in DNA repair, but it also plays roles in other cellular processes, such as RNA transcription or energy metabolism which also affect the survival of normal cells. The safety of these inhibitors should be intensively evaluated in clinical trials. Therefore, developing kinases and phosphatases inhibitors that specifically target the DNA-repair pathway but with fewer side-effects on other signaling is important. Another challenge is that inhibitors of DNA-repair-related kinases and phosphatases have shown efficacy in treating certain types of cancer, whereas their use has not been approved for all types of cancer in clinical trials. Therefore, identifying biomarkers that can predict response to the inhibitors of DNA-repair-related kinases or phosphatases will help clinicians to identify which groups of patients are more likely to respond to these inhibitors. Predictive biomarkers can help optimize treatment strategies and improve patient outcomes. The use of deep sequencing approaches to identify specific genetic alterations in cancer cells may provide opportunities for precision therapies that selectively target kinases or phosphatases in DNA-repair pathways.
Despite these challenges, recent studies have provided several promising options for the development of cancer therapeutics targeting kinases and phosphatases in DNA repair. Multiple inhibitors targeting kinases or phosphatases involved in DNA repair have shown promise in pre-clinical studies or clinical trials which may provide us with alternative choices in the toolbox for cancer therapeutics. Besides, the combination of traditional radio-therapy or chemo-therapy with kinase or phosphatase inhibitors is also a potential strategy for cancer treatment. Cancer cells frequently develop resistance to radio-therapy or single-agent chemo-therapies by acquiring altered DNA-repair activity. Therefore, a combination treatment with kinase or phosphatase inhibitors may disrupt the DNA-repair ability of cancer cells under radio- or chemo-therapy and may enhance therapeutics efficacy.
In conclusion, studies on kinases and phosphatases in DNA repair are rapidly expanding fields that provide valuable insights into the mechanisms involved in genome stability, cancer development, and cancer therapeutics. While there are still significant challenges and limitations on developing targeted therapeutics, further studies on the development of novel inhibitors and combination therapies are crucial for improving cancer therapeutics and reducing mortality rates.

Author Contributions

Conceptualization, F.Z. and W.K.; writing—original draft preparation, S.Q. and I.K.; writing—review and editing, F.Z., W.K., Y.H., and S.Q.; supervision, F.Z. and W.K. project administration, W.K.; funding acquisition, W.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by funding from the National Research Foundation of Korea (NRF) (grant number 2022R1A2C1091563 and grant number 2019M3E5D3073092) and the Soonchunhyang University Research Fund.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no competing interests.

References

  1. Zhao, F.; Kim, W.; Kloeber, J.A.; Lou, Z. DNA end resection and its role in DNA replication and DSB repair choice in mammalian cells. Exp. Mol. Med. 2020, 52, 1705–1714. [Google Scholar] [CrossRef] [PubMed]
  2. Huen, M.S.; Chen, J. The DNA damage response pathways: At the crossroad of protein modifications. Cell Res. 2008, 18, 8–16. [Google Scholar] [CrossRef] [Green Version]
  3. Bakkenist, C.J.; Kastan, M.B. Phosphatases join kinases in DNA-damage response pathways. Trends Cell Biol. 2004, 14, 339–341. [Google Scholar] [CrossRef] [PubMed]
  4. Blackford, A.N.; Jackson, S.P. ATM, ATR, and DNA-PK: The Trinity at the Heart of the DNA Damage Response. Mol. Cell 2017, 66, 801–817. [Google Scholar] [CrossRef] [Green Version]
  5. Shimada, M.; Nakanishi, M. Response to DNA damage: Why do we need to focus on protein phosphatases? Front. Oncol. 2013, 3, 8. [Google Scholar] [CrossRef] [Green Version]
  6. Lanz, M.C.; Dibitetto, D.; Smolka, M.B. DNA damage kinase signaling: Checkpoint and repair at 30 years. EMBO J. 2019, 38, e101801. [Google Scholar] [CrossRef]
  7. Chen, S.H.; Albuquerque, C.P.; Liang, J.; Suhandynata, R.T.; Zhou, H. A proteome-wide analysis of kinase-substrate network in the DNA damage response. J. Biol. Chem. 2010, 285, 12803–12812. [Google Scholar] [CrossRef] [Green Version]
  8. Summers, K.C.; Shen, F.; Sierra Potchanant, E.A.; Phipps, E.A.; Hickey, R.J.; Malkas, L.H. Phosphorylation: The molecular switch of double-strand break repair. Int. J. Proteom. 2011, 2011, 373816. [Google Scholar] [CrossRef] [Green Version]
  9. Menolfi, D.; Zha, S. ATM, ATR and DNA-PKcs kinases-the lessons from the mouse models: Inhibition not equal deletion. Cell Biosci. 2020, 10, 8. [Google Scholar] [CrossRef]
  10. Shibata, A.; Jeggo, P.A. ATM’s Role in the Repair of DNA Double-Strand Breaks. Genes 2021, 12, 1370. [Google Scholar] [CrossRef]
  11. Hartlerode, A.J.; Scully, R. Mechanisms of double-strand break repair in somatic mammalian cells. Biochem. J. 2009, 423, 157–168. [Google Scholar] [CrossRef] [Green Version]
  12. Pannunzio, N.R.; Watanabe, G.; Lieber, M.R. Nonhomologous DNA end-joining for repair of DNA double-strand breaks. J. Biol. Chem. 2018, 293, 10512–10523. [Google Scholar] [CrossRef] [Green Version]
  13. Paques, F.; Haber, J.E. Multiple pathways of recombination induced by double-strand breaks in Saccharomyces cerevisiae. Microbiol. Mol. Biol. Rev. 1999, 63, 349–404. [Google Scholar] [CrossRef] [Green Version]
  14. Sung, P.; Klein, H. Mechanism of homologous recombination: Mediators and helicases take on regulatory functions. Nat. Rev. Mol. Cell Biol. 2006, 7, 739–750. [Google Scholar] [CrossRef]
  15. Beucher, A.; Birraux, J.; Tchouandong, L.; Barton, O.; Shibata, A.; Conrad, S.; Goodarzi, A.A.; Krempler, A.; Jeggo, P.A.; Lobrich, M. ATM and Artemis promote homologous recombination of radiation-induced DNA double-strand breaks in G2. EMBO J. 2009, 28, 3413–3427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Karanam, K.; Kafri, R.; Loewer, A.; Lahav, G. Quantitative live cell imaging reveals a gradual shift between DNA repair mechanisms and a maximal use of HR in mid S phase. Mol. Cell 2012, 47, 320–329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Bakkenist, C.J.; Kastan, M.B. DNA damage activates ATM through intermolecular autophosphorylation and dimer dissociation. Nature 2003, 421, 499–506. [Google Scholar] [CrossRef] [PubMed]
  18. Bakr, A.; Oing, C.; Kocher, S.; Borgmann, K.; Dornreiter, I.; Petersen, C.; Dikomey, E.; Mansour, W.Y. Involvement of ATM in homologous recombination after end resection and RAD51 nucleofilament formation. Nucleic Acids Res. 2015, 43, 3154–3166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Khanna, K.K.; Lavin, M.F.; Jackson, S.P.; Mulhern, T.D. ATM, a central controller of cellular responses to DNA damage. Cell Death Differ. 2001, 8, 1052–1065. [Google Scholar] [CrossRef] [Green Version]
  20. Marechal, A.; Zou, L. DNA damage sensing by the ATM and ATR kinases. Cold Spring Harb. Perspect. Biol. 2013, 5, a012716. [Google Scholar] [CrossRef]
  21. Saldivar, J.C.; Cortez, D.; Cimprich, K.A. The essential kinase ATR: Ensuring faithful duplication of a challenging genome. Nat. Rev. Mol. Cell Biol. 2017, 18, 622–636. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Falck, J.; Coates, J.; Jackson, S.P. Conserved modes of recruitment of ATM, ATR and DNA-PKcs to sites of DNA damage. Nature 2005, 434, 605–611. [Google Scholar] [CrossRef] [PubMed]
  23. Singleton, B.K.; Torres-Arzayus, M.I.; Rottinghaus, S.T.; Taccioli, G.E.; Jeggo, P.A. The C terminus of Ku80 activates the DNA-dependent protein kinase catalytic subunit. Mol. Cell. Biol. 1999, 19, 3267–3277. [Google Scholar] [CrossRef] [Green Version]
  24. Zou, L.; Elledge, S.J. Sensing DNA damage through ATRIP recognition of RPA-ssDNA complexes. Science 2003, 300, 1542–1548. [Google Scholar] [CrossRef] [Green Version]
  25. Shibata, A. Regulation of repair pathway choice at two-ended DNA double-strand breaks. Mutat. Res. 2017, 803–805, 51–55. [Google Scholar] [CrossRef] [PubMed]
  26. Shibata, A.; Jeggo, P. A historical reflection on our understanding of radiation-induced DNA double strand break repair in somatic mammalian cells; interfacing the past with the present. Int. J. Radiat. Biol. 2019, 95, 945–956. [Google Scholar] [CrossRef] [Green Version]
  27. Chen, B.P.; Uematsu, N.; Kobayashi, J.; Lerenthal, Y.; Krempler, A.; Yajima, H.; Lobrich, M.; Shiloh, Y.; Chen, D.J. Ataxia telangiectasia mutated (ATM) is essential for DNA-PKcs phosphorylations at the Thr-2609 cluster upon DNA double strand break. J. Biol. Chem. 2007, 282, 6582–6587. [Google Scholar] [CrossRef] [Green Version]
  28. Yajima, H.; Lee, K.J.; Zhang, S.; Kobayashi, J.; Chen, B.P. DNA double-strand break formation upon UV-induced replication stress activates ATM and DNA-PKcs kinases. J. Mol. Biol. 2009, 385, 800–810. [Google Scholar] [CrossRef] [Green Version]
  29. Ma, Y.; Pannicke, U.; Lu, H.; Niewolik, D.; Schwarz, K.; Lieber, M.R. The DNA-dependent protein kinase catalytic subunit phosphorylation sites in human Artemis. J. Biol. Chem. 2005, 280, 33839–33846. [Google Scholar] [CrossRef] [Green Version]
  30. Yannone, S.M.; Roy, S.; Chan, D.W.; Murphy, M.B.; Huang, S.; Campisi, J.; Chen, D.J. Werner syndrome protein is regulated and phosphorylated by DNA-dependent protein kinase. J. Biol. Chem. 2001, 276, 38242–38248. [Google Scholar] [CrossRef]
  31. Kozlov, S.V.; Graham, M.E.; Jakob, B.; Tobias, F.; Kijas, A.W.; Tanuji, M.; Chen, P.; Robinson, P.J.; Taucher-Scholz, G.; Suzuki, K.; et al. Autophosphorylation and ATM activation: Additional sites add to the complexity. J. Biol. Chem. 2011, 286, 9107–9119. [Google Scholar] [CrossRef] [Green Version]
  32. Kozlov, S.V.; Graham, M.E.; Peng, C.; Chen, P.; Robinson, P.J.; Lavin, M.F. Involvement of novel autophosphorylation sites in ATM activation. EMBO J. 2006, 25, 3504–3514. [Google Scholar] [CrossRef] [Green Version]
  33. Phan, L.M.; Rezaeian, A.H. ATM: Main Features, Signaling Pathways, and Its Diverse Roles in DNA Damage Response, Tumor Suppression, and Cancer Development. Genes 2021, 12, 845. [Google Scholar] [CrossRef] [PubMed]
  34. Liu, S.; Shiotani, B.; Lahiri, M.; Marechal, A.; Tse, A.; Leung, C.C.; Glover, J.N.; Yang, X.H.; Zou, L. ATR autophosphorylation as a molecular switch for checkpoint activation. Mol. Cell 2011, 43, 192–202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Matsuoka, S.; Rotman, G.; Ogawa, A.; Shiloh, Y.; Tamai, K.; Elledge, S.J. Ataxia telangiectasia-mutated phosphorylates Chk2 in vivo and in vitro. Proc. Natl. Acad. Sci. USA 2000, 97, 10389–10394. [Google Scholar] [CrossRef] [Green Version]
  36. Xu, X.; Tsvetkov, L.M.; Stern, D.F. Chk2 activation and phosphorylation-dependent oligomerization. Mol. Cell. Biol. 2002, 22, 4419–4432. [Google Scholar] [CrossRef] [Green Version]
  37. Melchionna, R.; Chen, X.B.; Blasina, A.; McGowan, C.H. Threonine 68 is required for radiation-induced phosphorylation and activation of Cds1. Nat. Cell Biol. 2000, 2, 762–765. [Google Scholar] [CrossRef] [PubMed]
  38. Leung-Pineda, V.; Ryan, C.E.; Piwnica-Worms, H. Phosphorylation of Chk1 by ATR is antagonized by a Chk1-regulated protein phosphatase 2A circuit. Mol. Cell. Biol. 2006, 26, 7529–7538. [Google Scholar] [CrossRef] [Green Version]
  39. Patil, M.; Pabla, N.; Dong, Z. Checkpoint kinase 1 in DNA damage response and cell cycle regulation. Cell. Mol. Life Sci. 2013, 70, 4009–4021. [Google Scholar] [CrossRef] [Green Version]
  40. Stolz, A.; Ertych, N.; Bastians, H. Tumor suppressor CHK2: Regulator of DNA damage response and mediator of chromosomal stability. Clin. Cancer Res. 2011, 17, 401–405. [Google Scholar] [CrossRef] [Green Version]
  41. Mahajan, K.; Mahajan, N.P. Cross talk of tyrosine kinases with the DNA damage signaling pathways. Nucleic Acids Res. 2015, 43, 10588–10601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Moon, J.; Kitty, I.; Renata, K.; Qin, S.; Zhao, F.; Kim, W. DNA Damage and Its Role in Cancer Therapeutics. Int. J. Mol. Sci. 2023, 24, 4741. [Google Scholar] [CrossRef] [PubMed]
  43. Bartkova, J.; Horejsi, Z.; Koed, K.; Kramer, A.; Tort, F.; Zieger, K.; Guldberg, P.; Sehested, M.; Nesland, J.M.; Lukas, C.; et al. DNA damage response as a candidate anti-cancer barrier in early human tumorigenesis. Nature 2005, 434, 864–870. [Google Scholar] [CrossRef]
  44. Macheret, M.; Halazonetis, T.D. DNA replication stress as a hallmark of cancer. Annu. Rev. Pathol. 2015, 10, 425–448. [Google Scholar] [CrossRef] [Green Version]
  45. Chen, Y.; Li, Y.; Xiong, J.; Lan, B.; Wang, X.; Liu, J.; Lin, J.; Fei, Z.; Zheng, X.; Chen, C. Role of PRKDC in cancer initiation, progression, and treatment. Cancer Cell Int. 2021, 21, 563. [Google Scholar] [CrossRef]
  46. Bychkovsky, B.L.; Agaoglu, N.B.; Horton, C.; Zhou, J.; Yussuf, A.; Hemyari, P.; Richardson, M.E.; Young, C.; LaDuca, H.; McGuinness, D.L.; et al. Differences in Cancer Phenotypes Among Frequent CHEK2 Variants and Implications for Clinical Care-Checking CHEK2. JAMA Oncol. 2022, 8, 1598–1606. [Google Scholar] [CrossRef]
  47. Zhang, Y.; Hunter, T. Roles of Chk1 in cell biology and cancer therapy. Int. J. Cancer 2014, 134, 1013–1023. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Rothblum-Oviatt, C.; Wright, J.; Lefton-Greif, M.A.; McGrath-Morrow, S.A.; Crawford, T.O.; Lederman, H.M. Ataxia telangiectasia: A review. Orphanet J. Rare Dis. 2016, 11, 159. [Google Scholar] [CrossRef] [Green Version]
  49. Sharma, R.; Lewis, S.; Wlodarski, M.W. DNA Repair Syndromes and Cancer: Insights Into Genetics and Phenotype Patterns. Front. Pediatr. 2020, 8, 570084. [Google Scholar] [CrossRef]
  50. Ichisima, J.; Suzuki, N.M.; Samata, B.; Awaya, T.; Takahashi, J.; Hagiwara, M.; Nakahata, T.; Saito, M.K. Verification and rectification of cell type-specific splicing of a Seckel syndrome-associated ATR mutation using iPS cell model. J. Hum. Genet. 2019, 64, 445–458. [Google Scholar] [CrossRef] [Green Version]
  51. Morio, T. Recent advances in the study of immunodeficiency and DNA damage response. Int. J. Hematol. 2017, 106, 357–365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Mathieu, A.L.; Verronese, E.; Rice, G.I.; Fouyssac, F.; Bertrand, Y.; Picard, C.; Chansel, M.; Walter, J.E.; Notarangelo, L.D.; Butte, M.J.; et al. PRKDC mutations associated with immunodeficiency, granuloma, and autoimmune regulator-dependent autoimmunity. J. Allergy Clin. Immunol. 2015, 135, 1578–1588.e5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Landrum, M.J.; Lee, J.M.; Benson, M.; Brown, G.R.; Chao, C.; Chitipiralla, S.; Gu, B.; Hart, J.; Hoffman, D.; Jang, W.; et al. ClinVar: Improving access to variant interpretations and supporting evidence. Nucleic Acids Res. 2018, 46, D1062–D1067. [Google Scholar] [CrossRef] [Green Version]
  54. Tate, J.G.; Bamford, S.; Jubb, H.C.; Sondka, Z.; Beare, D.M.; Bindal, N.; Boutselakis, H.; Cole, C.G.; Creatore, C.; Dawson, E.; et al. COSMIC: The Catalogue Of Somatic Mutations In Cancer. Nucleic Acids Res. 2019, 47, D941–D947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Chakravarty, D.; Gao, J.; Phillips, S.M.; Kundra, R.; Zhang, H.; Wang, J.; Rudolph, J.E.; Yaeger, R.; Soumerai, T.; Nissan, M.H.; et al. OncoKB: A Precision Oncology Knowledge Base. JCO Precis. Oncol. 2017, 2017, 1–16. [Google Scholar] [CrossRef]
  56. Bertolotti, A. The split protein phosphatase system. Biochem. J. 2018, 475, 3707–3723. [Google Scholar] [CrossRef] [Green Version]
  57. Campos, A.; Clemente-Blanco, A. Cell Cycle and DNA Repair Regulation in the Damage Response: Protein Phosphatases Take Over the Reins. Int. J. Mol. Sci. 2020, 21, 446. [Google Scholar] [CrossRef] [Green Version]
  58. Peng, A.; Yamamoto, T.M.; Goldberg, M.L.; Maller, J.L. A novel role for greatwall kinase in recovery from DNA damage. Cell Cycle 2010, 9, 4364–4369. [Google Scholar] [CrossRef] [Green Version]
  59. Landsverk, H.B.; Mora-Bermudez, F.; Landsverk, O.J.; Hasvold, G.; Naderi, S.; Bakke, O.; Ellenberg, J.; Collas, P.; Syljuasen, R.G.; Kuntziger, T. The protein phosphatase 1 regulator PNUTS is a new component of the DNA damage response. EMBO Rep. 2010, 11, 868–875. [Google Scholar] [CrossRef] [Green Version]
  60. Kuntziger, T.; Landsverk, H.B.; Collas, P.; Syljuasen, R.G. Protein phosphatase 1 regulators in DNA damage signaling. Cell Cycle 2011, 10, 1356–1362. [Google Scholar] [CrossRef] [Green Version]
  61. Bradshaw, R.A.; Dennis, E.A. Handbook of Cell Signaling; Academic Press: San Diego, CA, USA, 2004. [Google Scholar]
  62. Ramos, F.; Villoria, M.T.; Alonso-Rodriguez, E.; Clemente-Blanco, A. Role of protein phosphatases PP1, PP2A, PP4 and Cdc14 in the DNA damage response. Cell Stress 2019, 3, 70–85. [Google Scholar] [CrossRef] [Green Version]
  63. Chen, M.X.; McPartlin, A.E.; Brown, L.; Chen, Y.H.; Barker, H.M.; Cohen, P.T. A novel human protein serine/threonine phosphatase, which possesses four tetratricopeptide repeat motifs and localizes to the nucleus. EMBO J. 1994, 13, 4278–4290. [Google Scholar] [CrossRef]
  64. Wechsler, T.; Chen, B.P.; Harper, R.; Morotomi-Yano, K.; Huang, B.C.; Meek, K.; Cleaver, J.E.; Chen, D.J.; Wabl, M. DNA-PKcs function regulated specifically by protein phosphatase 5. Proc. Natl. Acad. Sci. USA 2004, 101, 1247–1252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Wang, J.; Shen, T.; Zhu, W.; Dou, L.; Gu, H.; Zhang, L.; Yang, Z.; Chen, H.; Zhou, Q.; Sanchez, E.R.; et al. Protein phosphatase 5 and the tumor suppressor p53 down-regulate each other’s activities in mice. J. Biol. Chem. 2018, 293, 18218–18229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Douglas, P.; Zhong, J.; Ye, R.; Moorhead, G.B.; Xu, X.; Lees-Miller, S.P. Protein phosphatase 6 interacts with the DNA-dependent protein kinase catalytic subunit and dephosphorylates gamma-H2AX. Mol. Cell. Biol. 2010, 30, 1368–1381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Zhong, J.; Liao, J.; Liu, X.; Wang, P.; Liu, J.; Hou, W.; Zhu, B.; Yao, L.; Wang, J.; Li, J.; et al. Protein phosphatase PP6 is required for homology-directed repair of DNA double-strand breaks. Cell Cycle 2011, 10, 1411–1419. [Google Scholar] [CrossRef] [Green Version]
  68. Takekawa, M.; Adachi, M.; Nakahata, A.; Nakayama, I.; Itoh, F.; Tsukuda, H.; Taya, Y.; Imai, K. p53-inducible wip1 phosphatase mediates a negative feedback regulation of p38 MAPK-p53 signaling in response to UV radiation. EMBO J. 2000, 19, 6517–6526. [Google Scholar] [CrossRef]
  69. Lu, X.; Nannenga, B.; Donehower, L.A. PPM1D dephosphorylates Chk1 and p53 and abrogates cell cycle checkpoints. Genes Dev. 2005, 19, 1162–1174. [Google Scholar] [CrossRef] [Green Version]
  70. Fujimoto, H.; Onishi, N.; Kato, N.; Takekawa, M.; Xu, X.Z.; Kosugi, A.; Kondo, T.; Imamura, M.; Oishi, I.; Yoda, A.; et al. Regulation of the antioncogenic Chk2 kinase by the oncogenic Wip1 phosphatase. Cell Death Differ. 2006, 13, 1170–1180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Cha, H.; Lowe, J.M.; Li, H.; Lee, J.S.; Belova, G.I.; Bulavin, D.V.; Fornace, A.J., Jr. Wip1 directly dephosphorylates gamma-H2AX and attenuates the DNA damage response. Cancer Res. 2010, 70, 4112–4122. [Google Scholar] [CrossRef] [Green Version]
  72. Macurek, L.; Lindqvist, A.; Voets, O.; Kool, J.; Vos, H.R.; Medema, R.H. Wip1 phosphatase is associated with chromatin and dephosphorylates gammaH2AX to promote checkpoint inhibition. Oncogene 2010, 29, 2281–2291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Kimura, H.; Takizawa, N.; Allemand, E.; Hori, T.; Iborra, F.J.; Nozaki, N.; Muraki, M.; Hagiwara, M.; Krainer, A.R.; Fukagawa, T.; et al. A novel histone exchange factor, protein phosphatase 2Cgamma, mediates the exchange and dephosphorylation of H2A-H2B. J. Cell Biol. 2006, 175, 389–400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Pei, J.J.; Sersen, E.; Iqbal, K.; Grundke-Iqbal, I. Expression of protein phosphatases (PP-1, PP-2A, PP-2B and PTP-1B) and protein kinases (MAP kinase and P34cdc2) in the hippocampus of patients with Alzheimer disease and normal aged individuals. Brain Res. 1994, 655, 70–76. [Google Scholar] [CrossRef]
  75. Gao, C.; Cao, N.; Wang, Y. Metal dependent protein phosphatase PPM family in cardiac health and diseases. Cell. Signal. 2021, 85, 110061. [Google Scholar] [CrossRef]
  76. Vintem, A.P.; Henriques, A.G.; da Cruz, E.S.O.A.; da Cruz, E.S.E.F. PP1 inhibition by Abeta peptide as a potential pathological mechanism in Alzheimer’s disease. Neurotoxicol. Teratol. 2009, 31, 85–88. [Google Scholar] [CrossRef]
  77. Baskaran, R.; Velmurugan, B.K. Protein phosphatase 2A as therapeutic targets in various disease models. Life Sci. 2018, 210, 40–46. [Google Scholar] [CrossRef] [PubMed]
  78. Cerami, E.; Gao, J.; Dogrusoz, U.; Gross, B.E.; Sumer, S.O.; Aksoy, B.A.; Jacobsen, A.; Byrne, C.J.; Heuer, M.L.; Larsson, E.; et al. The cBio cancer genomics portal: An open platform for exploring multidimensional cancer genomics data. Cancer Discov. 2012, 2, 401–404. [Google Scholar] [CrossRef] [Green Version]
  79. Takakura, S.; Kohno, T.; Manda, R.; Okamoto, A.; Tanaka, T.; Yokota, J. Genetic alterations and expression of the protein phosphatase 1 genes in human cancers. Int. J. Oncol. 2001, 18, 817–824. [Google Scholar] [CrossRef]
  80. Freeman, A.K.; Monteiro, A.N. Phosphatases in the cellular response to DNA damage. Cell Commun. Signal. 2010, 8, 27. [Google Scholar] [CrossRef] [Green Version]
  81. Stiff, T.; Walker, S.A.; Cerosaletti, K.; Goodarzi, A.A.; Petermann, E.; Concannon, P.; O’Driscoll, M.; Jeggo, P.A. ATR-dependent phosphorylation and activation of ATM in response to UV treatment or replication fork stalling. EMBO J. 2006, 25, 5775–5782. [Google Scholar] [CrossRef]
  82. Yang, C.; Tang, X.; Guo, X.; Niikura, Y.; Kitagawa, K.; Cui, K.; Wong, S.T.; Fu, L.; Xu, B. Aurora-B mediated ATM serine 1403 phosphorylation is required for mitotic ATM activation and the spindle checkpoint. Mol. Cell 2011, 44, 597–608. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Tian, B.; Yang, Q.; Mao, Z. Phosphorylation of ATM by Cdk5 mediates DNA damage signalling and regulates neuronal death. Nat. Cell Biol. 2009, 11, 211–218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Lee, H.J.; Lan, L.; Peng, G.; Chang, W.C.; Hsu, M.C.; Wang, Y.N.; Cheng, C.C.; Wei, L.; Nakajima, S.; Chang, S.S.; et al. Tyrosine 370 phosphorylation of ATM positively regulates DNA damage response. Cell Res. 2015, 25, 225–236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Shreeram, S.; Demidov, O.N.; Hee, W.K.; Yamaguchi, H.; Onishi, N.; Kek, C.; Timofeev, O.N.; Dudgeon, C.; Fornace, A.J.; Anderson, C.W.; et al. Wip1 phosphatase modulates ATM-dependent signaling pathways. Mol. Cell 2006, 23, 757–764. [Google Scholar] [CrossRef]
  86. Goodarzi, A.A.; Jonnalagadda, J.C.; Douglas, P.; Young, D.; Ye, R.; Moorhead, G.B.; Lees-Miller, S.P.; Khanna, K.K. Autophosphorylation of ataxia-telangiectasia mutated is regulated by protein phosphatase 2A. EMBO J. 2004, 23, 4451–4461. [Google Scholar] [CrossRef] [Green Version]
  87. Wang, X.; Zeng, L.; Wang, J.; Chau, J.F.; Lai, K.P.; Jia, D.; Poonepalli, A.; Hande, M.P.; Liu, H.; He, G.; et al. A positive role for c-Abl in Atm and Atr activation in DNA damage response. Cell Death Differ. 2011, 18, 5–15. [Google Scholar] [CrossRef] [Green Version]
  88. Jarrett, S.G.; Horrell, E.M.W.; Christian, P.A.; Vanover, J.C.; Boulanger, M.C.; Zou, Y.; D’Orazio, J.A. PKA-mediated phosphorylation of ATR promotes recruitment of XPA to UV-induced DNA damage. Mol. Cell 2014, 54, 999–1011. [Google Scholar] [CrossRef] [Green Version]
  89. Yajima, H.; Lee, K.J.; Chen, B.P. ATR-dependent phosphorylation of DNA-dependent protein kinase catalytic subunit in response to UV-induced replication stress. Mol. Cell. Biol. 2006, 26, 7520–7528. [Google Scholar] [CrossRef] [Green Version]
  90. Douglas, P.; Sapkota, G.P.; Morrice, N.; Yu, Y.; Goodarzi, A.A.; Merkle, D.; Meek, K.; Alessi, D.R.; Lees-Miller, S.P. Identification of in vitro and in vivo phosphorylation sites in the catalytic subunit of the DNA-dependent protein kinase. Biochem. J. 2002, 368 Pt 1, 243–251. [Google Scholar] [CrossRef] [Green Version]
  91. Fan, X.J.; Wang, Y.L.; Zhao, W.W.; Bai, S.M.; Ma, Y.; Yin, X.K.; Feng, L.L.; Feng, W.X.; Wang, Y.N.; Liu, Q.; et al. NONO phase separation enhances DNA damage repair by accelerating nuclear EGFR-induced DNA-PK activation. Am. J. Cancer Res. 2021, 11, 2838–2852. [Google Scholar]
  92. Douglas, P.; Ye, R.; Trinkle-Mulcahy, L.; Neal, J.A.; De Wever, V.; Morrice, N.A.; Meek, K.; Lees-Miller, S.P. Polo-like kinase 1 (PLK1) and protein phosphatase 6 (PP6) regulate DNA-dependent protein kinase catalytic subunit (DNA-PKcs) phosphorylation in mitosis. Biosci. Rep. 2014, 34, e00113. [Google Scholar] [CrossRef] [PubMed]
  93. Caporali, S.; Falcinelli, S.; Starace, G.; Russo, M.T.; Bonmassar, E.; Jiricny, J.; D’Atri, S. DNA damage induced by temozolomide signals to both ATM and ATR: Role of the mismatch repair system. Mol. Pharmacol. 2004, 66, 478–491. [Google Scholar] [PubMed]
  94. Shtivelman, E.; Sussman, J.; Stokoe, D. A role for PI 3-kinase and PKB activity in the G2/M phase of the cell cycle. Curr. Biol. 2002, 12, 919–924. [Google Scholar] [CrossRef] [Green Version]
  95. Shiromizu, T.; Goto, H.; Tomono, Y.; Bartek, J.; Totsukawa, G.; Inoko, A.; Nakanishi, M.; Matsumura, F.; Inagaki, M. Regulation of mitotic function of Chk1 through phosphorylation at novel sites by cyclin-dependent kinase 1 (Cdk1). Genes Cells 2006, 11, 477–485. [Google Scholar] [CrossRef]
  96. Moiseeva, T.N.; Yin, Y.; Calderon, M.J.; Qian, C.; Schamus-Haynes, S.; Sugitani, N.; Osmanbeyoglu, H.U.; Rothenberg, E.; Watkins, S.C.; Bakkenist, C.J. An ATR and CHK1 kinase signaling mechanism that limits origin firing during unperturbed DNA replication. Proc. Natl. Acad. Sci. USA 2019, 116, 13374–13383. [Google Scholar] [CrossRef] [Green Version]
  97. Li, P.; Goto, H.; Kasahara, K.; Matsuyama, M.; Wang, Z.; Yatabe, Y.; Kiyono, T.; Inagaki, M. P90 RSK arranges Chk1 in the nucleus for monitoring of genomic integrity during cell proliferation. Mol. Biol. Cell 2012, 23, 1582–1592. [Google Scholar] [CrossRef] [PubMed]
  98. Lovly, C.M.; Yan, L.; Ryan, C.E.; Takada, S.; Piwnica-Worms, H. Regulation of Chk2 ubiquitination and signaling through autophosphorylation of serine 379. Mol. Cell. Biol. 2008, 28, 5874–5885. [Google Scholar] [CrossRef] [Green Version]
  99. Li, J.; Stern, D.F. Regulation of CHK2 by DNA-dependent protein kinase. J. Biol. Chem. 2005, 280, 12041–12050. [Google Scholar] [CrossRef] [Green Version]
  100. Chowdhury, M.A.N.; Wang, S.W.; Suen, C.S.; Hwang, M.J.; Hsueh, Y.A.; Shieh, S.Y. JAK2-CHK2 signaling safeguards the integrity of the mitotic spindle assembly checkpoint and genome stability. Cell Death Dis. 2022, 13, 619. [Google Scholar] [CrossRef]
  101. Tsvetkov, L.; Xu, X.; Li, J.; Stern, D.F. Polo-like kinase 1 and Chk2 interact and co-localize to centrosomes and the midbody. J. Biol. Chem. 2003, 278, 8468–8475. [Google Scholar] [CrossRef] [Green Version]
  102. Tosti, E.; Waldbaum, L.; Warshaw, G.; Gross, E.A.; Ruggieri, R. The stress kinase MRK contributes to regulation of DNA damage checkpoints through a p38gamma-independent pathway. J. Biol. Chem. 2004, 279, 47652–47660. [Google Scholar] [CrossRef] [Green Version]
  103. Carlessi, L.; Buscemi, G.; Fontanella, E.; Delia, D. A protein phosphatase feedback mechanism regulates the basal phosphorylation of Chk2 kinase in the absence of DNA damage. Biochim. Biophys. Acta 2010, 1803, 1213–1223. [Google Scholar] [CrossRef] [Green Version]
  104. Ziegler, V.; Deussen, M.; Schumacher, L.; Roos, W.P.; Fritz, G. Anticancer drug and ionizing radiation-induced DNA damage differently influences transcription activity and DDR-related stress responses of an endothelial monolayer. Biochim. Biophys. Acta Mol. Cell Res. 2020, 1867, 118678. [Google Scholar] [CrossRef]
  105. Li, A.; Yu, Y.; Lee, S.C.; Ishibashi, T.; Lees-Miller, S.P.; Ausio, J. Phosphorylation of histone H2A. X by DNA-dependent protein kinase is not affected by core histone acetylation, but it alters nucleosome stability and histone H1 binding. J. Biol. Chem. 2010, 285, 17778–17788. [Google Scholar] [CrossRef] [Green Version]
  106. Basnet, H.; Su, X.B.; Tan, Y.; Meisenhelder, J.; Merkurjev, D.; Ohgi, K.A.; Hunter, T.; Pillus, L.; Rosenfeld, M.G. Tyrosine phosphorylation of histone H2A by CK2 regulates transcriptional elongation. Nature 2014, 516, 267–271. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Liu, Y.; Long, Y.H.; Wang, S.Q.; Zhang, Y.Y.; Li, Y.F.; Mi, J.S.; Yu, C.H.; Li, D.Y.; Zhang, J.H.; Zhang, X.J. JMJD6 regulates histone H2A. X phosphorylation and promotes autophagy in triple-negative breast cancer cells via a novel tyrosine kinase activity. Oncogene 2019, 38, 980–997. [Google Scholar] [CrossRef]
  108. Wen, W.; Zhu, F.; Zhang, J.; Keum, Y.S.; Zykova, T.; Yao, K.; Peng, C.; Zheng, D.; Cho, Y.Y.; Ma, W.Y.; et al. MST1 promotes apoptosis through phosphorylation of histone H2AX. J. Biol. Chem. 2010, 285, 39108–39116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Guo, Z.F.; Kong, F.L. Akt regulates RSK2 to alter phosphorylation level of H2A.X in breast cancer. Oncol. Lett. 2021, 21, 187. [Google Scholar] [CrossRef]
  110. Liu, Y.; Long, Y.H.; Wang, S.Q.; Li, Y.F.; Zhang, J.H. Phosphorylation of H2A.X(T)(yr39) positively regulates DNA damage response and is linked to cancer progression. FEBS J. 2016, 283, 4462–4473. [Google Scholar] [CrossRef] [PubMed]
  111. Cook, P.J.; Ju, B.G.; Telese, F.; Wang, X.; Glass, C.K.; Rosenfeld, M.G. Tyrosine dephosphorylation of H2AX modulates apoptosis and survival decisions. Nature 2009, 458, 591–596. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Moon, S.H.; Lin, L.; Zhang, X.; Nguyen, T.A.; Darlington, Y.; Waldman, A.S.; Lu, X.; Donehower, L.A. Wild-type p53-induced phosphatase 1 dephosphorylates histone variant gamma-H2AX and suppresses DNA double strand break repair. J. Biol. Chem. 2010, 285, 12935–12947. [Google Scholar] [CrossRef] [Green Version]
  113. Nakada, S.; Chen, G.I.; Gingras, A.C.; Durocher, D. PP4 is a gamma H2AX phosphatase required for recovery from the DNA damage checkpoint. EMBO Rep. 2008, 9, 1019–1026. [Google Scholar] [CrossRef] [PubMed]
  114. Chowdhury, D.; Xu, X.; Zhong, X.; Ahmed, F.; Zhong, J.; Liao, J.; Dykxhoorn, D.M.; Weinstock, D.M.; Pfeifer, G.P.; Lieberman, J. A PP4-phosphatase complex dephosphorylates gamma-H2AX generated during DNA replication. Mol. Cell 2008, 31, 33–46. [Google Scholar] [CrossRef] [Green Version]
  115. Liu, J.; Luo, S.; Zhao, H.; Liao, J.; Li, J.; Yang, C.; Xu, B.; Stern, D.F.; Xu, X.; Ye, K. Structural mechanism of the phosphorylation-dependent dimerization of the MDC1 forkhead-associated domain. Nucleic Acids Res. 2012, 40, 3898–3912. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Luo, K.; Yuan, J.; Lou, Z. Oligomerization of MDC1 protein is important for proper DNA damage response. J. Biol. Chem. 2011, 286, 28192–28199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Stokes, M.P.; Rush, J.; Macneill, J.; Ren, J.M.; Sprott, K.; Nardone, J.; Yang, V.; Beausoleil, S.A.; Gygi, S.P.; Livingstone, M.; et al. Profiling of UV-induced ATM/ATR signaling pathways. Proc. Natl. Acad. Sci. USA 2007, 104, 19855–19860. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Wu, L.; Luo, K.; Lou, Z.; Chen, J. MDC1 regulates intra-S-phase checkpoint by targeting NBS1 to DNA double-strand breaks. Proc. Natl. Acad. Sci. USA 2008, 105, 11200–11205. [Google Scholar] [CrossRef] [Green Version]
  119. Li, Z.; Shao, C.; Kong, Y.; Carlock, C.; Ahmad, N.; Liu, X. DNA Damage Response-Independent Role for MDC1 in Maintaining Genomic Stability. Mol. Cell. Biol. 2017, 37, e00595-16. [Google Scholar] [CrossRef] [Green Version]
  120. Ando, K.; Ozaki, T.; Hirota, T.; Nakagawara, A. NFBD1/MDC1 is phosphorylated by PLK1 and controls G2/M transition through the regulation of a TOPOIIalpha-mediated decatenation checkpoint. PLoS ONE 2013, 8, e82744. [Google Scholar] [CrossRef]
  121. Lee, H.; Kwak, H.J.; Cho, I.T.; Park, S.H.; Lee, C.H. S1219 residue of 53BP1 is phosphorylated by ATM kinase upon DNA damage and required for proper execution of DNA damage response. Biochem. Biophys. Res. Commun. 2009, 378, 32–36. [Google Scholar] [CrossRef]
  122. Jiang, Y.; Dong, Y.; Luo, Y.; Jiang, S.; Meng, F.L.; Tan, M.; Li, J.; Zang, Y. AMPK-mediated phosphorylation on 53BP1 promotes c-NHEJ. Cell Rep. 2021, 34, 108713. [Google Scholar] [CrossRef]
  123. Jowsey, P.; Morrice, N.A.; Hastie, C.J.; McLauchlan, H.; Toth, R.; Rouse, J. Characterisation of the sites of DNA damage-induced 53BP1 phosphorylation catalysed by ATM and ATR. DNA Repair 2007, 6, 1536–1544. [Google Scholar] [CrossRef]
  124. Wang, H.; Peng, B.; Pandita, R.K.; Engler, D.A.; Matsunami, R.K.; Xu, X.; Hegde, P.M.; Butler, B.E.; Pandita, T.K.; Mitra, S.; et al. Aurora kinase B dependent phosphorylation of 53BP1 is required for resolving merotelic kinetochore-microtubule attachment errors during mitosis. Oncotarget 2017, 8, 48671–48687. [Google Scholar] [CrossRef] [Green Version]
  125. Benada, J.; Burdova, K.; Lidak, T.; von Morgen, P.; Macurek, L. Polo-like kinase 1 inhibits DNA damage response during mitosis. Cell Cycle 2015, 14, 219–231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. 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] [PubMed] [Green Version]
  127. Olson, E.; Nievera, C.J.; Klimovich, V.; Fanning, E.; Wu, X. RPA2 is a direct downstream target for ATR to regulate the S-phase checkpoint. J. Biol. Chem. 2006, 281, 39517–39533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Shiotani, B.; Nguyen, H.D.; Hakansson, P.; Marechal, A.; Tse, A.; Tahara, H.; Zou, L. Two distinct modes of ATR activation orchestrated by Rad17 and Nbs1. Cell Rep. 2013, 3, 1651–1662. [Google Scholar] [CrossRef] [Green Version]
  129. 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] [Green Version]
  130. Niu, H.; Erdjument-Bromage, H.; Pan, Z.Q.; Lee, S.H.; Tempst, P.; Hurwitz, J. Mapping of amino acid residues in the p34 subunit of human single-stranded DNA-binding protein phosphorylated by DNA-dependent protein kinase and Cdc2 kinase in vitro. J. Biol. Chem. 1997, 272, 12634–12641. [Google Scholar] [CrossRef] [Green Version]
  131. 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] [Green Version]
  132. Lee, D.H.; Pan, Y.; Kanner, S.; Sung, P.; Borowiec, J.A.; Chowdhury, D. A PP4 phosphatase complex dephosphorylates RPA2 to facilitate DNA repair via homologous recombination. Nat. Struct. Mol. Biol. 2010, 17, 365–372. [Google Scholar] [CrossRef] [PubMed]
  133. Schellenbauer, A.; Guilly, M.N.; Grall, R.; Le Bars, R.; Paget, V.; Kortulewski, T.; Sutcu, H.; Mathe, C.; Hullo, M.; Biard, D.; et al. Phospho-Ku70 induced by DNA damage interacts with RNA Pol II and promotes the formation of phospho-53BP1 foci to ensure optimal cNHEJ. Nucleic Acids Res. 2021, 49, 11728–11745. [Google Scholar] [CrossRef] [PubMed]
  134. Bouley, J.; Saad, L.; Grall, R.; Schellenbauer, A.; Biard, D.; Paget, V.; Morel-Altmeyer, S.; Guipaud, O.; Chambon, C.; Salles, B.; et al. A new phosphorylated form of Ku70 identified in resistant leukemic cells confers fast but unfaithful DNA repair in cancer cell lines. Oncotarget 2015, 6, 27980–28000. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Tanaka, H.; Kondo, K.; Fujita, K.; Homma, H.; Tagawa, K.; Jin, X.; Jin, M.; Yoshioka, Y.; Takayama, S.; Masuda, H.; et al. HMGB1 signaling phosphorylates Ku70 and impairs DNA damage repair in Alzheimer’s disease pathology. Commun. Biol. 2021, 4, 1175. [Google Scholar] [CrossRef]
  136. Morii, M.; Kubota, S.; Honda, T.; Yuki, R.; Morinaga, T.; Kuga, T.; Tomonaga, T.; Yamaguchi, N.; Yamaguchi, N. Src Acts as an Effector for Ku70-dependent Suppression of Apoptosis through Phosphorylation of Ku70 at Tyr-530. J. Biol. Chem. 2017, 292, 1648–1665. [Google Scholar] [CrossRef] [Green Version]
  137. Wu, X.; Ranganathan, V.; Weisman, D.S.; Heine, W.F.; Ciccone, D.N.; O’Neill, T.B.; Crick, K.E.; Pierce, K.A.; Lane, W.S.; Rathbun, G.; et al. ATM phosphorylation of Nijmegen breakage syndrome protein is required in a DNA damage response. Nature 2000, 405, 477–482. [Google Scholar] [CrossRef]
  138. Zhang, B.; Wang, E.; Dai, H.; Shen, J.; Hsieh, H.J.; Lu, X.; Peng, G. Phosphorylation of the BRCA1 C terminus (BRCT) repeat inhibitor of hTERT (BRIT1) protein coordinates TopBP1 protein recruitment and amplifies ataxia telangiectasia-mutated and Rad3-related (ATR) Signaling. J. Biol. Chem. 2014, 289, 34284–34295. [Google Scholar] [CrossRef] [Green Version]
  139. Gatei, M.; Young, D.; Cerosaletti, K.M.; Desai-Mehta, A.; Spring, K.; Kozlov, S.; Lavin, M.F.; Gatti, R.A.; Concannon, P.; Khanna, K. ATM-dependent phosphorylation of nibrin in response to radiation exposure. Nat. Genet. 2000, 25, 115–119. [Google Scholar] [CrossRef]
  140. Lim, D.S.; Kim, S.T.; Xu, B.; Maser, R.S.; Lin, J.; Petrini, J.H.; Kastan, M.B. ATM phosphorylates p95/nbs1 in an S-phase checkpoint pathway. Nature 2000, 404, 613–617. [Google Scholar] [CrossRef]
  141. Kim, S.T.; Lim, D.S.; Canman, C.E.; Kastan, M.B. Substrate specificities and identification of putative substrates of ATM kinase family members. J. Biol. Chem. 1999, 274, 37538–37543. [Google Scholar] [CrossRef] [Green Version]
  142. Zhao, S.; Weng, Y.C.; Yuan, S.S.; Lin, Y.T.; Hsu, H.C.; Lin, S.C.; Gerbino, E.; Song, M.H.; Zdzienicka, M.Z.; Gatti, R.A.; et al. Functional link between ataxia-telangiectasia and Nijmegen breakage syndrome gene products. Nature 2000, 405, 473–477. [Google Scholar] [CrossRef]
  143. Olson, E.; Nievera, C.J.; Lee, A.Y.; Chen, L.; Wu, X. The Mre11-Rad50-Nbs1 complex acts both upstream and downstream of ataxia telangiectasia mutated and Rad3-related protein (ATR) to regulate the S-phase checkpoint following UV treatment. J. Biol. Chem. 2007, 282, 22939–22952. [Google Scholar] [CrossRef] [Green Version]
  144. Wohlbold, L.; Merrick, K.A.; De, S.; Amat, R.; Kim, J.H.; Larochelle, S.; Allen, J.J.; Zhang, C.; Shokat, K.M.; Petrini, J.H.; et al. Chemical genetics reveals a specific requirement for Cdk2 activity in the DNA damage response and identifies Nbs1 as a Cdk2 substrate in human cells. PLoS Genet. 2012, 8, e1002935. [Google Scholar] [CrossRef]
  145. Rai, R.; Hu, C.; Broton, C.; Chen, Y.; Lei, M.; Chang, S. NBS1 Phosphorylation Status Dictates Repair Choice of Dysfunctional Telomeres. Mol. Cell 2017, 65, 801–817.e4. [Google Scholar] [CrossRef] [Green Version]
  146. Khanam, T.; Munoz, I.; Weiland, F.; Carroll, T.; Morgan, M.; Borsos, B.N.; Pantazi, V.; Slean, M.; Novak, M.; Toth, R.; et al. CDKL5 kinase controls transcription-coupled responses to DNA damage. EMBO J. 2021, 40, e108271. [Google Scholar] [CrossRef]
  147. Chini, C.C.; Chen, J. Repeated phosphopeptide motifs in human Claspin are phosphorylated by Chk1 and mediate Claspin function. J. Biol. Chem. 2006, 281, 33276–33282. [Google Scholar] [CrossRef] [Green Version]
  148. Peschiaroli, A.; Dorrello, N.V.; Guardavaccaro, D.; Venere, M.; Halazonetis, T.; Sherman, N.E.; Pagano, M. SCFbetaTrCP-mediated degradation of Claspin regulates recovery from the DNA replication checkpoint response. Mol. Cell 2006, 23, 319–329. [Google Scholar] [CrossRef] [PubMed]
  149. Nelson, A.C.; Lyons, T.R.; Young, C.D.; Hansen, K.C.; Anderson, S.M.; Holt, J.T. AKT regulates BRCA1 stability in response to hormone signaling. Mol. Cell. Endocrinol. 2010, 319, 129–142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Altiok, S.; Batt, D.; Altiok, N.; Papautsky, A.; Downward, J.; Roberts, T.M.; Avraham, H. Heregulin induces phosphorylation of BRCA1 through phosphatidylinositol 3-Kinase/AKT in breast cancer cells. J. Biol. Chem. 1999, 274, 32274–32278. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Gatei, M.; Scott, S.P.; Filippovitch, I.; Soronika, N.; Lavin, M.F.; Weber, B.; Khanna, K.K. Role for ATM in DNA damage-induced phosphorylation of BRCA1. Cancer Res. 2000, 60, 3299–3304. [Google Scholar]
  152. Tibbetts, R.S.; Cortez, D.; Brumbaugh, K.M.; Scully, R.; Livingston, D.; Elledge, S.J.; Abraham, R.T. Functional interactions between BRCA1 and the checkpoint kinase ATR during genotoxic stress. Genes Dev. 2000, 14, 2989–3002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Gatei, M.; Zhou, B.B.; Hobson, K.; Scott, S.; Young, D.; Khanna, K.K. Ataxia telangiectasia mutated (ATM) kinase and ATM and Rad3 related kinase mediate phosphorylation of Brca1 at distinct and overlapping sites. In vivo assessment using phospho-specific antibodies. J. Biol. Chem. 2001, 276, 17276–17280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Cortez, D.; Wang, Y.; Qin, J.; Elledge, S.J. Requirement of ATM-dependent phosphorylation of brca1 in the DNA damage response to double-strand breaks. Science 1999, 286, 1162–1166. [Google Scholar] [CrossRef] [PubMed]
  155. Ouchi, M.; Fujiuchi, N.; Sasai, K.; Katayama, H.; Minamishima, Y.A.; Ongusaha, P.P.; Deng, C.; Sen, S.; Lee, S.W.; Ouchi, T. BRCA1 phosphorylation by Aurora-A in the regulation of G2 to M transition. J. Biol. Chem. 2004, 279, 19643–19648. [Google Scholar] [CrossRef]
  156. Johnson, N.; Cai, D.; Kennedy, R.D.; Pathania, S.; Arora, M.; Li, Y.C.; D’Andrea, A.D.; Parvin, J.D.; Shapiro, G.I. Cdk1 participates in BRCA1-dependent S phase checkpoint control in response to DNA damage. Mol. Cell 2009, 35, 327–339. [Google Scholar] [CrossRef] [Green Version]
  157. Ruffner, H.; Jiang, W.; Craig, A.G.; Hunter, T.; Verma, I.M. BRCA1 is phosphorylated at serine 1497 in vivo at a cyclin-dependent kinase 2 phosphorylation site. Mol. Cell. Biol. 1999, 19, 4843–4854. [Google Scholar] [CrossRef] [Green Version]
  158. Kehn, K.; Berro, R.; Alhaj, A.; Bottazzi, M.E.; Yeh, W.I.; Klase, Z.; Van Duyne, R.; Fu, S.; Kashanchi, F. Functional consequences of cyclin D1/BRCA1 interaction in breast cancer cells. Oncogene 2007, 26, 5060–5069. [Google Scholar] [CrossRef] [Green Version]
  159. Lee, J.S.; Collins, K.M.; Brown, A.L.; Lee, C.H.; Chung, J.H. hCds1-mediated phosphorylation of BRCA1 regulates the DNA damage response. Nature 2000, 404, 201–204. [Google Scholar] [CrossRef]
  160. Kim, J.L.; Ha, G.H.; Campo, L.; Denning, M.F.; Patel, T.B.; Osipo, C.; Lin, S.Y.; Breuer, E.K. The role of Rak in the regulation of stability and function of BRCA1. Oncotarget 2017, 8, 86799–86815. [Google Scholar] [CrossRef] [Green Version]
  161. Chabalier-Taste, C.; Brichese, L.; Racca, C.; Canitrot, Y.; Calsou, P.; Larminat, F. Polo-like kinase 1 mediates BRCA1 phosphorylation and recruitment at DNA double-strand breaks. Oncotarget 2016, 7, 2269–2283. [Google Scholar] [CrossRef] [Green Version]
  162. Ertych, N.; Stolz, A.; Valerius, O.; Braus, G.H.; Bastians, H. CHK2-BRCA1 tumor-suppressor axis restrains oncogenic Aurora-A kinase to ensure proper mitotic microtubule assembly. Proc. Natl. Acad. Sci. USA 2016, 113, 1817–1822. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Zhang, F.; Yan, P.; Yu, H.; Le, H.; Li, Z.; Chen, J.; Liang, X.; Wang, S.; Wei, W.; Liu, L.; et al. LARP7 Is a BRCA1 Ubiquitinase Substrate and Regulates Genome Stability and Tumorigenesis. Cell Rep. 2020, 32, 108058. [Google Scholar] [CrossRef] [PubMed]
  164. Lin, H.R.; Ting, N.S.; Qin, J.; Lee, W.H. M phase-specific phosphorylation of BRCA2 by Polo-like kinase 1 correlates with the dissociation of the BRCA2-P/CAF complex. J. Biol. Chem. 2003, 278, 35979–35987. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Ehlen, A.; Martin, C.; Miron, S.; Julien, M.; Theillet, F.X.; Ropars, V.; Sessa, G.; Beaurepere, R.; Boucherit, V.; Duchambon, P.; et al. Proper chromosome alignment depends on BRCA2 phosphorylation by PLK1. Nat. Commun. 2020, 11, 1819. [Google Scholar] [CrossRef] [Green Version]
  166. Xia, L.; Wang, X.R.; Wang, X.L.; Liu, S.H.; Ding, X.W.; Chen, G.Q.; Lu, Y. A Novel Role for Pyruvate Kinase M2 as a Corepressor for P53 during the DNA Damage Response in Human Tumor Cells. J. Biol. Chem. 2016, 291, 26138–26150. [Google Scholar] [CrossRef] [Green Version]
  167. Bowen, C.; Ju, J.H.; Lee, J.H.; Paull, T.T.; Gelmann, E.P. Functional activation of ATM by the prostate cancer suppressor NKX3.1. Cell Rep. 2013, 4, 516–529. [Google Scholar] [CrossRef] [Green Version]
  168. Kodama, M.; Otsubo, C.; Hirota, T.; Yokota, J.; Enari, M.; Taya, Y. Requirement of ATM for rapid p53 phosphorylation at Ser46 without Ser/Thr-Gln sequences. Mol. Cell. Biol. 2010, 30, 1620–1633. [Google Scholar] [CrossRef] [Green Version]
  169. Takagi, M.; Tsuchida, R.; Oguchi, K.; Shigeta, T.; Nakada, S.; Shimizu, K.; Ohki, M.; Delia, D.; Chessa, L.; Taya, Y.; et al. Identification and characterization of polymorphic variations of the ataxia telangiectasia mutated (ATM) gene in childhood Hodgkin disease. Blood 2004, 103, 283–290. [Google Scholar] [CrossRef]
  170. Oguchi, K.; Takagi, M.; Tsuchida, R.; Taya, Y.; Ito, E.; Isoyama, K.; Ishii, E.; Zannini, L.; Delia, D.; Mizutani, S. Missense mutation and defective function of ATM in a childhood acute leukemia patient with MLL gene rearrangement. Blood 2003, 101, 3622–3627. [Google Scholar] [CrossRef] [Green Version]
  171. Miyakoda, M.; Suzuki, K.; Kodama, S.; Watanabe, M. Activation of ATM and phosphorylation of p53 by heat shock. Oncogene 2002, 21, 1090–1096. [Google Scholar] [CrossRef]
  172. Ye, R.; Bodero, A.; Zhou, B.B.; Khanna, K.K.; Lavin, M.F.; Lees-Miller, S.P. The plant isoflavenoid genistein activates p53 and Chk2 in an ATM-dependent manner. J. Biol. Chem. 2001, 276, 4828–4833. [Google Scholar] [CrossRef] [Green Version]
  173. Chan, D.W.; Son, S.C.; Block, W.; Ye, R.; Khanna, K.K.; Wold, M.S.; Douglas, P.; Goodarzi, A.A.; Pelley, J.; Taya, Y.; et al. Purification and characterization of ATM from human placenta. A manganese-dependent, wortmannin-sensitive serine/threonine protein kinase. J. Biol. Chem. 2000, 275, 7803–7810. [Google Scholar] [CrossRef] [Green Version]
  174. Hirao, A.; Kong, Y.Y.; Matsuoka, S.; Wakeham, A.; Ruland, J.; Yoshida, H.; Liu, D.; Elledge, S.J.; Mak, T.W. DNA damage-induced activation of p53 by the checkpoint kinase Chk2. Science 2000, 287, 1824–1827. [Google Scholar] [CrossRef]
  175. Tibbetts, R.S.; Brumbaugh, K.M.; Williams, J.M.; Sarkaria, J.N.; Cliby, W.A.; Shieh, S.Y.; Taya, Y.; Prives, C.; Abraham, R.T. A role for ATR in the DNA damage-induced phosphorylation of p53. Genes Dev. 1999, 13, 152–157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Khanna, K.K.; Keating, K.E.; Kozlov, S.; Scott, S.; Gatei, M.; Hobson, K.; Taya, Y.; Gabrielli, B.; Chan, D.; Lees-Miller, S.P.; et al. ATM associates with and phosphorylates p53: Mapping the region of interaction. Nat. Genet. 1998, 20, 398–400. [Google Scholar] [CrossRef] [PubMed]
  177. Canman, C.E.; Lim, D.S.; Cimprich, K.A.; Taya, Y.; Tamai, K.; Sakaguchi, K.; Appella, E.; Kastan, M.B.; Siliciano, J.D. Activation of the ATM kinase by ionizing radiation and phosphorylation of p53. Science 1998, 281, 1677–1679. [Google Scholar] [CrossRef] [PubMed]
  178. Boutell, C.; Everett, R.D. Herpes simplex virus type 1 infection induces the stabilization of p53 in a USP7- and ATM-independent manner. J. Virol. 2004, 78, 8068–8077. [Google Scholar] [CrossRef] [Green Version]
  179. Cui, D.; Xiong, X.; Shu, J.; Dai, X.; Sun, Y.; Zhao, Y. FBXW7 Confers Radiation Survival by Targeting p53 for Degradation. Cell Rep. 2020, 30, 497–509.e4. [Google Scholar] [CrossRef]
  180. Lindsey-Boltz, L.A.; Sercin, O.; Choi, J.H.; Sancar, A. Reconstitution of human claspin-mediated phosphorylation of Chk1 by the ATR (ataxia telangiectasia-mutated and rad3-related) checkpoint kinase. J. Biol. Chem. 2009, 284, 33107–33114. [Google Scholar] [CrossRef] [Green Version]
  181. Hsueh, K.W.; Fu, S.L.; Chang, C.B.; Chang, Y.L.; Lin, C.H. A novel Aurora-A-mediated phosphorylation of p53 inhibits its interaction with MDM2. Biochim. Biophys. Acta 2013, 1834, 508–515. [Google Scholar] [CrossRef]
  182. Liu, Q.; Kaneko, S.; Yang, L.; Feldman, R.I.; Nicosia, S.V.; Chen, J.; Cheng, J.Q. Aurora-A abrogation of p53 DNA binding and transactivation activity by phosphorylation of serine 215. J. Biol. Chem. 2004, 279, 52175–52182. [Google Scholar] [CrossRef] [Green Version]
  183. Choi, B.K.; Dayaram, T.; Parikh, N.; Wilkins, A.D.; Nagarajan, M.; Novikov, I.B.; Bachman, B.J.; Jung, S.Y.; Haas, P.J.; Labrie, J.L.; et al. Literature-based automated discovery of tumor suppressor p53 phosphorylation and inhibition by NEK2. Proc. Natl. Acad. Sci. USA 2018, 115, 10666–10671. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Gully, C.P.; Velazquez-Torres, G.; Shin, J.H.; Fuentes-Mattei, E.; Wang, E.; Carlock, C.; Chen, J.; Rothenberg, D.; Adams, H.P.; Choi, H.H.; et al. Aurora B kinase phosphorylates and instigates degradation of p53. Proc. Natl. Acad. Sci. USA 2012, 109, E1513–E1522. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Wu, L.; Ma, C.A.; Zhao, Y.; Jain, A. Aurora B interacts with NIR-p53, leading to p53 phosphorylation in its DNA-binding domain and subsequent functional suppression. J. Biol. Chem. 2011, 286, 2236–2244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Althubiti, M.; Rada, M.; Samuel, J.; Escorsa, J.M.; Najeeb, H.; Lee, K.G.; Lam, K.P.; Jones, G.D.; Barlev, N.A.; Macip, S. BTK Modulates p53 Activity to Enhance Apoptotic and Senescent Responses. Cancer Res. 2016, 76, 5405–5414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Nantajit, D.; Fan, M.; Duru, N.; Wen, Y.; Reed, J.C.; Li, J.J. Cyclin B1/Cdk1 phosphorylation of mitochondrial p53 induces anti-apoptotic response. PLoS ONE 2010, 5, e12341. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Blaydes, J.P.; Luciani, M.G.; Pospisilova, S.; Ball, H.M.; Vojtesek, B.; Hupp, T.R. Stoichiometric phosphorylation of human p53 at Ser315 stimulates p53-dependent transcription. J. Biol. Chem. 2001, 276, 4699–4708. [Google Scholar] [CrossRef] [Green Version]
  189. Wang, Y.; Prives, C. Increased and altered DNA binding of human p53 by S and G2/M but not G1 cyclin-dependent kinases. Nature 1995, 376, 88–91. [Google Scholar] [CrossRef]
  190. Liao, P.; Zeng, S.X.; Zhou, X.; Chen, T.; Zhou, F.; Cao, B.; Jung, J.H.; Del Sal, G.; Luo, S.; Lu, H. Mutant p53 Gains Its Function via c-Myc Activation upon CDK4 Phosphorylation at Serine 249 and Consequent PIN1 Binding. Mol. Cell 2017, 68, 1134–1146.e6. [Google Scholar] [CrossRef] [Green Version]
  191. Ajay, A.K.; Upadhyay, A.K.; Singh, S.; Vijayakumar, M.V.; Kumari, R.; Pandey, V.; Boppana, R.; Bhat, M.K. Cdk5 phosphorylates non-genotoxically overexpressed p53 following inhibition of PP2A to induce cell cycle arrest/apoptosis and inhibits tumor progression. Mol. Cancer 2010, 9, 204. [Google Scholar] [CrossRef] [Green Version]
  192. Wang, H.; Chen, L.; Zhou, T.; Zhang, Z.; Zeng, C. Nicotine Promotes WRL68 Cells Proliferation Due to the Mutant p53 Gain-of-Function by Activating CDK6-p53-RS-PIN1-STAT1 Signaling Pathway. Chem. Res. Toxicol. 2020, 33, 2361–2373. [Google Scholar] [CrossRef] [PubMed]
  193. Wang, H.; Liao, P.; Zeng, S.X.; Lu, H. Co-targeting p53-R249S and CDK4 synergistically suppresses survival of hepatocellular carcinoma cells. Cancer Biol. Ther. 2020, 21, 269–277. [Google Scholar] [CrossRef] [PubMed]
  194. Radhakrishnan, S.K.; Gartel, A.L. CDK9 phosphorylates p53 on serine residues 33, 315 and 392. Cell Cycle 2006, 5, 519–521. [Google Scholar] [CrossRef] [PubMed]
  195. Goudelock, D.M.; Jiang, K.; Pereira, E.; Russell, B.; Sanchez, Y. Regulatory interactions between the checkpoint kinase Chk1 and the proteins of the DNA-dependent protein kinase complex. J. Biol. Chem. 2003, 278, 29940–29947. [Google Scholar] [CrossRef] [Green Version]
  196. Shieh, S.Y.; Ahn, J.; Tamai, K.; Taya, Y.; Prives, C. The human homologs of checkpoint kinases Chk1 and Cds1 (Chk2) phosphorylate p53 at multiple DNA damage-inducible sites. Genes Dev. 2000, 14, 289–300. [Google Scholar] [CrossRef]
  197. Craig, A.L.; Chrystal, J.A.; Fraser, J.A.; Sphyris, N.; Lin, Y.; Harrison, B.J.; Scott, M.T.; Dornreiter, I.; Hupp, T.R. The MDM2 ubiquitination signal in the DNA-binding domain of p53 forms a docking site for calcium calmodulin kinase superfamily members. Mol. Cell. Biol. 2007, 27, 3542–3555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Ou, Y.H.; Chung, P.H.; Sun, T.P.; Shieh, S.Y. p53 C-terminal phosphorylation by CHK1 and CHK2 participates in the regulation of DNA-damage-induced C-terminal acetylation. Mol. Biol. Cell 2005, 16, 1684–1695. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Chehab, N.H.; Malikzay, A.; Appel, M.; Halazonetis, T.D. Chk2/hCds1 functions as a DNA damage checkpoint in G(1) by stabilizing p53. Genes Dev. 2000, 14, 278–288. [Google Scholar] [CrossRef] [PubMed]
  200. Venerando, A.; Marin, O.; Cozza, G.; Bustos, V.H.; Sarno, S.; Pinna, L.A. Isoform specific phosphorylation of p53 by protein kinase CK1. Cell. Mol. Life Sci. 2010, 67, 1105–1118. [Google Scholar] [CrossRef]
  201. MacLaine, N.J.; Oster, B.; Bundgaard, B.; Fraser, J.A.; Buckner, C.; Lazo, P.A.; Meek, D.W.; Hollsberg, P.; Hupp, T.R. A central role for CK1 in catalyzing phosphorylation of the p53 transactivation domain at serine 20 after HHV-6B viral infection. J. Biol. Chem. 2008, 283, 28563–28573. [Google Scholar] [CrossRef] [Green Version]
  202. Gasparian, A.V.; Burkhart, C.A.; Purmal, A.A.; Brodsky, L.; Pal, M.; Saranadasa, M.; Bosykh, D.A.; Commane, M.; Guryanova, O.A.; Pal, S.; et al. Curaxins: Anticancer compounds that simultaneously suppress NF-kappaB and activate p53 by targeting FACT. Sci. Transl. Med. 2011, 3, 95ra74. [Google Scholar] [CrossRef] [PubMed]
  203. Cruet-Hennequart, S.; Glynn, M.T.; Murillo, L.S.; Coyne, S.; Carty, M.P. Enhanced DNA-PK-mediated RPA2 hyperphosphorylation in DNA polymerase eta-deficient human cells treated with cisplatin and oxaliplatin. DNA Repair 2008, 7, 582–596. [Google Scholar] [CrossRef] [PubMed]
  204. Li, D.W.; Liu, J.P.; Schmid, P.C.; Schlosser, R.; Feng, H.; Liu, W.B.; Yan, Q.; Gong, L.; Sun, S.M.; Deng, M.; et al. Protein serine/threonine phosphatase-1 dephosphorylates p53 at Ser-15 and Ser-37 to modulate its transcriptional and apoptotic activities. Oncogene 2006, 25, 3006–3022. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Levy, N.; Martz, A.; Bresson, A.; Spenlehauer, C.; de Murcia, G.; Menissier-de Murcia, J. XRCC1 is phosphorylated by DNA-dependent protein kinase in response to DNA damage. Nucleic Acids Res. 2006, 34, 32–41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Komiyama, S.; Taniguchi, S.; Matsumoto, Y.; Tsunoda, E.; Ohto, T.; Suzuki, Y.; Yin, H.L.; Tomita, M.; Enomoto, A.; Morita, A.; et al. Potentiality of DNA-dependent protein kinase to phosphorylate Ser46 of human p53. Biochem. Biophys. Res. Commun. 2004, 323, 816–822. [Google Scholar] [CrossRef] [PubMed]
  207. Soubeyrand, S.; Schild-Poulter, C.; Hache, R.J. Structured DNA promotes phosphorylation of p53 by DNA-dependent protein kinase at serine 9 and threonine 18. Eur. J. Biochem. 2004, 271, 3776–3784. [Google Scholar] [CrossRef]
  208. Sakaguchi, K.; Saito, S.; Higashimoto, Y.; Roy, S.; Anderson, C.W.; Appella, E. Damage-mediated phosphorylation of human p53 threonine 18 through a cascade mediated by a casein 1-like kinase. Effect on Mdm2 binding. J. Biol. Chem. 2000, 275, 9278–9283. [Google Scholar] [CrossRef] [Green Version]
  209. Kapoor, M.; Hamm, R.; Yan, W.; Taya, Y.; Lozano, G. Cooperative phosphorylation at multiple sites is required to activate p53 in response to UV radiation. Oncogene 2000, 19, 358–364. [Google Scholar] [CrossRef] [Green Version]
  210. Dumaz, N.; Milne, D.M.; Meek, D.W. Protein kinase CK1 is a p53-threonine 18 kinase which requires prior phosphorylation of serine 15. FEBS Lett. 1999, 463, 312–316. [Google Scholar] [CrossRef] [Green Version]
  211. Shieh, S.Y.; Ikeda, M.; Taya, Y.; Prives, C. DNA damage-induced phosphorylation of p53 alleviates inhibition by MDM2. Cell 1997, 91, 325–334. [Google Scholar] [CrossRef] [Green Version]
  212. Lees-Miller, S.P.; Sakaguchi, K.; Ullrich, S.J.; Appella, E.; Anderson, C.W. Human DNA-activated protein kinase phosphorylates serines 15 and 37 in the amino-terminal transactivation domain of human p53. Mol. Cell. Biol. 1992, 12, 5041–5049. [Google Scholar] [PubMed] [Green Version]
  213. Park, J.; Oh, Y.; Yoo, L.; Jung, M.S.; Song, W.J.; Lee, S.H.; Seo, H.; Chung, K.C. Dyrk1A phosphorylates p53 and inhibits proliferation of embryonic neuronal cells. J. Biol. Chem. 2010, 285, 31895–31906. [Google Scholar] [CrossRef] [Green Version]
  214. Taira, N.; Nihira, K.; Yamaguchi, T.; Miki, Y.; Yoshida, K. DYRK2 is targeted to the nucleus and controls p53 via Ser46 phosphorylation in the apoptotic response to DNA damage. Mol. Cell 2007, 25, 725–738. [Google Scholar] [CrossRef]
  215. Persons, D.L.; Yazlovitskaya, E.M.; Pelling, J.C. Effect of extracellular signal-regulated kinase on p53 accumulation in response to cisplatin. J. Biol. Chem. 2000, 275, 35778–35785. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Melnikova, V.O.; Santamaria, A.B.; Bolshakov, S.V.; Ananthaswamy, H.N. Mutant p53 is constitutively phosphorylated at Serine 15 in UV-induced mouse skin tumors: Involvement of ERK1/2 MAP kinase. Oncogene 2003, 22, 5958–5966. [Google Scholar] [CrossRef] [Green Version]
  217. She, Q.B.; Ma, W.Y.; Dong, Z. Role of MAP kinases in UVB-induced phosphorylation of p53 at serine 20. Oncogene 2002, 21, 1580–1589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Yeh, P.Y.; Chuang, S.E.; Yeh, K.H.; Song, Y.C.; Cheng, A.L. Nuclear extracellular signal-regulated kinase 2 phosphorylates p53 at Thr55 in response to doxorubicin. Biochem. Biophys. Res. Commun. 2001, 284, 880–886. [Google Scholar] [CrossRef]
  219. Gambardella, J.; Fiordelisi, A.; Santulli, G.; Ciccarelli, M.; Cerasuolo, F.A.; Sala, M.; Sommella, E.; Campiglia, P.; Illario, M.; Iaccarino, G.; et al. Exploiting GRK2 Inhibition as a Therapeutic Option in Experimental Cancer Treatment: Role of p53-Induced Mitochondrial Apoptosis. Cancers 2020, 12, 3530. [Google Scholar] [CrossRef]
  220. Chen, X.; Zhu, H.; Yuan, M.; Fu, J.; Zhou, Y.; Ma, L. G-protein-coupled receptor kinase 5 phosphorylates p53 and inhibits DNA damage-induced apoptosis. J. Biol. Chem. 2010, 285, 12823–12830. [Google Scholar] [CrossRef] [Green Version]
  221. D’Orazi, G.; Cecchinelli, B.; Bruno, T.; Manni, I.; Higashimoto, Y.; Saito, S.; Gostissa, M.; Coen, S.; Marchetti, A.; Del Sal, G.; et al. Homeodomain-interacting protein kinase-2 phosphorylates p53 at Ser 46 and mediates apoptosis. Nat. Cell Biol. 2002, 4, 11–19. [Google Scholar] [CrossRef]
  222. Hofmann, T.G.; Moller, A.; Sirma, H.; Zentgraf, H.; Taya, Y.; Droge, W.; Will, H.; Schmitz, M.L. Regulation of p53 activity by its interaction with homeodomain-interacting protein kinase-2. Nat. Cell Biol. 2002, 4, 1–10. [Google Scholar] [CrossRef]
  223. Buschmann, T.; Potapova, O.; Bar-Shira, A.; Ivanov, V.N.; Fuchs, S.Y.; Henderson, S.; Fried, V.A.; Minamoto, T.; Alarcon-Vargas, D.; Pincus, M.R.; et al. Jun NH2-terminal kinase phosphorylation of p53 on Thr-81 is important for p53 stabilization and transcriptional activities in response to stress. Mol. Cell. Biol. 2001, 21, 2743–2754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Zeng, P.Y.; Berger, S.L. LKB1 is recruited to the p21/WAF1 promoter by p53 to mediate transcriptional activation. Cancer Res. 2006, 66, 10701–10708. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Woods, M.L.; Weiss, A.; Sokol, A.M.; Graumann, J.; Boettger, T.; Richter, A.M.; Schermuly, R.T.; Dammann, R.H. Epigenetically silenced apoptosis-associated tyrosine kinase (AATK) facilitates a decreased expression of Cyclin D1 and WEE1, phosphorylates TP53 and reduces cell proliferation in a kinase-dependent manner. Cancer Gene Ther. 2022, 29, 1975–1987. [Google Scholar] [CrossRef] [PubMed]
  226. Ho, D.H.; Kim, H.; Kim, J.; Sim, H.; Ahn, H.; Kim, J.; Seo, H.; Chung, K.C.; Park, B.J.; Son, I.; et al. Leucine-Rich Repeat Kinase 2 (LRRK2) phosphorylates p53 and induces p21(WAF1/CIP1) expression. Mol. Brain 2015, 8, 54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Lee, C.W.; Ferreon, J.C.; Ferreon, A.C.; Arai, M.; Wright, P.E. Graded enhancement of p53 binding to CREB-binding protein (CBP) by multisite phosphorylation. Proc. Natl. Acad. Sci. USA 2010, 107, 19290–19295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Sun, P.; Yoshizuka, N.; New, L.; Moser, B.A.; Li, Y.; Liao, R.; Xie, C.; Chen, J.; Deng, Q.; Yamout, M.; et al. PRAK is essential for ras-induced senescence and tumor suppression. Cell 2007, 128, 295–308. [Google Scholar] [CrossRef]
  229. Hou, X.; Liu, J.E.; Liu, W.; Liu, C.Y.; Liu, Z.Y.; Sun, Z.Y. A new role of NUAK1: Directly phosphorylating p53 and regulating cell proliferation. Oncogene 2011, 30, 2933–2942. [Google Scholar] [CrossRef]
  230. Wen, J.; Chen, W.; Zhao, B.; Xu, Q.; Liu, C.; Zhang, Q.; Xie, Z.; Yan, Y.; Guo, J.; Huang, J.; et al. A carbazole compound, 9-ethyl-9H-carbazole-3-carbaldehyde, plays an antitumor function through reactivation of the p53 pathway in human melanoma cells. Cell Death Dis. 2021, 12, 591. [Google Scholar] [CrossRef]
  231. Kim, H.D.; Kim, T.S.; Kim, J. Aberrant ribosome biogenesis activates c-Myc and ASK1 pathways resulting in p53-dependent G1 arrest. Oncogene 2011, 30, 3317–3327. [Google Scholar] [CrossRef] [Green Version]
  232. Kim, S.J.; Hwang, S.G.; Shin, D.Y.; Kang, S.S.; Chun, J.S. p38 kinase regulates nitric oxide-induced apoptosis of articular chondrocytes by accumulating p53 via NFkappa B-dependent transcription and stabilization by serine 15 phosphorylation. J. Biol. Chem. 2002, 277, 33501–33508. [Google Scholar] [CrossRef] [Green Version]
  233. Kwong, J.; Hong, L.; Liao, R.; Deng, Q.; Han, J.; Sun, P. p38alpha and p38gamma mediate oncogenic ras-induced senescence through differential mechanisms. J. Biol. Chem. 2009, 284, 11237–11246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Bulavin, D.V.; Saito, S.; Hollander, M.C.; Sakaguchi, K.; Anderson, C.W.; Appella, E.; Fornace, A.J., Jr. Phosphorylation of human p53 by p38 kinase coordinates N-terminal phosphorylation and apoptosis in response to UV radiation. EMBO J. 1999, 18, 6845–6854. [Google Scholar] [CrossRef] [PubMed]
  235. Xu, H.T.; Lai, W.L.; Liu, H.F.; Wong, L.L.; Ng, I.O.; Ching, Y.P. PAK4 Phosphorylates p53 at Serine 215 to Promote Liver Cancer Metastasis. Cancer Res. 2016, 76, 5732–5742. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Yoshida, K.; Liu, H.; Miki, Y. Protein kinase C delta regulates Ser46 phosphorylation of p53 tumor suppressor in the apoptotic response to DNA damage. J. Biol. Chem. 2006, 281, 5734–5740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Cuddihy, A.R.; Wong, A.H.; Tam, N.W.; Li, S.; Koromilas, A.E. The double-stranded RNA activated protein kinase PKR physically associates with the tumor suppressor p53 protein and phosphorylates human p53 on serine 392 in vitro. Oncogene 1999, 18, 2690–2702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Li, Z.; Niu, J.; Uwagawa, T.; Peng, B.; Chiao, P.J. Function of polo-like kinase 3 in NF-kappaB-mediated proapoptotic response. J. Biol. Chem. 2005, 280, 16843–16850. [Google Scholar] [CrossRef] [Green Version]
  239. Xie, S.; Wu, H.; Wang, Q.; Cogswell, J.P.; Husain, I.; Conn, C.; Stambrook, P.; Jhanwar-Uniyal, M.; Dai, W. Plk3 functionally links DNA damage to cell cycle arrest and apoptosis at least in part via the p53 pathway. J. Biol. Chem. 2001, 276, 43305–43312. [Google Scholar] [CrossRef] [Green Version]
  240. Xie, S.; Wang, Q.; Wu, H.; Cogswell, J.; Lu, L.; Jhanwar-Uniyal, M.; Dai, W. Reactive oxygen species-induced phosphorylation of p53 on serine 20 is mediated in part by polo-like kinase-3. J. Biol. Chem. 2001, 276, 36194–36199. [Google Scholar] [CrossRef] [Green Version]
  241. Facchin, S.; Ruzzene, M.; Peggion, C.; Sartori, G.; Carignani, G.; Marin, O.; Brustolon, F.; Lopreiato, R.; Pinna, L.A. Phosphorylation and activation of the atypical kinase p53-related protein kinase (PRPK) by Akt/PKB. Cell. Mol. Life Sci. 2007, 64, 2680–2689. [Google Scholar] [CrossRef]
  242. Abe, Y.; Matsumoto, S.; Wei, S.; Nezu, K.; Miyoshi, A.; Kito, K.; Ueda, N.; Shigemoto, K.; Hitsumoto, Y.; Nikawa, J.; et al. Cloning and characterization of a p53-related protein kinase expressed in interleukin-2-activated cytotoxic T-cells, epithelial tumor cell lines, and the testes. J. Biol. Chem. 2001, 276, 44003–44011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Huang, Y.F.; Bulavin, D.V. Oncogene-mediated regulation of p53 ISGylation and functions. Oncotarget 2014, 5, 5808–5818. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Wu, Y.; Lee, S.; Bobadilla, S.; Duan, S.Z.; Liu, X. High glucose-induced p53 phosphorylation contributes to impairment of endothelial antioxidant system. Biochim. Biophys. Acta Mol. Basis Dis. 2017, 1863, 2355–2362. [Google Scholar] [CrossRef]
  245. Wu, Y.; Lin, J.C.; Piluso, L.G.; Dhahbi, J.M.; Bobadilla, S.; Spindler, S.R.; Liu, X. Phosphorylation of p53 by TAF1 inactivates p53-dependent transcription in the DNA damage response. Mol. Cell 2014, 53, 63–74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Li, H.H.; Li, A.G.; Sheppard, H.M.; Liu, X. Phosphorylation on Thr-55 by TAF1 mediates degradation of p53: A role for TAF1 in cell G1 progression. Mol. Cell 2004, 13, 867–878. [Google Scholar] [CrossRef]
  247. Vega, F.M.; Sevilla, A.; Lazo, P.A. p53 Stabilization and accumulation induced by human vaccinia-related kinase 1. Mol. Cell. Biol. 2004, 24, 10366–10380. [Google Scholar] [CrossRef] [Green Version]
  248. Li, L.; Ljungman, M.; Dixon, J.E. The human Cdc14 phosphatases interact with and dephosphorylate the tumor suppressor protein p53. J. Biol. Chem. 2000, 275, 2410–2414. [Google Scholar] [CrossRef] [Green Version]
  249. Chen, Y.C.; Hsieh, H.H.; Chang, H.C.; Wang, H.C.; Lin, W.J.; Lin, J.J. CDC25B induces cellular senescence and correlates with tumor suppression in a p53-dependent manner. J. Biol. Chem. 2021, 296, 100564. [Google Scholar] [CrossRef]
  250. Shang, X.; Vasudevan, S.A.; Yu, Y.; Ge, N.; Ludwig, A.D.; Wesson, C.L.; Wang, K.; Burlingame, S.M.; Zhao, Y.J.; Rao, P.H.; et al. Dual-specificity phosphatase 26 is a novel p53 phosphatase and inhibits p53 tumor suppressor functions in human neuroblastoma. Oncogene 2010, 29, 4938–4946. [Google Scholar] [CrossRef] [Green Version]
  251. Ichwan, S.J.; Yamada, S.; Sumrejkanchanakij, P.; Ibrahim-Auerkari, E.; Eto, K.; Ikeda, M.A. Defect in serine 46 phosphorylation of p53 contributes to acquisition of p53 resistance in oral squamous cell carcinoma cells. Oncogene 2006, 25, 1216–1224. [Google Scholar] [CrossRef] [Green Version]
  252. Dohoney, K.M.; Guillerm, C.; Whiteford, C.; Elbi, C.; Lambert, P.F.; Hager, G.L.; Brady, J.N. Phosphorylation of p53 at serine 37 is important for transcriptional activity and regulation in response to DNA damage. Oncogene 2004, 23, 49–57. [Google Scholar] [CrossRef] [Green Version]
  253. Park, J.H.; Smith, R.J.; Shieh, S.Y.; Roeder, R.G. The GAS41-PP2Cbeta complex dephosphorylates p53 at serine 366 and regulates its stability. J. Biol. Chem. 2011, 286, 10911–10917. [Google Scholar] [CrossRef] [Green Version]
  254. Li, H.H.; Cai, X.; Shouse, G.P.; Piluso, L.G.; Liu, X. A specific PP2A regulatory subunit, B56gamma, mediates DNA damage-induced dephosphorylation of p53 at Thr55. EMBO J. 2007, 26, 402–411. [Google Scholar] [CrossRef]
  255. Dias, S.S.; Milne, D.M.; Meek, D.W. c-Abl phosphorylates Hdm2 at tyrosine 276 in response to DNA damage and regulates interaction with ARF. Oncogene 2006, 25, 6666–6671. [Google Scholar] [CrossRef] [Green Version]
  256. Goldberg, Z.; Vogt Sionov, R.; Berger, M.; Zwang, Y.; Perets, R.; Van Etten, R.A.; Oren, M.; Taya, Y.; Haupt, Y. Tyrosine phosphorylation of Mdm2 by c-Abl: Implications for p53 regulation. EMBO J. 2002, 21, 3715–3727. [Google Scholar] [CrossRef]
  257. Milne, D.; Kampanis, P.; Nicol, S.; Dias, S.; Campbell, D.G.; Fuller-Pace, F.; Meek, D. A novel site of AKT-mediated phosphorylation in the human MDM2 onco-protein. FEBS Lett. 2004, 577, 270–276. [Google Scholar] [CrossRef] [Green Version]
  258. Feng, J.; Tamaskovic, R.; Yang, Z.; Brazil, D.P.; Merlo, A.; Hess, D.; Hemmings, B.A. Stabilization of Mdm2 via decreased ubiquitination is mediated by protein kinase B/Akt-dependent phosphorylation. J. Biol. Chem. 2004, 279, 35510–35517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  259. Ashcroft, M.; Ludwig, R.L.; Woods, D.B.; Copeland, T.D.; Weber, H.O.; MacRae, E.J.; Vousden, K.H. Phosphorylation of HDM2 by Akt. Oncogene 2002, 21, 1955–1962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  260. Zhou, B.P.; Liao, Y.; Xia, W.; Zou, Y.; Spohn, B.; Hung, M.C. HER-2/neu induces p53 ubiquitination via Akt-mediated MDM2 phosphorylation. Nat. Cell Biol. 2001, 3, 973–982. [Google Scholar] [CrossRef] [PubMed]
  261. Mayo, L.D.; Donner, D.B. A phosphatidylinositol 3-kinase/Akt pathway promotes translocation of Mdm2 from the cytoplasm to the nucleus. Proc. Natl. Acad. Sci. USA 2001, 98, 11598–11603. [Google Scholar] [CrossRef] [Green Version]
  262. Ogawara, Y.; Kishishita, S.; Obata, T.; Isazawa, Y.; Suzuki, T.; Tanaka, K.; Masuyama, N.; Gotoh, Y. Akt enhances Mdm2-mediated ubiquitination and degradation of p53. J. Biol. Chem. 2002, 277, 21843–21850. [Google Scholar] [CrossRef] [Green Version]
  263. Maya, R.; Balass, M.; Kim, S.T.; Shkedy, D.; Leal, J.F.; Shifman, O.; Moas, M.; Buschmann, T.; Ronai, Z.; Shiloh, Y.; et al. ATM-dependent phosphorylation of Mdm2 on serine 395: Role in p53 activation by DNA damage. Genes Dev. 2001, 15, 1067–1077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Cheng, Q.; Chen, L.; Li, Z.; Lane, W.S.; Chen, J. ATM activates p53 by regulating MDM2 oligomerization and E3 processivity. EMBO J. 2009, 28, 3857–3867. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Cheng, Q.; Cross, B.; Li, B.; Chen, L.; Li, Z.; Chen, J. Regulation of MDM2 E3 ligase activity by phosphorylation after DNA damage. Mol. Cell. Biol. 2011, 31, 4951–4963. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  266. Shinozaki, T.; Nota, A.; Taya, Y.; Okamoto, K. Functional role of Mdm2 phosphorylation by ATR in attenuation of p53 nuclear export. Oncogene 2003, 22, 8870–8880. [Google Scholar] [CrossRef] [Green Version]
  267. Yao, J.Y.; Xu, S.; Sun, Y.N.; Xu, Y.; Guo, Q.L.; Wei, L.B. Novel CDK9 inhibitor oroxylin A promotes wild-type P53 stability and prevents hepatocellular carcinoma progression by disrupting both MDM2 and SIRT1 signaling. Acta Pharmacol. Sin. 2022, 43, 1033–1045. [Google Scholar] [CrossRef]
  268. Inuzuka, H.; Tseng, A.; Gao, D.; Zhai, B.; Zhang, Q.; Shaik, S.; Wan, L.; Ang, X.L.; Mock, C.; Yin, H.; et al. Phosphorylation by casein kinase I promotes the turnover of the Mdm2 oncoprotein via the SCF(beta-TRCP) ubiquitin ligase. Cancer Cell 2010, 18, 147–159. [Google Scholar] [CrossRef] [Green Version]
  269. Zhang, S.; Zhou, L.; El-Deiry, W.S. Small-Molecule NSC59984 Induces Mutant p53 Degradation through a ROS-ERK2-MDM2 Axis in Cancer Cells. Mol. Cancer Res. 2022, 20, 622–636. [Google Scholar] [CrossRef]
  270. Weber, H.O.; Ludwig, R.L.; Morrison, D.; Kotlyarov, A.; Gaestel, M.; Vousden, K.H. HDM2 phosphorylation by MAPKAP kinase 2. Oncogene 2005, 24, 1965–1972. [Google Scholar] [CrossRef] [Green Version]
  271. Liu, T.; Li, Y.; Gu, H.; Zhu, G.; Li, J.; Cao, L.; Li, F. p21-Activated kinase 6 (PAK6) inhibits prostate cancer growth via phosphorylation of androgen receptor and tumorigenic E3 ligase murine double minute-2 (Mdm2). J. Biol. Chem. 2013, 288, 3359–3369. [Google Scholar] [CrossRef] [Green Version]
  272. Hogan, C.; Hutchison, C.; Marcar, L.; Milne, D.; Saville, M.; Goodlad, J.; Kernohan, N.; Meek, D. Elevated levels of oncogenic protein kinase Pim-1 induce the p53 pathway in cultured cells and correlate with increased Mdm2 in mantle cell lymphoma. J. Biol. Chem. 2008, 283, 18012–18023. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Dias, S.S.; Hogan, C.; Ochocka, A.M.; Meek, D.W. Polo-like kinase-1 phosphorylates MDM2 at Ser260 and stimulates MDM2-mediated p53 turnover. FEBS Lett. 2009, 583, 3543–3548. [Google Scholar] [CrossRef] [Green Version]
  274. Zheng, Z.; Li, L.; Liu, X.; Wang, D.; Tu, B.; Wang, L.; Wang, H.; Zhu, W.G. 5-Aza-2′-deoxycytidine reactivates gene expression via degradation of pRb pocket proteins. FASEB J. 2012, 26, 449–459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  275. Okamoto, K.; Li, H.; Jensen, M.R.; Zhang, T.; Taya, Y.; Thorgeirsson, S.S.; Prives, C. Cyclin G recruits PP2A to dephosphorylate Mdm2. Mol. Cell 2002, 9, 761–771. [Google Scholar] [CrossRef] [PubMed]
  276. Yang, X.; Chen, G.; Li, W.; Peng, C.; Zhu, Y.; Yang, X.; Li, T.; Cao, C.; Pei, H. Cervical Cancer Growth Is Regulated by a c-ABL-PLK1 Signaling Axis. Cancer Res. 2017, 77, 1142–1154. [Google Scholar] [CrossRef] [Green Version]
  277. Raab, M.; Matthess, Y.; Raab, C.A.; Gutfreund, N.; Dotsch, V.; Becker, S.; Sanhaji, M.; Strebhardt, K. A dimerization-dependent mechanism regulates enzymatic activation and nuclear entry of PLK1. Oncogene 2022, 41, 372–386. [Google Scholar] [CrossRef]
  278. Tavernier, N.; Thomas, Y.; Vigneron, S.; Maisonneuve, P.; Orlicky, S.; Mader, P.; Regmi, S.G.; Van Hove, L.; Levinson, N.M.; Gasmi-Seabrook, G.; et al. Bora phosphorylation substitutes in trans for T-loop phosphorylation in Aurora A to promote mitotic entry. Nat. Commun. 2021, 12, 1899. [Google Scholar] [CrossRef]
  279. Li, W.; Wang, H.Y.; Zhao, X.; Duan, H.; Cheng, B.; Liu, Y.; Zhao, M.; Shu, W.; Mei, Y.; Wen, Z.; et al. A methylation-phosphorylation switch determines Plk1 kinase activity and function in DNA damage repair. Sci. Adv. 2019, 5, eaau7566. [Google Scholar] [CrossRef] [Green Version]
  280. Qi, F.; Chen, Q.; Chen, H.; Yan, H.; Chen, B.; Xiang, X.; Liang, C.; Yi, Q.; Zhang, M.; Cheng, H.; et al. WAC Promotes Polo-like Kinase 1 Activation for Timely Mitotic Entry. Cell Rep. 2018, 24, 546–556. [Google Scholar] [CrossRef] [Green Version]
  281. Macurek, L.; Lindqvist, A.; Lim, D.; Lampson, M.A.; Klompmaker, R.; Freire, R.; Clouin, C.; Taylor, S.S.; Yaffe, M.B.; Medema, R.H. Polo-like kinase-1 is activated by aurora A to promote checkpoint recovery. Nature 2008, 455, 119–123. [Google Scholar] [CrossRef]
  282. Peng, B.; Shi, R.; Bian, J.; Li, Y.; Wang, P.; Wang, H.; Liao, J.; Zhu, W.G.; Xu, X. PARP1 and CHK1 coordinate PLK1 enzymatic activity during the DNA damage response to promote homologous recombination-mediated repair. Nucleic Acids Res. 2021, 49, 7554–7570. [Google Scholar] [CrossRef] [PubMed]
  283. Wu, B.; Jiang, P.; Mu, Y.; Wilmouth, R.C. Cancer Osaka thyroid (Cot) phosphorylates Polo-like kinase (PLK1) at Ser137 but not at Thr210. Biol. Chem. 2009, 390, 1271–1277. [Google Scholar] [CrossRef] [PubMed]
  284. Tang, J.; Yang, X.; Liu, X. Phosphorylation of Plk1 at Ser326 regulates its functions during mitotic progression. Oncogene 2008, 27, 6635–6645. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  285. Maroto, B.; Ye, M.B.; von Lohneysen, K.; Schnelzer, A.; Knaus, U.G. P21-activated kinase is required for mitotic progression and regulates Plk1. Oncogene 2008, 27, 4900–4908. [Google Scholar] [CrossRef] [Green Version]
  286. Yu, L.; Lang, Y.; Hsu, C.C.; Chen, W.M.; Chiang, J.C.; Hsieh, J.T.; Story, M.D.; Shang, Z.F.; Chen, B.P.C.; Saha, D. Mitotic phosphorylation of tumor suppressor DAB2IP maintains spindle assembly checkpoint and chromosomal stability through activating PLK1-Mps1 signal pathway and stabilizing mitotic checkpoint complex. Oncogene 2022, 41, 489–501. [Google Scholar] [CrossRef]
  287. Matsumoto, T.; Wang, P.Y.; Ma, W.; Sung, H.J.; Matoba, S.; Hwang, P.M. Polo-like kinases mediate cell survival in mitochondrial dysfunction. Proc. Natl. Acad. Sci. USA 2009, 106, 14542–14546. [Google Scholar] [CrossRef] [Green Version]
  288. Zhu, H.; Li, Q.; Zhao, Y.; Peng, H.; Guo, L.; Zhu, J.; Jiang, Z.; Zeng, Z.; Xu, B.; Chen, S. Vaccinia-related kinase 2 drives pancreatic cancer progression by protecting Plk1 from Chfr-mediated degradation. Oncogene 2021, 40, 4663–4674. [Google Scholar] [CrossRef]
  289. Xia, D.; Stull, J.T.; Kamm, K.E. Myosin phosphatase targeting subunit 1 affects cell migration by regulating myosin phosphorylation and actin assembly. Exp. Cell Res. 2005, 304, 506–517. [Google Scholar] [CrossRef]
  290. Katayama, K.; Fujita, N.; Tsuruo, T. Akt/protein kinase B-dependent phosphorylation and inactivation of WEE1Hu promote cell cycle progression at G2/M transition. Mol. Cell. Biol. 2005, 25, 5725–5737. [Google Scholar] [CrossRef] [Green Version]
  291. Ovejero, S.; Ayala, P.; Bueno, A.; Sacristan, M.P. Human Cdc14A regulates Wee1 stability by counteracting CDK-mediated phosphorylation. Mol. Biol. Cell 2012, 23, 4515–4525. [Google Scholar] [CrossRef]
  292. Mazzolini, L.; Broban, A.; Froment, C.; Burlet-Schiltz, O.; Besson, A.; Manenti, S.; Dozier, C. Phosphorylation of CDC25A on SER283 in late S/G2 by CDK/cyclin complexes accelerates mitotic entry. Cell Cycle 2016, 15, 2742–2752. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Mailand, N.; Podtelejnikov, A.V.; Groth, A.; Mann, M.; Bartek, J.; Lukas, J. Regulation of G(2)/M events by Cdc25A through phosphorylation-dependent modulation of its stability. EMBO J. 2002, 21, 5911–5920. [Google Scholar] [CrossRef] [PubMed]
  294. Liang, J.; Cao, R.; Zhang, Y.; Xia, Y.; Zheng, Y.; Li, X.; Wang, L.; Yang, W.; Lu, Z. PKM2 dephosphorylation by Cdc25A promotes the Warburg effect and tumorigenesis. Nat. Commun. 2016, 7, 12431. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  295. Kasahara, K.; Goto, H.; Enomoto, M.; Tomono, Y.; Kiyono, T.; Inagaki, M. 14-3-3gamma mediates Cdc25A proteolysis to block premature mitotic entry after DNA damage. EMBO J. 2010, 29, 2802–2812. [Google Scholar] [CrossRef] [Green Version]
  296. Jin, J.; Ang, X.L.; Ye, X.; Livingstone, M.; Harper, J.W. Differential roles for checkpoint kinases in DNA damage-dependent degradation of the Cdc25A protein phosphatase. J. Biol. Chem. 2008, 283, 19322–19328. [Google Scholar] [CrossRef] [Green Version]
  297. Jin, J.; Shirogane, T.; Xu, L.; Nalepa, G.; Qin, J.; Elledge, S.J.; Harper, J.W. SCFbeta-TRCP links Chk1 signaling to degradation of the Cdc25A protein phosphatase. Genes Dev. 2003, 17, 3062–3074. [Google Scholar] [CrossRef] [Green Version]
  298. Sorensen, C.S.; Syljuasen, R.G.; Falck, J.; Schroeder, T.; Ronnstrand, L.; Khanna, K.K.; Zhou, B.B.; Bartek, J.; Lukas, J. Chk1 regulates the S phase checkpoint by coupling the physiological turnover and ionizing radiation-induced accelerated proteolysis of Cdc25A. Cancer Cell 2003, 3, 247–258. [Google Scholar] [CrossRef] [Green Version]
  299. Chen, M.S.; Ryan, C.E.; Piwnica-Worms, H. Chk1 kinase negatively regulates mitotic function of Cdc25A phosphatase through 14-3-3 binding. Mol. Cell. Biol. 2003, 23, 7488–7497. [Google Scholar] [CrossRef] [Green Version]
  300. Falck, J.; Mailand, N.; Syljuasen, R.G.; Bartek, J.; Lukas, J. The ATM-Chk2-Cdc25A checkpoint pathway guards against radioresistant DNA synthesis. Nature 2001, 410, 842–847. [Google Scholar] [CrossRef]
  301. Honaker, Y.; Piwnica-Worms, H. Casein kinase 1 functions as both penultimate and ultimate kinase in regulating Cdc25A destruction. Oncogene 2010, 29, 3324–3334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  302. Lara-Chica, M.; Correa-Saez, A.; Jimenez-Izquierdo, R.; Garrido-Rodriguez, M.; Ponce, F.J.; Moreno, R.; Morrison, K.; Di Vona, C.; Arato, K.; Jimenez-Jimenez, C.; et al. A novel CDC25A/DYRK2 regulatory switch modulates cell cycle and survival. Cell Death Differ. 2022, 29, 105–117. [Google Scholar] [CrossRef] [PubMed]
  303. Kang, T.; Wei, Y.; Honaker, Y.; Yamaguchi, H.; Appella, E.; Hung, M.C.; Piwnica-Worms, H. GSK-3 beta targets Cdc25A for ubiquitin-mediated proteolysis, and GSK-3 beta inactivation correlates with Cdc25A overproduction in human cancers. Cancer Cell 2008, 13, 36–47. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Melixetian, M.; Klein, D.K.; Sorensen, C.S.; Helin, K. NEK11 regulates CDC25A degradation and the IR-induced G2/M checkpoint. Nat. Cell Biol. 2009, 11, 1247–1253. [Google Scholar] [CrossRef] [PubMed]
  305. Wu, C.F.; Liu, S.; Lee, Y.C.; Wang, R.; Sun, S.; Yin, F.; Bornmann, W.G.; Yu-Lee, L.Y.; Gallick, G.E.; Zhang, W.; et al. RSK promotes G2/M transition through activating phosphorylation of Cdc25A and Cdc25B. Oncogene 2014, 33, 2385–2394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  306. Dutertre, S.; Cazales, M.; Quaranta, M.; Froment, C.; Trabut, V.; Dozier, C.; Mirey, G.; Bouche, J.P.; Theis-Febvre, N.; Schmitt, E.; et al. Phosphorylation of CDC25B by Aurora-A at the centrosome contributes to the G2-M transition. J. Cell Sci. 2004, 117 Pt 12, 2523–2531. [Google Scholar] [CrossRef] [Green Version]
  307. Astuti, P.; Boutros, R.; Ducommun, B.; Gabrielli, B. Mitotic phosphorylation of Cdc25B Ser321 disrupts 14-3-3 binding to the high affinity Ser323 site. J. Biol. Chem. 2010, 285, 34364–34370. [Google Scholar] [CrossRef] [Green Version]
  308. Schmitt, E.; Boutros, R.; Froment, C.; Monsarrat, B.; Ducommun, B.; Dozier, C. CHK1 phosphorylates CDC25B during the cell cycle in the absence of DNA damage. J. Cell Sci. 2006, 119 Pt 20, 4269–4275. [Google Scholar] [CrossRef] [Green Version]
  309. Mirey, G.; Chartrain, I.; Froment, C.; Quaranta, M.; Bouche, J.P.; Monsarrat, B.; Tassan, J.P.; Ducommun, B. CDC25B phosphorylated by pEg3 localizes to the centrosome and the spindle poles at mitosis. Cell Cycle 2005, 4, 806–811. [Google Scholar] [CrossRef] [Green Version]
  310. Lu, R.; Niida, H.; Nakanishi, M. Human SAD1 kinase is involved in UV-induced DNA damage checkpoint function. J. Biol. Chem. 2004, 279, 31164–31170. [Google Scholar] [CrossRef] [Green Version]
  311. Bulavin, D.V.; Higashimoto, Y.; Demidenko, Z.N.; Meek, S.; Graves, P.; Phillips, C.; Zhao, H.; Moody, S.A.; Appella, E.; Piwnica-Worms, H.; et al. Dual phosphorylation controls Cdc25 phosphatases and mitotic entry. Nat. Cell Biol. 2003, 5, 545–551. [Google Scholar] [CrossRef]
  312. Strausfeld, U.; Fernandez, A.; Capony, J.P.; Girard, F.; Lautredou, N.; Derancourt, J.; Labbe, J.C.; Lamb, N.J. Activation of p34cdc2 protein kinase by microinjection of human cdc25C into mammalian cells. Requirement for prior phosphorylation of cdc25C by p34cdc2 on sites phosphorylated at mitosis. J. Biol. Chem. 1994, 269, 5989–6000. [Google Scholar] [CrossRef]
  313. Bagheri-Yarmand, R.; Nanos-Webb, A.; Biernacka, A.; Bui, T.; Keyomarsi, K. Cyclin E deregulation impairs mitotic progression through premature activation of Cdc25C. Cancer Res. 2010, 70, 5085–5095. [Google Scholar] [CrossRef] [Green Version]
  314. Schwindling, S.L.; Noll, A.; Montenarh, M.; Gotz, C. Mutation of a CK2 phosphorylation site in cdc25C impairs importin alpha/beta binding and results in cytoplasmic retention. Oncogene 2004, 23, 4155–4165. [Google Scholar] [CrossRef] [Green Version]
  315. Gutierrez, G.J.; Tsuji, T.; Cross, J.V.; Davis, R.J.; Templeton, D.J.; Jiang, W.; Ronai, Z.A. JNK-mediated phosphorylation of Cdc25C regulates cell cycle entry and G(2)/M DNA damage checkpoint. J. Biol. Chem. 2010, 285, 14217–14228. [Google Scholar] [CrossRef] [Green Version]
  316. Goss, V.L.; Cross, J.V.; Ma, K.; Qian, Y.; Mola, P.W.; Templeton, D.J. SAPK/JNK regulates cdc2/cyclin B kinase through phosphorylation and inhibition of cdc25c. Cell Signal. 2003, 15, 709–718. [Google Scholar] [CrossRef]
  317. Gheghiani, L.; Loew, D.; Lombard, B.; Mansfeld, J.; Gavet, O. PLK1 Activation in Late G2 Sets Up Commitment to Mitosis. Cell Rep. 2017, 19, 2060–2073. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  318. Muller, J.; Ritt, D.A.; Copeland, T.D.; Morrison, D.K. Functional analysis of C-TAK1 substrate binding and identification of PKP2 as a new C-TAK1 substrate. EMBO J. 2003, 22, 4431–4442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  319. Dupre-Crochet, S.; Figueroa, A.; Hogan, C.; Ferber, E.C.; Bialucha, C.U.; Adams, J.; Richardson, E.C.; Fujita, Y. Casein kinase 1 is a novel negative regulator of E-cadherin-based cell-cell contacts. Mol. Cell. Biol. 2007, 27, 3804–3816. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  320. Chen, C.L.; Wang, S.H.; Chan, P.C.; Shen, M.R.; Chen, H.C. Phosphorylation of E-cadherin at threonine 790 by protein kinase Cdelta reduces beta-catenin binding and suppresses the function of E-cadherin. Oncotarget 2016, 7, 37260–37276. [Google Scholar] [CrossRef]
  321. Mukherjee, M.; Chow, S.Y.; Yusoff, P.; Seetharaman, J.; Ng, C.; Sinniah, S.; Koh, X.W.; Asgar, N.F.; Li, D.; Yim, D.; et al. Structure of a novel phosphotyrosine-binding domain in Hakai that targets E-cadherin. EMBO J. 2012, 31, 1308–1319. [Google Scholar] [CrossRef] [Green Version]
  322. Zhang, Y.; Sun, L.; Gao, X.; Guo, A.; Diao, Y.; Zhao, Y. RNF43 ubiquitinates and degrades phosphorylated E-cadherin by c-Src to facilitate epithelial-mesenchymal transition in lung adenocarcinoma. BMC Cancer 2019, 19, 670. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  323. Li, X.; Song, Y. Proteolysis-targeting chimera (PROTAC) for targeted protein degradation and cancer therapy. J. Hematol. Oncol. 2020, 13, 50. [Google Scholar] [CrossRef] [PubMed]
  324. Toulany, M. Targeting DNA Double-Strand Break Repair Pathways to Improve Radiotherapy Response. Genes 2019, 10, 25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  325. Garcia, M.E.G.; Kirsch, D.G.; Reitman, Z.J. Targeting the ATM Kinase to Enhance the Efficacy of Radiotherapy and Outcomes for Cancer Patients. Semin. Radiat. Oncol. 2022, 32, 3–14. [Google Scholar] [CrossRef] [PubMed]
  326. Riches, L.C.; Trinidad, A.G.; Hughes, G.; Jones, G.N.; Hughes, A.M.; Thomason, A.G.; Gavine, P.; Cui, A.; Ling, S.; Stott, J.; et al. Pharmacology of the ATM Inhibitor AZD0156: Potentiation of Irradiation and Olaparib Responses Preclinically. Mol. Cancer Ther. 2020, 19, 13–25. [Google Scholar] [CrossRef] [Green Version]
  327. Huang, C.; Filippone, N.R.; Reiner, T.; Roberts, S. Sensors and Inhibitors for the Detection of Ataxia Telangiectasia Mutated (ATM) Protein Kinase. Mol. Pharm. 2021, 18, 2470–2481. [Google Scholar] [CrossRef]
  328. Davis, S.L.; Hartman, S.J.; Bagby, S.M.; Schlaepfer, M.; Yacob, B.W.; Tse, T.; Simmons, D.M.; Diamond, J.R.; Lieu, C.H.; Leal, A.D.; et al. ATM kinase inhibitor AZD0156 in combination with irinotecan and 5-fluorouracil in preclinical models of colorectal cancer. BMC Cancer 2022, 22, 1107. [Google Scholar] [CrossRef]
  329. Zimmermann, A.; Zenke, F.T.; Chiu, L.Y.; Dahmen, H.; Pehl, U.; Fuchss, T.; Grombacher, T.; Blume, B.; Vassilev, L.T.; Blaukat, A. A New Class of Selective ATM Inhibitors as Combination Partners of DNA Double-Strand Break Inducing Cancer Therapies. Mol. Cancer Ther. 2022, 21, 859–870. [Google Scholar] [CrossRef]
  330. Waqar, S.N.; Robinson, C.; Olszanski, A.J.; Spira, A.; Hackmaster, M.; Lucas, L.; Sponton, L.; Jin, H.; Hering, U.; Cronier, D.; et al. Author Correction: Phase I trial of ATM inhibitor M3541 in combination with palliative radiotherapy in patients with solid tumors. Investig. New Drugs 2022, 40, 679. [Google Scholar] [CrossRef]
  331. Giacosa, S.; Pillet, C.; Seraudie, I.; Guyon, L.; Wallez, Y.; Roelants, C.; Battail, C.; Evrard, B.; Chalmel, F.; Barette, C.; et al. Cooperative Blockade of CK2 and ATM Kinases Drives Apoptosis in VHL-Deficient Renal Carcinoma Cells through ROS Overproduction. Cancers 2021, 13, 576. [Google Scholar] [CrossRef]
  332. Middleton, M.R.; Dean, E.; Evans, T.R.J.; Shapiro, G.I.; Pollard, J.; Hendriks, B.S.; Falk, M.; Diaz-Padilla, I.; Plummer, R. Phase 1 study of the ATR inhibitor berzosertib (formerly M6620, VX-970) combined with gemcitabine +/- cisplatin in patients with advanced solid tumours. Br. J. Cancer 2021, 125, 510–519. [Google Scholar] [CrossRef]
  333. 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]
  334. Li, S.; Wang, T.; Fei, X.; Zhang, M. ATR Inhibitors in Platinum-Resistant Ovarian Cancer. Cancers 2022, 14, 5902. [Google Scholar] [CrossRef] [PubMed]
  335. Yap, T.A.; Krebs, M.G.; Postel-Vinay, S.; El-Khouiery, A.; Soria, J.C.; Lopez, J.; Berges, A.; Cheung, S.Y.A.; Irurzun-Arana, I.; Goldwin, A.; et al. Ceralasertib (AZD6738), an Oral ATR Kinase Inhibitor, in Combination with Carboplatin in Patients with Advanced Solid Tumors: A Phase I Study. Clin. Cancer Res. 2021, 27, 5213–5224. [Google Scholar] [CrossRef] [PubMed]
  336. Kwon, M.; Kim, G.; Kim, R.; Kim, K.T.; Kim, S.T.; Smith, S.; Mortimer, P.G.S.; Hong, J.Y.; Loembe, A.B.; Irurzun-Arana, I.; et al. Phase II study of ceralasertib (AZD6738) in combination with durvalumab in patients with advanced gastric cancer. J. Immunother. Cancer 2022, 10, e005041. [Google Scholar] [CrossRef] [PubMed]
  337. Haines, E.; Nishida, Y.; Carr, M.I.; Montoya, R.H.; Ostermann, L.B.; Zhang, W.; Zenke, F.T.; Blaukat, A.; Andreeff, M.; Vassilev, L.T. DNA-PK inhibitor peposertib enhances p53-dependent cytotoxicity of DNA double-strand break inducing therapy in acute leukemia. Sci. Rep. 2021, 11, 12148. [Google Scholar] [CrossRef]
  338. van Bussel, M.T.J.; Awada, A.; de Jonge, M.J.A.; Mau-Sorensen, M.; Nielsen, D.; Schoffski, P.; Verheul, H.M.W.; Sarholz, B.; Berghoff, K.; El Bawab, S.; et al. A first-in-man phase 1 study of the DNA-dependent protein kinase inhibitor peposertib (formerly M3814) in patients with advanced solid tumours. Br. J. Cancer 2021, 124, 728–735. [Google Scholar] [CrossRef]
  339. Fok, J.H.L.; Ramos-Montoya, A.; Vazquez-Chantada, M.; Wijnhoven, P.W.G.; Follia, V.; James, N.; Farrington, P.M.; Karmokar, A.; Willis, S.E.; Cairns, J.; et al. AZD7648 is a potent and selective DNA-PK inhibitor that enhances radiation, chemotherapy and olaparib activity. Nat. Commun. 2019, 10, 5065. [Google Scholar] [CrossRef] [Green Version]
  340. Trenner, A.; Sartori, A.A. Harnessing DNA Double-Strand Break Repair for Cancer Treatment. Front. Oncol. 2019, 9, 1388. [Google Scholar] [CrossRef] [Green Version]
  341. Neizer-Ashun, F.; Bhattacharya, R. Reality CHEK: Understanding the biology and clinical potential of CHK1. Cancer Lett. 2021, 497, 202–211. [Google Scholar] [CrossRef]
  342. Cash, T.; Fox, E.; Liu, X.; Minard, C.G.; Reid, J.M.; Scheck, A.C.; Weigel, B.J.; Wetmore, C. A phase 1 study of prexasertib (LY2606368), a CHK1/2 inhibitor, in pediatric patients with recurrent or refractory solid tumors, including CNS tumors: A report from the Children’s Oncology Group Pediatric Early Phase Clinical Trials Network (ADVL1515). Pediatr. Blood Cancer 2021, 68, e29065. [Google Scholar] [CrossRef]
  343. King, C.; Diaz, H.B.; McNeely, S.; Barnard, D.; Dempsey, J.; Blosser, W.; Beckmann, R.; Barda, D.; Marshall, M.S. LY2606368 Causes Replication Catastrophe and Antitumor Effects through CHK1-Dependent Mechanisms. Mol. Cancer Ther. 2015, 14, 2004–2013. [Google Scholar] [CrossRef] [Green Version]
  344. Do, K.T.; Manuszak, C.; Thrash, E.; Giobbie-Hurder, A.; Hu, J.; Kelland, S.; Powers, A.; de Jonge, A.; Shapiro, G.I.; Severgnini, M. Immune modulating activity of the CHK1 inhibitor prexasertib and anti-PD-L1 antibody LY3300054 in patients with high-grade serous ovarian cancer and other solid tumors. Cancer Immunol. Immunother. 2021, 70, 2991–3000. [Google Scholar] [CrossRef] [PubMed]
  345. Everix, L.; Nair, S.; Driver, C.H.S.; Goethals, I.; Sathekge, M.M.; Ebenhan, T.; Vandevoorde, C.; Bolcaen, J. Perspective on the Use of DNA Repair Inhibitors as a Tool for Imaging and Radionuclide Therapy of Glioblastoma. Cancers 2022, 14, 1821. [Google Scholar] [CrossRef] [PubMed]
  346. Zhou, Z.R.; Yang, Z.Z.; Wang, S.J.; Zhang, L.; Luo, J.R.; Feng, Y.; Yu, X.L.; Chen, X.X.; Guo, X.M. The Chk1 inhibitor MK-8776 increases the radiosensitivity of human triple-negative breast cancer by inhibiting autophagy. Acta Pharmacol. Sin. 2017, 38, 513–523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  347. Daud, A.I.; Ashworth, M.T.; Strosberg, J.; Goldman, J.W.; Mendelson, D.; Springett, G.; Venook, A.P.; Loechner, S.; Rosen, L.S.; Shanahan, F.; et al. Phase I dose-escalation trial of checkpoint kinase 1 inhibitor MK-8776 as monotherapy and in combination with gemcitabine in patients with advanced solid tumors. J. Clin. Oncol. 2015, 33, 1060–1066. [Google Scholar] [CrossRef]
  348. Park, S.J.; Chang, S.J.; Suh, D.H.; Kong, T.W.; Song, H.; Kim, T.H.; Kim, J.W.; Kim, H.S.; Lee, S.J. A phase IA dose-escalation study of PHI-101, a new checkpoint kinase 2 inhibitor, for platinum-resistant recurrent ovarian cancer. BMC Cancer 2022, 22, 28. [Google Scholar] [CrossRef]
  349. Ronco, C.; Martin, A.R.; Demange, L.; Benhida, R. ATM, ATR, CHK1, CHK2 and WEE1 inhibitors in cancer and cancer stem cells. MedChemComm 2017, 8, 295–319. [Google Scholar] [CrossRef]
  350. Mazhar, S.; Taylor, S.E.; Sangodkar, J.; Narla, G. Targeting PP2A in cancer: Combination therapies. Biochim. Biophys. Acta Mol. Cell Res. 2019, 1866, 51–63. [Google Scholar] [CrossRef]
  351. Chung, V.; Mansfield, A.S.; Braiteh, F.; Richards, D.; Durivage, H.; Ungerleider, R.S.; Johnson, F.; Kovach, J.S. Safety, Tolerability, and Preliminary Activity of LB-100, an Inhibitor of Protein Phosphatase 2A, in Patients with Relapsed Solid Tumors: An Open-Label, Dose Escalation, First-in-Human, Phase I Trial. Clin. Cancer Res. 2017, 23, 3277–3284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  352. Otani, Y.; Sur, H.P.; Rachaiah, G.; Namagiri, S.; Chowdhury, A.; Lewis, C.T.; Shimizu, T.; Gangaplara, A.; Wang, X.; Vezina, A.; et al. Inhibiting protein phosphatase 2A increases the antitumor effect of protein arginine methyltransferase 5 inhibition in models of glioblastoma. Neuro Oncol. 2021, 23, 1481–1493. [Google Scholar] [CrossRef]
  353. Hu, C.; Yu, M.; Ren, Y.; Li, K.; Maggio, D.M.; Mei, C.; Ye, L.; Wei, J.; Jin, J.; Zhuang, Z.; et al. PP2A inhibition from LB100 therapy enhances daunorubicin cytotoxicity in secondary acute myeloid leukemia via miR-181b-1 upregulation. Sci. Rep. 2017, 7, 2894. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  354. Pechackova, S.; Burdova, K.; Benada, J.; Kleiblova, P.; Jenikova, G.; Macurek, L. Inhibition of WIP1 phosphatase sensitizes breast cancer cells to genotoxic stress and to MDM2 antagonist nutlin-3. Oncotarget 2016, 7, 14458–14475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  355. Deng, W.; Li, J.; Dorrah, K.; Jimenez-Tapia, D.; Arriaga, B.; Hao, Q.; Cao, W.; Gao, Z.; Vadgama, J.; Wu, Y. The role of PPM1D in cancer and advances in studies of its inhibitors. Biomed. Pharmacother. 2020, 125, 109956. [Google Scholar] [CrossRef] [PubMed]
  356. Miller, P.G.; Sathappa, M.; Moroco, J.A.; Jiang, W.; Qian, Y.; Iqbal, S.; Guo, Q.; Giacomelli, A.O.; Shaw, S.; Vernier, C.; et al. Allosteric inhibition of PPM1D serine/threonine phosphatase via an altered conformational state. Nat. Commun. 2022, 13, 3778. [Google Scholar] [CrossRef] [PubMed]
  357. Chen, Z.; Wang, L.; Yao, D.; Yang, T.; Cao, W.M.; Dou, J.; Pang, J.C.; Guan, S.; Zhang, H.; Yu, Y.; et al. Wip1 inhibitor GSK2830371 inhibits neuroblastoma growth by inducing Chk2/p53-mediated apoptosis. Sci. Rep. 2016, 6, 38011. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The Apical (sensor) kinases and co-factors in DNA double-strand break repair response.
Figure 1. The Apical (sensor) kinases and co-factors in DNA double-strand break repair response.
Ijms 24 10212 g001
Figure 2. The targets of kinases in DNA double-strand break repair response.
Figure 2. The targets of kinases in DNA double-strand break repair response.
Ijms 24 10212 g002
Table 1. A summary of main kinases, phosphatases, and their substrates in DSBs repair in mammalians.
Table 1. A summary of main kinases, phosphatases, and their substrates in DSBs repair in mammalians.
Subtract (Including Its Phosphorylated Form)Kinase
(In Vitro)
Phosphatase
(In Vitro)
ATMATM [31,32], ATR [81], AurB [82], CDK5 [83],
EGFR [84]
PPM1D [85] PPP2CA [86]
ATRATR [34], ABL [87], PKACA [88]N/A
DNAPKATM [27], ATR [89], DNAPK [90], EGFR [91],
PLK1 [92]
PPP6C [92]
CHK1ATM [93], ATR [34], AKT1 [94], CDK1 [95],
CHK1 [96], p90RSK [97]
PPM1D [69], PPP2CA [38]
CHK2ATM [35], ATR [35], CHK2 [98], DNAPK [99],
JAK2 [100], PLK1 [101], ZAK [102]
PPM1D [103], PPP1CA [103],
PPP2CA [103]
H2AXATM [104], DNA-PK [105], CK2A1 [106],
JMJD6 [107], MST1 [108], RSK2 [109]
EYA2 [110]. EYA3 [111],
PPM1D [85,112],
PPP2CA [113,114]
MDC1ATM [115,116], ATR [117], CK2A1 [118],
PLK1 [119,120]
N/A
53BP1ATM [121], AMPKA1 [122], AMPKA2 [122],
ATR [123], AurB [124], PLK1 [125]
N/A
RPA2ATM [126], ATR [127,128],
DNAPK [126,129,130]
PPP2CA [131], PPP4C [132]
Ku70/ Ku80ATM [133,134], DNAPK [134], PKCA [135],
Src [136]
N/A
NBS1ATM [137,138,139,140,141,142], ATR [143], CDK2 [144]PPP1CA [145]
ClaspinCDKL5 [146], CHEK1 [147], PLK1 [148]N/A
BRCA1Akt1 [149,150], ATM [141,151], ATR [152,153,154], AurA [155], CDK1 [156], CDK2 [157],
CDK4 [158], CHEK2 [159], FRK [160],
PLK1 [161]
PPP6C [162]
BRCA2CDK1 [163], PLK1 [164,165]N/A
P53ATM [141,166,167,168,169,170,171,172,173,174,175,176,177,178,179], ATR [175,180],
AurA [181,182,183], AurB [184,185], Btk [186],
CDK1 [187,188,189], CDK4 [190], CDK5 [191],
CDK6 [192,193], CDK9 [194], CHEK1 [195,196,197,198], CHEK2 [174,196,197,199], CK1A [200,201],
CK2B [202], DAPK1 [197], DNAPK
[195,196,197,203,204,205,206,207,208,209,210,211,212], DYRK1A [213],
DYRK2 [214], ERK1 [215], ERK2 [216,217,218],
GRK2 [219], GRK5 [220], HIPK2 [221,222],
JNK1 [217,223], JNK2 [217], LKB1 [224],
Lmr1 [225], LRRK2 [226],
MAPKAPK5 [227,228], NEK2 [183],
NuaK1 [229], P38A [227,230,231,232,233,234], P38G [233],
PAK4 [235], PKCD [236], PKR [237],
PLK3 [238,239,240], PRPK [241,242], SMG1,
Src [243], TAF1 [244,245,246], VRK1 [247]
CDC14A [248], CDC25B [249],
DUSP26 [250], PPM1D [85,251], PPP1CA [204], PPP2CA [252],
PPP2CB [253], PPP2R5C [254]
MDM2Abl [255,256], Akt1 [257,258,259,260,261,262], ATM [263,264,265],
ATR [266], CDK9 [267], CK1D [268],
ERK2 [269], MAPKAPK2 [270], PAK6 [271],
Pim1 [272], PLK1 [273]
PPP2CA [274,275]
PLK1Abl [276], AurA [277,278,279,280,281], CHEK1 [282],
Cot [283], MAPKAPK2 [284], PAK1 [285],
PLK1 [286], PLK2 [287], VRK2 [288]
PPP1CA [289]
WEE1Akt1 [290], CDK1 [291]CDC14A [291]
CDC25ACDK1 [292,293,294], CDK2 [292]
CHEK1 [295,296,297,298,299],
CHEK2 [298,300], CK1A [301],
DYRK2 [302], GSK3B [303], NEK11 [304],
p90RSK [305], PLK3 [303], Src [294]
N/A
CDC25BAurA [306], CDK1 [307], CHEK1 [308],
MELK [309], p90RSK [305]
N/A
CDC25CBRSK1 iso2 [310], CDK1 [311,312],
CDK2 [313], CK2A1 [314], JNK1,
JNK2 [315,316], PLK1 [317], TAK1 [318]
N/A
CDH1CK1A [319], CK1E [319], PKCD [320],
Src [321,322]
N/A
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

Qin, S.; Kitty, I.; Hao, Y.; Zhao, F.; Kim, W. Maintaining Genome Integrity: Protein Kinases and Phosphatases Orchestrate the Balancing Act of DNA Double-Strand Breaks Repair in Cancer. Int. J. Mol. Sci. 2023, 24, 10212. https://doi.org/10.3390/ijms241210212

AMA Style

Qin S, Kitty I, Hao Y, Zhao F, Kim W. Maintaining Genome Integrity: Protein Kinases and Phosphatases Orchestrate the Balancing Act of DNA Double-Strand Breaks Repair in Cancer. International Journal of Molecular Sciences. 2023; 24(12):10212. https://doi.org/10.3390/ijms241210212

Chicago/Turabian Style

Qin, Sisi, Ichiwa Kitty, Yalan Hao, Fei Zhao, and Wootae Kim. 2023. "Maintaining Genome Integrity: Protein Kinases and Phosphatases Orchestrate the Balancing Act of DNA Double-Strand Breaks Repair in Cancer" International Journal of Molecular Sciences 24, no. 12: 10212. https://doi.org/10.3390/ijms241210212

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