DNA-PKcs

Last updated
PRKDC
Identifiers
Aliases PRKDC , DNA-PKcs, DNAPK, DNPK1, HYRC, HYRC1, XRCC7, p350, IMD26, protein kinase, DNA-activated, catalytic polypeptide, DNA-PKC, protein kinase, DNA-activated, catalytic subunit, DNAPKc
External IDs OMIM: 600899 MGI: 104779 HomoloGene: 5037 GeneCards: PRKDC
Orthologs
SpeciesHumanMouse
Entrez
Ensembl
UniProt
RefSeq (mRNA)

NM_001081640
NM_006904

NM_011159

RefSeq (protein)

NP_001075109
NP_008835

NP_035289

Location (UCSC) Chr 8: 47.77 – 47.96 Mb Chr 16: 15.46 – 15.66 Mb
PubMed search [3] [4]
Wikidata
View/Edit Human View/Edit Mouse

DNA-dependent protein kinase, catalytic subunit, also known as DNA-PKcs, is an enzyme that in humans is encoded by the gene designated as PRKDC or XRCC7. [5] DNA-PKcs belongs to the phosphatidylinositol 3-kinase-related kinase protein family. The DNA-Pkcs protein is a serine/threonine protein kinase consisting of a single polypeptide chain of 4,128 amino acids. [6] [7]

Contents

Function

DNA-PKcs is the catalytic subunit of a nuclear DNA-dependent serine/threonine protein kinase called DNA-PK. The second component is the autoimmune antigen Ku. On its own, DNA-PKcs is inactive and relies on Ku to direct it to DNA ends and trigger its kinase activity. [8] DNA-PKcs is required for the non-homologous end joining (NHEJ) pathway of DNA repair, which rejoins double-strand breaks. It is also required for V(D)J recombination, a process that utilizes NHEJ to promote immune system diversity.

Many proteins have been identified as substrates for the kinase activity of DNA-PK. Autophosphorylation of DNA-PKcs appears to play a key role in NHEJ and is thought to induce a conformational change that allows end processing enzymes to access the ends of the double-strand break. [9] DNA-PK also cooperates with ATR and ATM to phosphorylate proteins involved in the DNA damage checkpoint.

Disease

DNA-PKcs knockout mice have severe combined immunodeficiency due to their V(D)J recombination defect. Natural analogs of this knockout happen in mice, horses and dogs, also causing SCID. [10] Human SCID usually have other causes, but two cases related to mutations in this gene are also known. [11]

Apoptosis

DNA-PKcs activates p53 to regulate apoptosis [12] . In response to ionizing radiation, DNA-PKcs can serve as an upstream effector for p53 protein activation, thus linking DNA damage to apoptosis [12] . Both Repair of DNA damages and apoptosis are catalytic activities required for maintaining integrity of the human genome. Cells that have insufficient DNA repair capability tend to accumulate DNA damages, and when such cells are additionally defective in apoptosis they tend to survive even though excessive DNA damages are present [13] . The replication of DNA in such deficient cells can generate mutations and such mutations may cause cancer. Thus DNA-PKcs appears to have two functions related to the prevention of cancer, where the first function is to participate in the repair of DNA double-strand breaks by the NHEJ repair pathway and the second function is to induce apoptosis if the level of such DNA breaks exceed the cell’s repair capability [13]

Cancer

DNA damage appears to be the primary underlying cause of cancer, [14] and deficiencies in DNA repair genes likely underlie many forms of cancer. [15] [16] If DNA repair is deficient, DNA damage tends to accumulate. Such excess DNA damage may increase mutations due to error-prone translesion synthesis. Excess DNA damage may also increase epigenetic alterations due to errors during DNA repair. [17] [18] Such mutations and epigenetic alterations may give rise to cancer.

PRKDC (DNA-PKcs) mutations were found in 3 out of 10 of endometriosis-associated ovarian cancers, as well as in the field defects from which they arose. [19] They were also found in 10% of breast and pancreatic cancers. [20]

Reductions in expression of DNA repair genes (usually caused by epigenetic alterations) are very common in cancers, and are ordinarily even more frequent than mutational defects in DNA repair genes in cancers.[ citation needed ] DNA-PKcs expression was reduced by 23% to 57% in six cancers as indicated in the table.

Frequency of reduced expression of DNA-PKcs in sporadic cancers
CancerFrequency of reduction in cancerRef.
Breast cancer57% [21]
Prostate cancer51% [22]
Cervical carcinoma32% [23]
Nasopharyngeal carcinoma30% [24]
Epithelial ovarian cancer29% [25]
Gastric cancer23% [26]

It is not clear what causes reduced expression of DNA-PKcs in cancers. MicroRNA-101 targets DNA-PKcs via binding to the 3'- UTR of DNA-PKcs mRNA and efficiently reduces protein levels of DNA-PKcs. [27] But miR-101 is more often decreased in cancers, rather than increased. [28] [29]

HMGA2 protein could also have an effect on DNA-PKcs. HMGA2 delays the release of DNA-PKcs from sites of double-strand breaks, interfering with DNA repair by non-homologous end joining and causing chromosomal aberrations. [30] The let-7a microRNA normally represses the HMGA2 gene. [31] [32] In normal adult tissues, almost no HMGA2 protein is present. In many cancers, let-7 microRNA is repressed. As an example, in breast cancers the promoter region controlling let-7a-3/let-7b microRNA is frequently repressed by hypermethylation. [33] Epigenetic reduction or absence of let-7a microRNA allows high expression of the HMGA2 protein and this would lead to defective expression of DNA-PKcs.

DNA-PKcs can be up-regulated by stressful conditions such as in Helicobacter pylori-associated gastritis. [34] After ionizing radiation DNA-PKcs was increased in the surviving cells of oral squamous cell carcinoma tissues. [35]

The ATM protein is important in homologous recombinational repair (HRR) of DNA double strand breaks. When cancer cells are deficient in ATM the cells are "addicted" to DNA-PKcs, important in the alternative DNA repair pathway for double-strand breaks, non-homologous end joining (NHEJ). [36] That is, in ATM-mutant cells, an inhibitor of DNA-PKcs causes high levels of apoptotic cell death. In ATM mutant cells, additional loss of DNA-PKcs leaves the cells without either major pathway (HRR and NHEJ) for repair of DNA double-strand breaks.

Elevated DNA-PKcs expression is found in a large fraction (40% to 90%) of some cancers (the remaining fraction of cancers often has reduced or absent expression of DNA-PKcs). The elevation of DNA-PKcs is thought to reflect the induction of a compensatory DNA repair capability, due to the genome instability in these cancers. [37] (As indicated in the article Genome instability, such genome instability may be due to deficiencies in other DNA repair genes present in the cancers.) Elevated DNA-PKcs is thought to be "beneficial to the tumor cells", [37] though it would be at the expense of the patient. As indicated in a table listing 12 types of cancer reported in 20 publications, [37] the fraction of cancers with over-expression of DNA-PKcs is often associated with an advanced stage of the cancer and shorter survival time for the patient. However, the table also indicates that for some cancers, the fraction of cancers with reduced or absent DNA-PKcs is also associated with advanced stage and poor patient survival.

Aging

Non-homologous end joining (NHEJ) is the principal DNA repair process used by mammalian somatic cells to cope with double-strand breaks that continually occur in the genome. DNA-PKcs is one of the key components of the NHEJ machinery. DNA-PKcs deficient mice have a shorter lifespan and show an earlier onset of numerous aging related pathologies than corresponding wild-type littermates. [38] [39] These findings suggest that failure to efficiently repair DNA double-strand breaks results in premature aging, consistent with the DNA damage theory of aging. (See also Bernstein et al. [40] )

Interactions

DNA-PKcs has been shown to interact with:

DNA-PKcs Inhibitors

AZD7648, [56] M3814 (peposertib), [57] M9831 (VX-984) [58] and BAY-8400 [59] have been described as potent and selective DNA-PKcs inhibitors.

See also

Related Research Articles

<span class="mw-page-title-main">BRCA1</span> Gene known for its role in breast cancer

Breast cancer type 1 susceptibility protein is a protein that in humans is encoded by the BRCA1 gene. Orthologs are common in other vertebrate species, whereas invertebrate genomes may encode a more distantly related gene. BRCA1 is a human tumor suppressor gene and is responsible for repairing DNA.

<span class="mw-page-title-main">DNA repair</span> Cellular mechanism

DNA repair is a collection of processes by which a cell identifies and corrects damage to the DNA molecules that encodes its genome. In human cells, both normal metabolic activities and environmental factors such as radiation can cause DNA damage, resulting in tens of thousands of individual molecular lesions per cell per day. Many of these lesions cause structural damage to the DNA molecule and can alter or eliminate the cell's ability to transcribe the gene that the affected DNA encodes. Other lesions induce potentially harmful mutations in the cell's genome, which affect the survival of its daughter cells after it undergoes mitosis. As a consequence, the DNA repair process is constantly active as it responds to damage in the DNA structure. When normal repair processes fail, and when cellular apoptosis does not occur, irreparable DNA damage may occur. This can eventually lead to malignant tumors, or cancer as per the two-hit hypothesis.

In biochemistry, in the biological context of organisms' regulation of gene expression and production of gene products, downregulation is the process by which a cell decreases the production and quantities of its cellular components, such as RNA and proteins, in response to an external stimulus. The complementary process that involves increase in quantities of cellular components is called upregulation.

<span class="mw-page-title-main">Non-homologous end joining</span> Pathway that repairs double-strand breaks in DNA

Non-homologous end joining (NHEJ) is a pathway that repairs double-strand breaks in DNA. It is called "non-homologous" because the break ends are directly ligated without the need for a homologous template, in contrast to homology directed repair (HDR), which requires a homologous sequence to guide repair. NHEJ is active in both non-dividing and proliferating cells, while HDR is not readily accessible in non-dividing cells. The term "non-homologous end joining" was coined in 1996 by Moore and Haber.

<span class="mw-page-title-main">ATM serine/threonine kinase</span> Mammalian protein found in Homo sapiens

ATM serine/threonine kinase or Ataxia-telangiectasia mutated, symbol ATM, is a serine/threonine protein kinase that is recruited and activated by DNA double-strand breaks, oxidative stress, topoisomerase cleavage complexes, splicing intermediates, R-loops and in some cases by single-strand DNA breaks. It phosphorylates several key proteins that initiate activation of the DNA damage checkpoint, leading to cell cycle arrest, DNA repair or apoptosis. Several of these targets, including p53, CHK2, BRCA1, NBS1 and H2AX are tumor suppressors.

<span class="mw-page-title-main">Werner syndrome helicase</span> Protein-coding gene in the species Homo sapiens

Werner syndrome ATP-dependent helicase, also known as DNA helicase, RecQ-like type 3, is an enzyme that in humans is encoded by the WRN gene. WRN is a member of the RecQ Helicase family. Helicase enzymes generally unwind and separate double-stranded DNA. These activities are necessary before DNA can be copied in preparation for cell division. Helicase enzymes are also critical for making a blueprint of a gene for protein production, a process called transcription. Further evidence suggests that Werner protein plays a critical role in repairing DNA. Overall, this protein helps maintain the structure and integrity of a person's DNA.

<span class="mw-page-title-main">Ku (protein)</span>

Ku is a dimeric protein complex that binds to DNA double-strand break ends and is required for the non-homologous end joining (NHEJ) pathway of DNA repair. Ku is evolutionarily conserved from bacteria to humans. The ancestral bacterial Ku is a homodimer. Eukaryotic Ku is a heterodimer of two polypeptides, Ku70 (XRCC6) and Ku80 (XRCC5), so named because the molecular weight of the human Ku proteins is around 70 kDa and 80 kDa. The two Ku subunits form a basket-shaped structure that threads onto the DNA end. Once bound, Ku can slide down the DNA strand, allowing more Ku molecules to thread onto the end. In higher eukaryotes, Ku forms a complex with the DNA-dependent protein kinase catalytic subunit (DNA-PKcs) to form the full DNA-dependent protein kinase, DNA-PK. Ku is thought to function as a molecular scaffold to which other proteins involved in NHEJ can bind, orienting the double-strand break for ligation.

<span class="mw-page-title-main">Ataxia telangiectasia and Rad3 related</span> Protein kinase that detects DNA damage and halts cell division

Serine/threonine-protein kinase ATR, also known as ataxia telangiectasia and Rad3-related protein (ATR) or FRAP-related protein 1 (FRP1), is an enzyme that, in humans, is encoded by the ATR gene. It is a large kinase of about 301.66 kDa. ATR belongs to the phosphatidylinositol 3-kinase-related kinase protein family. ATR is activated in response to single strand breaks, and works with ATM to ensure genome integrity.

<span class="mw-page-title-main">DNA repair protein XRCC4</span> Protein-coding gene in the species Homo sapiens

DNA repair protein XRCC4 also known as X-ray repair cross-complementing protein 4 or XRCC4 is a protein that in humans is encoded by the XRCC4 gene. In addition to humans, the XRCC4 protein is also expressed in many other metazoans, fungi and in plants. The X-ray repair cross-complementing protein 4 is one of several core proteins involved in the non-homologous end joining (NHEJ) pathway to repair DNA double strand breaks (DSBs).

<span class="mw-page-title-main">HMGA2</span> Protein-coding gene in the species Homo sapiens

High-mobility group AT-hook 2, also known as HMGA2, is a protein that, in humans, is encoded by the HMGA2 gene.

<span class="mw-page-title-main">CHEK1</span> Protein-coding gene in humans

Checkpoint kinase 1, commonly referred to as Chk1, is a serine/threonine-specific protein kinase that, in humans, is encoded by the CHEK1 gene. Chk1 coordinates the DNA damage response (DDR) and cell cycle checkpoint response. Activation of Chk1 results in the initiation of cell cycle checkpoints, cell cycle arrest, DNA repair and cell death to prevent damaged cells from progressing through the cell cycle.

<span class="mw-page-title-main">Ku70</span> Protein-coding gene in the species Homo sapiens

Ku70 is a protein that, in humans, is encoded by the XRCC6 gene.

<span class="mw-page-title-main">Ku80</span> Protein-coding gene in the species Homo sapiens

Ku80 is a protein that, in humans, is encoded by the XRCC5 gene. Together, Ku70 and Ku80 make up the Ku heterodimer, which binds to DNA double-strand break ends and is required for the non-homologous end joining (NHEJ) pathway of DNA repair. It is also required for V(D)J recombination, which utilizes the NHEJ pathway to promote antigen diversity in the mammalian immune system.

<span class="mw-page-title-main">ERCC1</span> Protein-coding gene in the species Homo sapiens

DNA excision repair protein ERCC-1 is a protein that in humans is encoded by the ERCC1 gene. Together with ERCC4, ERCC1 forms the ERCC1-XPF enzyme complex that participates in DNA repair and DNA recombination.

The MRN complex is a protein complex consisting of Mre11, Rad50 and Nbs1. In eukaryotes, the MRN/X complex plays an important role in the initial processing of double-strand DNA breaks prior to repair by homologous recombination or non-homologous end joining. The MRN complex binds avidly to double-strand breaks both in vitro and in vivo and may serve to tether broken ends prior to repair by non-homologous end joining or to initiate DNA end resection prior to repair by homologous recombination. The MRN complex also participates in activating the checkpoint kinase ATM in response to DNA damage. Production of short single-strand oligonucleotides by Mre11 endonuclease activity has been implicated in ATM activation by the MRN complex.

Genome instability refers to a high frequency of mutations within the genome of a cellular lineage. These mutations can include changes in nucleic acid sequences, chromosomal rearrangements or aneuploidy. Genome instability does occur in bacteria. In multicellular organisms genome instability is central to carcinogenesis, and in humans it is also a factor in some neurodegenerative diseases such as amyotrophic lateral sclerosis or the neuromuscular disease myotonic dystrophy.

<span class="mw-page-title-main">LIG3</span> Protein-coding gene in the species Homo sapiens

DNA ligase 3 is an enzyme that, in humans, is encoded by the LIG3 gene. The human LIG3 gene encodes ATP-dependent DNA ligases that seal interruptions in the phosphodiester backbone of duplex DNA.

<span class="mw-page-title-main">Cancer epigenetics</span> Field of study in cancer research

Cancer epigenetics is the study of epigenetic modifications to the DNA of cancer cells that do not involve a change in the nucleotide sequence, but instead involve a change in the way the genetic code is expressed. Epigenetic mechanisms are necessary to maintain normal sequences of tissue specific gene expression and are crucial for normal development. They may be just as important, if not even more important, than genetic mutations in a cell's transformation to cancer. The disturbance of epigenetic processes in cancers, can lead to a loss of expression of genes that occurs about 10 times more frequently by transcription silencing than by mutations. As Vogelstein et al. points out, in a colorectal cancer there are usually about 3 to 6 driver mutations and 33 to 66 hitchhiker or passenger mutations. However, in colon tumors compared to adjacent normal-appearing colonic mucosa, there are about 600 to 800 heavily methylated CpG islands in the promoters of genes in the tumors while these CpG islands are not methylated in the adjacent mucosa. Manipulation of epigenetic alterations holds great promise for cancer prevention, detection, and therapy. In different types of cancer, a variety of epigenetic mechanisms can be perturbed, such as the silencing of tumor suppressor genes and activation of oncogenes by altered CpG island methylation patterns, histone modifications, and dysregulation of DNA binding proteins. There are several medications which have epigenetic impact, that are now used in a number of these diseases.

<span class="mw-page-title-main">DNA polymerase alpha catalytic subunit</span> Protein-coding gene in humans

DNA polymerase alpha catalytic subunit is an enzyme that in humans is encoded by the POLA1 gene.

<span class="mw-page-title-main">Double-strand break repair model</span>

A double-strand break repair model refers to the various models of pathways that cells undertake to repair double strand-breaks (DSB). DSB repair is an important cellular process, as the accumulation of unrepaired DSB could lead to chromosomal rearrangements, tumorigenesis or even cell death. In human cells, there are two main DSB repair mechanisms: Homologous recombination (HR) and non-homologous end joining (NHEJ). HR relies on undamaged template DNA as reference to repair the DSB, resulting in the restoration of the original sequence. NHEJ modifies and ligates the damaged ends regardless of homology. In terms of DSB repair pathway choice, most mammalian cells appear to favor NHEJ rather than HR. This is because the employment of HR may lead to gene deletion or amplification in cells which contains repetitive sequences. In terms of repair models in the cell cycle, HR is only possible during the S and G2 phases, while NHEJ can occur throughout whole process. These repair pathways are all regulated by the overarching DNA damage response mechanism. Besides HR and NHEJ, there are also other repair models which exists in cells. Some are categorized under HR, such as synthesis-dependent strain annealing, break-induced replication, and single-strand annealing; while others are an entirely alternate repair model, namely, the pathway microhomology-mediated end joining (MMEJ).

References

  1. 1 2 3 GRCh38: Ensembl release 89: ENSG00000253729 - Ensembl, May 2017
  2. 1 2 3 GRCm38: Ensembl release 89: ENSMUSG00000022672 - Ensembl, May 2017
  3. "Human PubMed Reference:". National Center for Biotechnology Information, U.S. National Library of Medicine.
  4. "Mouse PubMed Reference:". National Center for Biotechnology Information, U.S. National Library of Medicine.
  5. Sipley JD, Menninger JC, Hartley KO, Ward DC, Jackson SP, Anderson CW (August 1995). "Gene for the catalytic subunit of the human DNA-activated protein kinase maps to the site of the XRCC7 gene on chromosome 8". Proceedings of the National Academy of Sciences of the United States of America. 92 (16): 7515–7519. Bibcode:1995PNAS...92.7515S. doi: 10.1073/pnas.92.16.7515 . PMC   41370 . PMID   7638222.
  6. Sibanda BL, Chirgadze DY, Blundell TL (January 2010). "Crystal structure of DNA-PKcs reveals a large open-ring cradle comprised of HEAT repeats". Nature. 463 (7277): 118–121. doi:10.1038/nature08648. PMC   2811870 . PMID   20023628.
  7. Hartley KO, Gell D, Smith GC, Zhang H, Divecha N, Connelly MA, et al. (September 1995). "DNA-dependent protein kinase catalytic subunit: a relative of phosphatidylinositol 3-kinase and the ataxia telangiectasia gene product". Cell. 82 (5): 849–856. doi: 10.1016/0092-8674(95)90482-4 . PMID   7671312.
  8. "Entrez Gene: PRKDC protein kinase, DNA-activated, catalytic polypeptide".
  9. Meek K, Dang V, Lees-Miller SP (2008). Chapter 2 DNA-PK. Advances in Immunology. Vol. 99. pp. 33–58. doi:10.1016/S0065-2776(08)00602-0. ISBN   9780123743251. PMID   19117531.
  10. Meek K, Jutkowitz A, Allen L, Glover J, Convery E, Massa A, et al. (August 2009). "SCID dogs: similar transplant potential but distinct intra-uterine growth defects and premature replicative senescence compared with SCID mice". Journal of Immunology. 183 (4): 2529–2536. doi:10.4049/jimmunol.0801406. PMC   4047667 . PMID   19635917.
  11. Anne Esguerra Z, Watanabe G, Okitsu CY, Hsieh CL, Lieber MR (April 2020). "DNA-PKcs chemical inhibition versus genetic mutation: Impact on the junctional repair steps of V(D)J recombination". Molecular Immunology. 120: 93–100. doi:10.1016/j.molimm.2020.01.018. PMC   7184946 . PMID   32113132.
  12. 1 2 Wang S, Guo M, Ouyang H, Li X, Cordon-Cardo C, Kurimasa A, Chen DJ, Fuks Z, Ling CC, Li GC. The catalytic subunit of DNA-dependent protein kinase selectively regulates p53-dependent apoptosis but not cell-cycle arrest. Proc Natl Acad Sci U S A. 2000 Feb 15;97(4):1584-8. doi: 10.1073/pnas.97.4.1584. PMID: 10677503; PMCID: PMC26478
  13. 1 2 Bernstein C, Bernstein H, Payne CM, Garewal H. DNA repair/pro-apoptotic dual-role proteins in five major DNA repair pathways: fail-safe protection against carcinogenesis. Mutat Res. 2002 Jun;511(2):145-78. doi: 10.1016/s1383-5742(02)00009-1. PMID: 12052432
  14. Kastan MB (April 2008). "DNA damage responses: mechanisms and roles in human disease: 2007 G.H.A. Clowes Memorial Award Lecture". Molecular Cancer Research. 6 (4): 517–524. doi: 10.1158/1541-7786.MCR-08-0020 . PMID   18403632.
  15. Harper JW, Elledge SJ (December 2007). "The DNA damage response: ten years after". Molecular Cell. 28 (5): 739–745. doi: 10.1016/j.molcel.2007.11.015 . PMID   18082599.
  16. Dietlein F, Reinhardt HC (December 2014). "Molecular pathways: exploiting tumor-specific molecular defects in DNA repair pathways for precision cancer therapy". Clinical Cancer Research. 20 (23): 5882–5887. doi:10.1158/1078-0432.CCR-14-1165. PMID   25451105.
  17. O'Hagan HM, Mohammad HP, Baylin SB (August 2008). "Double strand breaks can initiate gene silencing and SIRT1-dependent onset of DNA methylation in an exogenous promoter CpG island". PLOS Genetics. 4 (8): e1000155. doi: 10.1371/journal.pgen.1000155 . PMC   2491723 . PMID   18704159.
  18. Cuozzo C, Porcellini A, Angrisano T, Morano A, Lee B, Di Pardo A, et al. (July 2007). "DNA damage, homology-directed repair, and DNA methylation". PLOS Genetics. 3 (7): e110. doi: 10.1371/journal.pgen.0030110 . PMC   1913100 . PMID   17616978.
  19. Er TK, Su YF, Wu CC, Chen CC, Wang J, Hsieh TH, et al. (July 2016). "Targeted next-generation sequencing for molecular diagnosis of endometriosis-associated ovarian cancer". Journal of Molecular Medicine. 94 (7): 835–847. doi:10.1007/s00109-016-1395-2. PMID   26920370. S2CID   16399834.
  20. Wang X, Szabo C, Qian C, Amadio PG, Thibodeau SN, Cerhan JR, et al. (February 2008). "Mutational analysis of thirty-two double-strand DNA break repair genes in breast and pancreatic cancers". Cancer Research. 68 (4): 971–975. doi: 10.1158/0008-5472.CAN-07-6272 . PMID   18281469.
  21. Söderlund Leifler K, Queseth S, Fornander T, Askmalm MS (December 2010). "Low expression of Ku70/80, but high expression of DNA-PKcs, predict good response to radiotherapy in early breast cancer". International Journal of Oncology. 37 (6): 1547–1554. doi:10.3892/ijo_00000808. PMID   21042724.
  22. Bouchaert P, Guerif S, Debiais C, Irani J, Fromont G (December 2012). "DNA-PKcs expression predicts response to radiotherapy in prostate cancer". International Journal of Radiation Oncology, Biology, Physics. 84 (5): 1179–1185. doi:10.1016/j.ijrobp.2012.02.014. PMID   22494583.
  23. Zhuang L, Yu SY, Huang XY, Cao Y, Xiong HH (July 2007). "[Potentials of DNA-PKcs, Ku80, and ATM in enhancing radiosensitivity of cervical carcinoma cells]". AI Zheng = Aizheng = Chinese Journal of Cancer (in Chinese). 26 (7): 724–729. PMID   17626748.
  24. Lee SW, Cho KJ, Park JH, Kim SY, Nam SY, Lee BJ, et al. (August 2005). "Expressions of Ku70 and DNA-PKcs as prognostic indicators of local control in nasopharyngeal carcinoma". International Journal of Radiation Oncology, Biology, Physics. 62 (5): 1451–1457. doi:10.1016/j.ijrobp.2004.12.049. PMID   16029807.
  25. Abdel-Fatah TM, Arora A, Moseley P, Coveney C, Perry C, Johnson K, et al. (December 2014). "ATM, ATR and DNA-PKcs expressions correlate to adverse clinical outcomes in epithelial ovarian cancers". BBA Clinical. 2: 10–17. doi:10.1016/j.bbacli.2014.08.001. PMC   4633921 . PMID   26674120.
  26. Lee HS, Yang HK, Kim WH, Choe G (April 2005). "Loss of DNA-dependent protein kinase catalytic subunit (DNA-PKcs) expression in gastric cancers". Cancer Research and Treatment. 37 (2): 98–102. doi:10.4143/crt.2005.37.2.98. PMC   2785401 . PMID   19956487.
  27. Yan D, Ng WL, Zhang X, Wang P, Zhang Z, Mo YY, et al. (July 2010). "Targeting DNA-PKcs and ATM with miR-101 sensitizes tumors to radiation". PLOS ONE. 5 (7): e11397. Bibcode:2010PLoSO...511397Y. doi: 10.1371/journal.pone.0011397 . PMC   2895662 . PMID   20617180.
  28. Li M, Tian L, Ren H, Chen X, Wang Y, Ge J, et al. (August 2015). "MicroRNA-101 is a potential prognostic indicator of laryngeal squamous cell carcinoma and modulates CDK8". Journal of Translational Medicine. 13: 271. doi: 10.1186/s12967-015-0626-6 . PMC   4545549 . PMID   26286725.
  29. Liu Z, Wang J, Mao Y, Zou B, Fan X (January 2016). "MicroRNA-101 suppresses migration and invasion via targeting vascular endothelial growth factor-C in hepatocellular carcinoma cells". Oncology Letters. 11 (1): 433–438. doi:10.3892/ol.2015.3832. PMC   4727073 . PMID   26870229.
  30. Li AY, Boo LM, Wang SY, Lin HH, Wang CC, Yen Y, et al. (July 2009). "Suppression of nonhomologous end joining repair by overexpression of HMGA2". Cancer Research. 69 (14): 5699–5706. doi:10.1158/0008-5472.CAN-08-4833. PMC   2737594 . PMID   19549901.
  31. Motoyama K, Inoue H, Nakamura Y, Uetake H, Sugihara K, Mori M (April 2008). "Clinical significance of high mobility group A2 in human gastric cancer and its relationship to let-7 microRNA family". Clinical Cancer Research. 14 (8): 2334–2340. doi: 10.1158/1078-0432.CCR-07-4667 . PMID   18413822.
  32. Wu A, Wu K, Li J, Mo Y, Lin Y, Wang Y, et al. (March 2015). "Let-7a inhibits migration, invasion and epithelial-mesenchymal transition by targeting HMGA2 in nasopharyngeal carcinoma". Journal of Translational Medicine. 13: 105. doi: 10.1186/s12967-015-0462-8 . PMC   4391148 . PMID   25884389.
  33. Vrba L, Muñoz-Rodríguez JL, Stampfer MR, Futscher BW (2013). "miRNA gene promoters are frequent targets of aberrant DNA methylation in human breast cancer". PLOS ONE. 8 (1): e54398. Bibcode:2013PLoSO...854398V. doi: 10.1371/journal.pone.0054398 . PMC   3547033 . PMID   23342147.
  34. Lee HS, Choe G, Park KU, Park DJ, Yang HK, Lee BL, Kim WH (October 2007). "Altered expression of DNA-dependent protein kinase catalytic subunit (DNA-PKcs) during gastric carcinogenesis and its clinical implications on gastric cancer". International Journal of Oncology. 31 (4): 859–866. doi: 10.3892/ijo.31.4.859 . PMID   17786318.
  35. Shintani S, Mihara M, Li C, Nakahara Y, Hino S, Nakashiro K, Hamakawa H (October 2003). "Up-regulation of DNA-dependent protein kinase correlates with radiation resistance in oral squamous cell carcinoma". Cancer Science. 94 (10): 894–900. doi:10.1111/j.1349-7006.2003.tb01372.x. PMID   14556663. S2CID   2126685.
  36. Riabinska A, Daheim M, Herter-Sprie GS, Winkler J, Fritz C, Hallek M, et al. (June 2013). "Therapeutic targeting of a robust non-oncogene addiction to PRKDC in ATM-defective tumors". Science Translational Medicine. 5 (189): 189ra78. doi:10.1126/scitranslmed.3005814. PMID   23761041. S2CID   206681916.
  37. 1 2 3 Hsu FM, Zhang S, Chen BP (June 2012). "Role of DNA-dependent protein kinase catalytic subunit in cancer development and treatment". Translational Cancer Research. 1 (1): 22–34. doi:10.3978/j.issn.2218-676X.2012.04.01. PMC   3431019 . PMID   22943041.
  38. Espejel S, Martín M, Klatt P, Martín-Caballero J, Flores JM, Blasco MA (May 2004). "Shorter telomeres, accelerated ageing and increased lymphoma in DNA-PKcs-deficient mice". EMBO Reports. 5 (5): 503–509. doi:10.1038/sj.embor.7400127. PMC   1299048 . PMID   15105825.
  39. Reiling E, Dollé ME, Youssef SA, Lee M, Nagarajah B, Roodbergen M, et al. (2014). "The progeroid phenotype of Ku80 deficiency is dominant over DNA-PKCS deficiency". PLOS ONE. 9 (4): e93568. Bibcode:2014PLoSO...993568R. doi: 10.1371/journal.pone.0093568 . PMC   3989187 . PMID   24740260.
  40. Bernstein H, Payne CM, Bernstein C, Garewal H, Dvorak K (2008). Cancer and aging as consequences of un-repaired DNA damage. In: New Research on DNA Damages (Editors: Honoka Kimura and Aoi Suzuki) Nova Science Publishers, Inc., New York, Chapter 1, pp. 1-47. open access, but read only https://www.novapublishers.com/catalog/product_info.php?products_id=43247 Archived 2014-10-25 at the Wayback Machine ISBN   978-1604565812
  41. 1 2 3 4 Kim ST, Lim DS, Canman CE, Kastan MB (December 1999). "Substrate specificities and identification of putative substrates of ATM kinase family members". The Journal of Biological Chemistry. 274 (53): 37538–37543. doi: 10.1074/jbc.274.53.37538 . PMID   10608806.
  42. Suzuki K, Kodama S, Watanabe M (September 1999). "Recruitment of ATM protein to double strand DNA irradiated with ionizing radiation". The Journal of Biological Chemistry. 274 (36): 25571–25575. doi: 10.1074/jbc.274.36.25571 . PMID   10464290.
  43. 1 2 Yavuzer U, Smith GC, Bliss T, Werner D, Jackson SP (July 1998). "DNA end-independent activation of DNA-PK mediated via association with the DNA-binding protein C1D". Genes & Development. 12 (14): 2188–2199. doi:10.1101/gad.12.14.2188. PMC   317006 . PMID   9679063.
  44. Ajuh P, Kuster B, Panov K, Zomerdijk JC, Mann M, Lamond AI (December 2000). "Functional analysis of the human CDC5L complex and identification of its components by mass spectrometry". The EMBO Journal. 19 (23): 6569–6581. doi:10.1093/emboj/19.23.6569. PMC   305846 . PMID   11101529.
  45. 1 2 Goudelock DM, Jiang K, Pereira E, Russell B, Sanchez Y (August 2003). "Regulatory interactions between the checkpoint kinase Chk1 and the proteins of the DNA-dependent protein kinase complex". The Journal of Biological Chemistry. 278 (32): 29940–29947. doi: 10.1074/jbc.M301765200 . PMID   12756247.
  46. Liu L, Kwak YT, Bex F, García-Martínez LF, Li XH, Meek K, et al. (July 1998). "DNA-dependent protein kinase phosphorylation of IkappaB alpha and IkappaB beta regulates NF-kappaB DNA binding properties". Molecular and Cellular Biology. 18 (7): 4221–4234. doi:10.1128/MCB.18.7.4221. PMC   109006 . PMID   9632806.
  47. Wu X, Lieber MR (October 1997). "Interaction between DNA-dependent protein kinase and a novel protein, KIP". Mutation Research. 385 (1): 13–20. doi:10.1016/s0921-8777(97)00035-9. PMID   9372844.
  48. Ma Y, Pannicke U, Schwarz K, Lieber MR (March 2002). "Hairpin opening and overhang processing by an Artemis/DNA-dependent protein kinase complex in nonhomologous end joining and V(D)J recombination". Cell. 108 (6): 781–794. doi: 10.1016/s0092-8674(02)00671-2 . PMID   11955432.
  49. 1 2 Ting NS, Kao PN, Chan DW, Lintott LG, Lees-Miller SP (January 1998). "DNA-dependent protein kinase interacts with antigen receptor response element binding proteins NF90 and NF45". The Journal of Biological Chemistry. 273 (4): 2136–2145. CiteSeerX   10.1.1.615.1747 . doi: 10.1074/jbc.273.4.2136 . PMID   9442054. S2CID   8781571.
  50. Jin S, Kharbanda S, Mayer B, Kufe D, Weaver DT (October 1997). "Binding of Ku and c-Abl at the kinase homology region of DNA-dependent protein kinase catalytic subunit". The Journal of Biological Chemistry. 272 (40): 24763–24766. doi: 10.1074/jbc.272.40.24763 . PMID   9312071.
  51. Matheos D, Ruiz MT, Price GB, Zannis-Hadjopoulos M (October 2002). "Ku antigen, an origin-specific binding protein that associates with replication proteins, is required for mammalian DNA replication". Biochimica et Biophysica Acta (BBA) - Gene Structure and Expression. 1578 (1–3): 59–72. doi:10.1016/s0167-4781(02)00497-9. PMID   12393188.
  52. Gell D, Jackson SP (September 1999). "Mapping of protein-protein interactions within the DNA-dependent protein kinase complex". Nucleic Acids Research. 27 (17): 3494–3502. doi:10.1093/nar/27.17.3494. PMC   148593 . PMID   10446239.
  53. Ko L, Cardona GR, Chin WW (May 2000). "Thyroid hormone receptor-binding protein, an LXXLL motif-containing protein, functions as a general coactivator". Proceedings of the National Academy of Sciences of the United States of America. 97 (11): 6212–6217. Bibcode:2000PNAS...97.6212K. doi: 10.1073/pnas.97.11.6212 . PMC   18584 . PMID   10823961.
  54. Shao RG, Cao CX, Zhang H, Kohn KW, Wold MS, Pommier Y (March 1999). "Replication-mediated DNA damage by camptothecin induces phosphorylation of RPA by DNA-dependent protein kinase and dissociates RPA:DNA-PK complexes". The EMBO Journal. 18 (5): 1397–1406. doi:10.1093/emboj/18.5.1397. PMC   1171229 . PMID   10064605.
  55. Karmakar P, Piotrowski J, Brosh RM, Sommers JA, Miller SP, Cheng WH, et al. (May 2002). "Werner protein is a target of DNA-dependent protein kinase in vivo and in vitro, and its catalytic activities are regulated by phosphorylation". The Journal of Biological Chemistry. 277 (21): 18291–18302. doi: 10.1074/jbc.M111523200 . PMID   11889123.
  56. Goldberg FW, Finlay MR, Ting AK, Beattie D, Lamont GM, Fallan C, Wrigley GL, Schimpl M, Howard MR, Williamson B, Vazquez-Chantada M, Barratt DG, Davies BR, Cadogan EB, Ramos-Montoya A, Dean E (2020). "The Discovery of 7-Methyl-2-[(7-methyl[1,2,4]triazolo[1,5-a]pyridin-6-yl)amino]-9-(tetrahydro-2H-pyran-4-yl)-7,9-dihydro-8H-purin-8-one (AZD7648), a Potent and Selective DNA-Dependent Protein Kinase (DNA-PK) Inhibitor". Journal of Medicinal Chemistry. 63 (7): 3461–3471. doi: 10.1021/acs.jmedchem.9b01684 . PMID   31851518.
  57. "Pharmacologic Inhibitor of DNA-PK, M3814, Potentiates Radiotherapy and Regresses Human Tumors in Mouse Models". Molecular Cancer Therapeutics.
  58. Khan AJ, Misenko SM, Thandoni A, Schiff D, Jhawar SR, Bunting SF, Haffty BG (2018). "VX-984 is a selective inhibitor of non-homologous end joining, with possible preferential activity in transformed cells". Oncotarget. 9 (40): 25833–25841. doi:10.18632/oncotarget.25383. PMC   5995231 . PMID   29899825.
  59. Berger M, Wortmann L, Buchgraber P, Lücking U, Zitzmann-Kolbe S, Wengner AM, Bader B, Bömer U, Briem H, Eis K, Rehwinkel H, Bartels F, Moosmayer D, Eberspächer U, Lienau P, Hammer S, Schatz CA, Wang Q, Wang Q, Mumberg D, Nising CF, Siemeister G (2021). "BAY-8400: A Novel Potent and Selective DNA-PK Inhibitor which Shows Synergistic Efficacy in Combination with Targeted Alpha Therapies". Journal of Medicinal Chemistry. 64 (17): 12723–12737. doi: 10.1021/acs.jmedchem.1c00762 . PMID   34428039.