Open Access

Dual roles of SAMHD1 in tumor development and chemoresistance to anticancer drugs (Review)

  • Authors:
    • Zhangming Chen
    • Jie Hu
    • Songcheng Ying
    • Aman Xu
  • View Affiliations

  • Published online on: April 8, 2021     https://doi.org/10.3892/ol.2021.12712
  • Article Number: 451
  • Copyright: © Chen et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Human sterile alpha motif and HD‑domain‑­containing protein 1 (SAMHD1) has been identified as a GTP or dGTP‑dependent deoxynucleotide triphosphohydrolase (dNTPase) and acts as an antiviral factor against certain retroviruses and DNA viruses. Genetic mutation in SAMHD1 causes the inflammatory Aicardi‑Goutières Syndrome and abnormal intracellular deoxyribonucleoside triphosphates (dNTPs) pool. At present, the role of SAMHD1 in numerous types of cancer, such as chronic lymphocytic leukemia, lung cancer and colorectal cancer, is highly studied. Furthermore, it has been found that methylation, acetylation and phosphorylation are involved in the regulation of SAMHD1 expression, and that genetic mutations can cause changes in its activities, including dNTPase activity, long interspersed element type 1 (LINE‑1) suppression and DNA damage repair, which could lead to uncontrolled cell cycle progression and cancer development. In addition, SAMHD1 has been reported to have a negative regulatory role in the chemosensitivity to anticancer drugs through its dNTPase activity. The present review aimed to summarize the regulation of SAMHD1 expression in cancer and its function in tumor growth and chemotherapy sensitivity, and discussed controversial points and future directions.

Introduction

Human sterile α motif and HD domain-containing protein 1 (SAMHD1), also known as dendritic cell derived IFN-γ induced protein, was first identified as interferon (IFN)-γ induced protein in human dendritic cells (1). SAMHD1 has an N-terminal nuclear localization signal (1–150), which contains the transferable sequence (11KRPR14) (2), an N-terminal SAM domain, which interacts with proteins, and an HD domain, which is responsible for the hydrolysis of deoxyribonucleoside triphosphates (dNTPs) (3). SAMHD1 is widely expressed in most human tissues and cell lines (1). Rice et al (4) reported 14 mutations in exons of SAMHD1 gene in patients with Aicardi-Goutières syndrome (AGS), resulting in alterations of the amino acid sequence (4). Mass spectrometry results have demonstrated that SAMHD1 interacts with Vpx, which induces the degradation of SAMHD1 via the ubiquitin-proteasome pathway (5). Previous studies have reported that SAMHD1 can block the replication of human immunodeficiency virus type 1 (HIV-1) by hydrolyzing dNTPs, which depletes the pool of dNTPs available for reverse transcription essential to DNA replication of HIV-1 (6), hepatitis B virus (HBV) (7,8) and herpes simplex virus 1 (9).

The maintenance of an appropriate intracellular dNTPs pool is essential for cell cycle progression, and an imbalance in the pool of dNTPs can enhance mutagenesis and DNA replication, resulting in dysregulation of the cell cycle (10). SAMHD1 is the first dNTP triphosphohydrolase (dNTPase) discovered in mammalian cells and is crucial in maintaining the homeostasis of intracellular dNTPs (6). dGTP-Mg2+-dGTP binding at the allosteric site promotes the tetramerization of SAMHD1 protein, leading to the formation of catalytically active enzymes (11). A previous study demonstrated that SAMHD1 has an essential role in regulating the cell cycle in quiescent and cycling cells via regulation of the intracellular concentration of dNTPs (12). In addition, genetic mutations of SAMHD1 have been discovered in several types of cancer, including chronic lymphocytic leukemia (CLL) (13), lung cancer (14) and colorectal cancer (15). Methylation at the promoter of SAMHD1 has been demonstrated to decrease its mRNA and protein expression (14), and the functions of SAMHD1 are affected by acetylation (16) and phosphorylation (17). In particular, SAMHD1 is involved in DNA damage repair (DDR) (18), long interspersed element type 1 (LINE-1) suppression (19,20) and chemosensitivity to anticancer drugs (21,22). In addition, genetic mutations of SAMHD1 contribute to amino acid changes, which in part impair its functions (11,15).

The present review aimed to summarize the underlying mechanisms of SAMHD1 in cancer development and chemosensitivity to anticancer drugs, and discussed numerous controversial hypotheses.

Role of methylation, acetylation and phosphorylation in the regulation of SAMHD1 expression in cancer

The transcriptional regulation of SAMHD1 is the result of methylation of its promoter. This has been observed in several types of tumor tissue and cell line. Human SAMHD1 is expressed at different levels in cell lines and tissues (1). de Silva et al (23) reported that the protein expression of SAMHD1 is correlated with its mRNA expression level in THP-1 and CD4+ T cells. In peripheral blood mononuclear cells from patients with cutaneous T-cell lymphoma (CTCL), both the mRNA and protein expression levels are significantly lower compared with those in healthy donors (24), which was also observed in lung cancer tissues compared with adjacent normal tissues (14). The different expression levels of SAMHD1 mRNA and protein in various types of cancer are summarized in Table I. The level of methylation has also been demonstrated to be inversely associated with the expression of SAMHD1 in 466 skin cutaneous melanoma samples from The Cancer Genome Atlas (TCGA) (25). Taken together, these findings indicate that epigenetic regulation plays an important role in the protein expression of SAMHD1. Results from methylation-specific PCR demonstrated that the SAMHD1 promoter is methylated in lung adenocarcinoma tissues compared with adjacent normal tissues (14). Furthermore, in patients with CTCL, methylated SAMHD1 has also been observed whereas it was absent in healthy donors (24). The DNA methylation inhibitor 5-aza-2′-deoxycytidine (5-AzadC) has been reported to increase the mRNA and protein expression levels of SAMHD1 in the lung cancer cell line A549 (14), and Jurkat and Sup-T1 cells (acute T-cell leukemia) (24). Conversely, the promoter of SAMHD1 has been demonstrated to be in an unmethylated state in THP-1 cells (23) and SAMHD1 protein expression does not increase following treatment with 5-AzadC. Furthermore, decitabine, which is a DNA methyltransferase inhibitor, has been demonstrated to be a substrate of SAMHD1, and its anticancer effect in acute myelocytic leukemia (AML) was found to be dependent of SAMHD1 protein expression (26). For other tumors characterized by methylation of the SAMHD1 promoter, further investigation is required to determine whether SAMHD1 could have a two-sided effect on chemotherapy with decitabine.

Table I.

SAMHD1 expression and its association with the chemosensitivity of anticancer drugs in various types of cancer.

Table I.

SAMHD1 expression and its association with the chemosensitivity of anticancer drugs in various types of cancer.

mRNAProteinCancer typeCell typeFunctionDrug ChemosensitivityReferences
DownregulatedDownregulatedChronic lymphocyticHeLaInhibitedCamptothecin,IncreasedClifford et al (13)
leukemia proliferationmitomycin C, etoposide
DownregulatedDownregulatedCutaneous T-cellHuT78Inhibitedara-CDecreasedde Silva et al (24)
lymphoma proliferation and stimulated apoptosis Herold et al (21)
DownregulatedNDAcute myeloidTHP-1,Inhibitedara-C;DecreasedJiang et al (74)
leukemiaOCI-AML3,Proliferatiovidarabine; Bonifati et al (42)
HELn and cellnelarabine; Kodigepalli et al (45,75)
transformationfludarabine; Schneider et al (59)
decitabine; Herold et al (21,22,76)
trifluridine; Knecht et al (61)
guadecitabine Oellerich et al (26)
DownregulatedDownregulatedT-cell acute lymphoblastic leukemiaMHH-CALL-4NDAra-GDecreasedRothenburger et al (77)
UpregulatedNDColorectal Cancer NDNDNDYang et al (78)
UpregulatedNDSerum of lung cancer patientsNDNDNDNDShang et al (79)
DownregulatedDownregulatedOropharyngeal cancerHPV16-positive keratinocytesInhibited proliferationNDNDJames et al (68)
DownregulatedNDSkin cutaneous melanomaNDNDNDNDChen et al (25)
NDNDHepatocellular carcinomaHepG2 CisplatinIncreasedShi et al (80)
DownregulatedDownregulatedLung cancerA549; H1299Suppressed proliferationNDNDWang et al (14)

[i] ND, not determined; NE, no effect; ara-C, cytosine arabinoside; ara-G, guanine arabinoside; HPV, human papillomavirus.

Acetylation has been reported to be involved in the post-transcriptional regulation of SAMHD1 expression. Trichostatin A (TSA), a histone deacetylase inhibitor, can increase the mRNA and protein expression of SAMHD1 in Jurkat and Sup-T1 cells (23). Lee et al (16) reported that SAMHD1 is acetylated at the K405 site via arrest defective protein 1 (ARD1), which leads to enhanced dNTPase activity (16). Furthermore, SAMHD1 and ARD1 protein expression has been found to be upregulated in hepatoma tissues compared with non-tumor tissues, which indicates that SAMHD1 acetylated by ARD1 may participate in tumorigenesis (16). However, the K580 residue of SAMHD1, which is another acetylated site, could not be acetylated by ARD1 in vitro (16). In particular, 5-AzadC and TSA were found to notably increase the mRNA expression of SAMHD1; however, only a slight increase in SAMHD1 protein expression was observed in Jurkat and Sup-T1 cells (23). Subsequently, besides methylation and acetylation, another regulatory mechanism may exist in SAMHD1 transcription or translation.

Phosphorylated SAMHD1 at the Thr592 site was first reported in 2013 (27). Another study revealed that SAMHD1 protein was phosphorylated at four major sites, including Thr592, Thr21, Ser6 and Ser33 (28). White et al (27) demonstrated that the expression of phosphorylated SAMHD1 at Ser33 site is equal in cycling and non-cycling U937 cells (27). It was reported that Thr592 has the largest effect on SAMHD1 phosphorylation, whereas mutation of Ser6 does not appear to affect SAMHD1 phosphorylation (28). The S phase is initiated by cyclin E/cyclin-dependent kinase (CDK) 2, cyclin A/CDK1 and cyclin A/CDK2 complexes, which are essential for the transition from S phase to M phase (29). The (S/T)PX(K/R) is a well-defined consensus sequence for the phosphorylation site in the amino acid sequence of a CDK substrate (30). SAMHD1 wild-type protein contains a sequence 592TPQK595, which could be the substrate for CDKs. Further in vitro experiments have shown that the C-terminal domain of SAMHD1 is the target sequence for the cyclin A2/CDK1 complex, which could not phosphorylate SAMHD1 mutant at the 592 site (Fig. 1) (31,32). Furthermore, the expression of SAMHD1 and phosphorylated SAMHD1 differed in the S phase compared with other cell cycle phases (33), indicating that phosphorylation modification of SAMHD1 by cyclin-CDKs complex is critical for transition of G1/G0 to S phase. In addition, phosphorylation of SAMHD1 can also be suppressed by the CDK2 inhibitor II in hepatoma cells (34). Palbociclib can also significantly decrease SAMHD1 phosphorylation in monocyte-derived macrophages and CD4+ T lymphocytes by downregulating the CDK6 downstream CDK2 phosphorylation at Thr160 (35,36). Furthermore, a previous study revealed that phosphorylated SAMHD1 appears at the G1/S border and lasts until mitosis, and could interact directly with cyclin E and cyclin A2 at its C terminus end (17). However, it remains unclear whether SAMHD1 mutations at the C terminus could attenuate the anticancer effect of CDK inhibitors. Depletion of SAMHD1 promotes G1/G0 phase arrest and cell senescence in HeLa cells, which is associated with an increased expression of p21 (32). Conversely, p21 has been demonstrated to have no effect on the expression of total SAMHD1 protein (37) but can reverse CDK-induced phosphorylation of SAMHD1 by taking on the role of CDK inhibitor (Fig. 1) (38). Ubiquitin-like specific protease 18 (USP18) suppresses the expression of p21 protein, leading to phosphorylation of SAMHD1 (Fig. 1) (39). IFN-β can also inhibit the phosphorylation of SAMHD1, but not decrease total SAMHD1 expression (39). Protein phosphatase 2 (PP2A)-B55α holoenzymes can interact with and dephosphorylate SAMHD1 at T592 in cycling cells (40). Repression of CDK1/2 via aryl hydrocarbon receptor activation has been demonstrated to be responsible for SAMHD1-mediated hydrolysis of dNTPs in macrophages (41). Taken together, these results suggested that SAMHD1 phosphorylation is regulated by CDKs-interacting proteins.

SAMHD1 inhibits the development of numerous types of tumor

SAMHD1 can inhibit cancer development through its dNTPase activity

In 2011, SAMHD1 was identified as a dNTPase that hydrolyzes cellular dNTPs (6). The expression of SAMHD1 in quiescent fibroblasts was found to be higher than that in cycling fibroblasts, and inversely associated with the percentage of lung fibroblasts in the S phase (12). Bonifati et al (42) reported that, compared with control groups, SAMHD1 knockdown can increase the concentration of dNTPs in THP-1 cells, enhance cell proliferation and decrease cell apoptosis (42). Furthermore, overexpression of SAMHD1 reduces the proliferation of HeLa (cervical cancer), A549 (lung cancer) and HuT78 (CTCL) cells (13,14,43), which is accompanied by increased cell apoptosis via activation of the caspase-3 pathway (43). Mutations of SAMHD1 gene at critical sites can cause changes in dNTPase activity (Fig. 1), which is summarized in Table II. Welbourn et al (44) discovered that SAMHD1 mutant H206A/D207A is disabled to hydrolyze the intracellular dNTPs in HeLa cells (44). Furthermore, mutations of SAMHD1 at catalysis-related sites, including K312A, Y315A, R366A, H370A and Y374G, lead to decreased dNTPase activity (11). Eight heterozygous mutations of SAMHD1 were identified in colorectal cancer according to TCGA database, including D497Y, V133I(X2), A388V, A525T, R366H, K596fs and A388T (15). The dNTPase activity of the mutations V133I, A338T and R366H is weakened to various degrees and is completely lost in the D497Y mutation compared with the wild-type SAMHD1 protein (15). However, the growth rate of THP-1 ×enograft tumors following SAMHD1 knockdown was reported to be decreased compared with the control group in vivo, which differs from the in vitro findings (45).

Table II.

The variants of SAMHD1 and its role to chemosensitivity of drugs in cancers.

Table II.

The variants of SAMHD1 and its role to chemosensitivity of drugs in cancers.

MutantsCancer tissuesFunctionDrug ChemosensitivityReferences
Insertion in exon4Hepatocellular carcinomaNDCisplatinIncreasedShi et al (80)
Deletions of exons8-9, 13 and 14Hepatocellular carcinomaNDCisplatinNEShi et al (80)
F545L, M1K, C522X, Y155C, L431S, D501Y, R371H, R145Q, W572X, L493R, H206R, I201N, E355K, R145X, T365P, I300L, P158S, D16Y, L244F, R290C, R451C, R451C, V500GChronic lymphocytic leukemiaNDNDNDClifford et al (13)
c.1411-2A> GCutaneous T-cell lymphomaNDNDNDMerati et al (81)
c.646_647delATChronicNDNDNDAmin et al (82)
c.299T>Clymphocytic
c.1352G>Aleukemia
c.503G>A
c.1562A>G
c.773G>A
c.1324C>T
c.1181A>T
c.646_647delAT
c.658C>T
D497Y, V133I(×2), A388V, A525T, R366H, K596fs A388TColon cancerReduced dNTPase activityNDNDRentoft et al (15)
K484TGastric cancerImpaired interaction of SAMHD1 with CtBP-interacting proteinCamptothecinDecreasedDaddacha et al (55)
rs6102991Adult acute myeloid leukemiaDecreased SAMHD1 mRNA expression Decreased risk of non-complete remissionNDNDZhu et al (83)

[i] ND, not determined; NE, no effect; SAMHD1, sterile a motif and HD domain-containing protein 1.

SAMHD1 can suppress LINE-1 retrotransposon in various types of cancer

LINE-1 is the only active autonomous retrotransposon in the human genome and has non-long terminal repeats, which can encode its mRNA and two open reading frame (ORF) proteins, ORF1p and ORF2p (46,47). In vitro, SAMHD1 has no effect on the expression of LINE-1 mRNA and ORF1p protein but suppresses expression of ORF2p (19). In addition, a previous study has reported more mutants of SAMHD1 protein in various types of cancer (Fig. 2) (20). Furthermore, SAMHD1 single-nucleotide polymorphisms (SNPs) at seven codon positions negatively regulate the retrotransposition of LINE-1 compared with its wild-type equivalent (48). Furthermore, SAMHD1 mutation deficiency of amino acids 162–335 attenuates the activity suppressing LINE-1; however, the mutant lacking SAM domain enhances its ability against LINE-1 (19). The enzymatically active HD domain of SAMHD1, SAM domain, is therefore crucial for the inhibition of LINE-1 retrotransposon. The dNTPase-inactive SAMHD1 mutants H206A/D207A and K288T can also inhibit the reverse transcription of LINE-1 in colon cancer (20,48). Removing the nuclear localization sequence of SAMHD1 or its K11A mutants could subsequently induce its redistribution in the cytoplasm, leading to ORF2p deletion and LINE-1 suppression, but with no effect on dNTPase activity (49,50). Therefore, SAMHD1-mediated LINE-1 inhibition is independent of its dNTPase activity. In addition, the activity of SAMHD1 against LINE-1 was found to be negatively regulated by phosphorylation at Thr592 (51). Furthermore, exogenous expression of wild-type SAMHD1, instead of H233A mutant, induces abnormal distribution of ORF1p and promotes the formation of large cytoplasmic foci, which suppresses LINE-1 retrotransposon dependent of the expression of stress granule markers (Ras GTPase-activating protein-binding protein 1 and Tia1 cytotoxic granule-associated RNA binding protein) (52). Interaction between SAMHD1 and ORF1p has not yet been discovered, and the underlying mechanism of SAMHD1-induced redistribution of ORF1p into stress granules requires further investigation.

SAMHD1 can promote DDR following irradiation and chemotherapy

RNA ribosomal 2-mediated dNTP biosynthesis and SAMHD1-mediated dNTP hydrolysis are two different strategies used to maintain dNTP homeostasis in mammalian cells (6,53) and are essential for DDR and cell survival. In addition, DNA damage also increases dNTPs pools through upregulating the activity of ribonucleotide reductase (53). Etoposide, an anticancer drug inducing DNA double-strand breaks, can also promote the dephosphorylation of SAMHD1 and the transition from G1 to G0 phase in the cell cycle, which is accompanied by a decrease in the protein expression of CDK1 (54). The cyclin A2/CDK1 complex interacts with and phosphorylates SAMHD1 at the Thr592 site (31). Subsequently, etoposide-induced dephosphorylation of SAMHD1 may result from etoposide-mediated inhibition of CDK1 when cells return from the G1 phase into the G0 phase. DNA synthesis is impeded by SAMHD1 mutants lacking the C-terminal end (618–626 residue), which is accompanied by the phosphorylation of checkpoint kinase 1 (CHK1) at Ser345 and activation of DNA damage checkpoints (33). A previous study reported that treatment with neocarzinostatin increases p21 expression, which inhibits CDK1/2, leading therefore to decreased phosphorylation of SAMHD1 (54). Dephosphorylation of SAMHD1 may thus be the result of increased p21 in DNA damage induced by ultraviolet light or neocrazinostatin (Fig. 1). Another study reported that SAMHD1 knockdown can increase U2OS (osteosarcoma), MCF7 (breast cancer) and HCT-116 (colon cancer) cell sensitivity to etoposide, camptothecin and ionizing radiation, which can be rescued by restoring the exogenous expression of SAMHD1 (55). Following treatment with camptothecin, immunofluorescence shows that SAMHD1 is colocalized with gH2A histone family member X and RAD51 in DNA damage sites, which are markers of DNA damage and homologous recombination, respectively (55). Furthermore, SAMHD1 is colocalized with p53-binding protein 1 foci, which is an indispensable regulatory protein involved in DDR, following camptothecin-induced DNA double-strand breaks (13). SAMHD1 has been observed at replication forks and interacts with double-strand break repair protein MRE11 exonuclease, which is followed by activation of the CHK1-ataxia telangiectasia and Rad3-related protein checkpoint and restart of the replication forks, eventually leading to acceleration of the normal fork progression, which could be restored by T592E (phosphorylation) and K312A (dNTPase-inactive) mutant of SAMHD1, but not by the T592A (non-phosphorylation) mutant (18). In addition, SAMHD1 can interact with and recruit CtBP-interacting protein to DNA damage sites, followed by promotion of DNA end resection, which could not be rescued by the gastric cancer-related K484T mutant of SAMHD1- (55). Further experiment shows that K484T mutant has no dNTPase activity and can not hydrolyze dNTPs (55). Therefore, SAMHD1 can promote DDR in cancer cells in response to DNA double-strand break-inducing agents through a phosphorylation-dependent mechanism at the Thr592 site rather than a dNTPase-dependent mechanism.

SAMHD1 is a negative regulator of nucleotide analogs in vitro and in vivo

At present, antiviral agents are nucleotide analogs that incorporate into DNA and impede DNA synthesis in the virus replication cycle. Ballana et al (56) demonstrated for the first time that degradation of SAMHD1 via Vpx can decrease the sensitivity of thymidine analogs against HIV-1, including zidovudine and stavudine (Fig. 3). Based on the understanding of dNTPase activity of SAMHD1, dNTP-based metabolite from anticancer drugs became the focus in numerous studies on anticancer drug chemosensitivity. Mutations in SAMHD1 have been found in 9.6% of patients with relapsed or refractory CLL (57) and were demonstrated to interact with proteins and implicated in DNA double-strand breaks (13). The antimetabolite cytosine arabinoside (ara-C) blocks DNA polymerase and gets incorporated into elongated DNA in the S phase of the cell cycle, leading to reduced DNA replication and eventually cell death (58). Herold et al (21) revealed that intracellular ara-CTP, a dCTP analog, is a substrate of SAMHD1 in THP-1 and HuT-78 cells. Results from high-performance liquid-chromatography demonstrated that ara-CTP is hydrolyzed to ara-C and inorganic triphosphate by SAMHD1 in the presence of allosteric activators (59). Furthermore, it was also confirmed that clofarabine is a substrate of SAMHD1 (60). In contrast to the research by Ballana et al (56), Vpx-mediated SAMHD1 degradation enhances the sensitivity of ara-CTP in AML and cutaneous T-cell lymphoma cells (21). Another study by Herold et al (22) reported that SAMHD1 could also hydrolyze triphosphates of other nucleotide analogs, such as vidarabine, nelarabine, fludarabin, decitabine and trifluridine (22), whereas gemcitabine and 6-thioguanine are neither substrates nor activators of SAMHD1 (21,61). In a mouse model of AML, xenograft tumors with SAMHD1 knockdown presented an improved response to ara-C chemotherapy compared with those expressing SAMHD1 (21). The chemosensitivity of ara-C was also found to be negatively regulated by SAMHD1 in patient-derived AML blasts (21). In 143 patients with AML who received ara-C-based chemotherapy, patients with low SAMHD1 expression levels had higher event-free survival rates than those with high SAMHD1 expression levels (37 vs. 19%, respectively), as well as higher 5-year overall survival rates (54 vs. 9%, respectively) (62). Furthermore, AML cells resistant to decitabine exhibit increased SAMHD1 protein expression, which could be a result of SAMHD1 promoter demethylation (26). In addition, SAMHD1 may also be considered as a useful biomarker for predicting the response to decitabine. Apart from its role in hematological malignancies, it is also unclear whether SAMHD1 has a crucial predictive function for the efficacy of nucleotide analogs in solid tumors. Wang et al (63) first reported that U251 (glioblastoma) cells with acquired resistance to olaparib, talazoparib and simmiparib, exhibited upregulated expression levels of SAMHD1 mRNA and protein than that in parental cell lines. Furthermore, U251 cells with resistance to olaparib displayed similar cross-resistance to 5-fluorouracil, and the IC50 was increased by >3-fold (63). However, it remains unclear whether upregulation of SAMHD1 expression is responsible for the chemoresistance of other solid malignancies to 5-fluorouracil. The structure and biochemical framework established by Knecht et al (61) is useful to determine novel anticancer drugs that could be hydrolyzed by SAMHD1. Ribonucleotide reductase subunit R2 inhibitors (gemcitabine, hydroxyurea and triapine) occupy the AS2 site and decrease the ara-CTPase of SAMHD1, leading to enhanced chemosensitivity of AML cells to ara-C (64). Hollenbaugh et al (65) demonstrated that Y374 in the catalytic pocket of SAMHD1 could be occupied by the (2′S)-2′-methyl group of SMDU-TP, resulting in decreased dGTP hydrolysis. Further investigation is therefore needed to determine whether a combination of SAMHD1 inhibitors and nucleotide analogs could improve chemosensitivity in the treatment of cancer.

Future directions and conclusion

Numerous variants of SAMHD1 have been reported in hematological (13) and solid tumors, including colon cancer (15), lung cancer (14), pancreatic cancer (66) and glioblastoma (67); however, their roles in tumorigenesis and cancer development remain to be elucidated. Deletion of SAMHD1 can accelerate the proliferation of keratinocytes induced by HPV-6, which is one of the contributing factors in the development of human oropharyngeal cancer (68). In turn, HPV-6 has also been found to inhibit the endogenous expression of SAMHD1 (68). It would therefore be crucial to verify whether SAMHD1 may serve a role in other virus-related tumors, such as Epstein-Barr virus-associated gastric cancer (69) and in the development of cervical cancer due to HPV infection (70). SAMHD1 is involved in homologous recombination in response to DNA damage, which only occurs in the S phase (55,71). However, SAMHD1 is expressed in the whole cell cycle and, thus, the function and mechanism of SAMHD1 in non-homologous recombination is worth further investigation. It is widely acknowledged that reduced dNTPase activity of SAMHD1 is the result of gene mutations and loss of expression, which leads to persistently high levels of dNTPs and abnormal cell proliferation in several types of tumor (13,14,43). Numerous studies have reported that SAMHD1 can interact with cyclin/CDK complexes (30), USP18 (39) and S-phase kinase-associated protein 2 (39), which are involved in the regulation of cell proliferation (30,39). However, the underlying mechanisms of SAMHD1 in the induction and regulation of tumorigenesis remain unknown, and it is hypothesized that SAMHD1 may mediate cell proliferation via dNTPase-dependent or -independent mechanisms. Previous studies have demonstrated that SAMHD1 is involved in regulating the therapeutic efficacy of nucleotide analogs in AML, which should also be investigated in solid tumors. Apart from retrospective studies, prospective cohort studies could provide more reliable evidence for this. Previous studies have also demonstrated that inhibitors of SAMHD1 dNTPase activity, including cephalosporin C zinc salt (72), dGTPαS, lomofungin, L-thyroxine, ergotamine, amrinone, retinoic acid, montelukast, hexestrol and sulindac (73), may be considered as potential sensitizers in anticancer therapy. In addition, modification of anticancer nucleotide analogs to eliminate SAMHD1-mediated chemoresistance may also be a useful strategy in the treatment of cancer. In conclusion, current evidence has demontrated that SAMHD1 has some dual effects in cancer, including some roles in tumor suppression and some roles in decreasing anticancer drug chemosensitivity.

Acknowledgements

Not applicable.

Funding

This work was supported by the National Natural Science Foundation of China (grant no. 81572350).

Availability of data and materials

Not applicable.

Authors' contributions

ZC and JH contributed to manuscript draft and figure design. ZC, SY and AX provided critical revision of the manuscript. Data sharing is not applicable. All authors read and approved the final manuscript.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

References

1 

Li N, Zhang W and Cao X: Identification of human homologue of mouse IFN-gamma induced protein from human dendritic cells. Immunol Lett. 74:221–224. 2000. View Article : Google Scholar : PubMed/NCBI

2 

Brandariz-Nuñez A, Valle-Casuso JC, White TE, Laguette N, Benkirane M, Brojatsch J and Diaz-Griffero F: Role of SAMHD1 nuclear localization in restriction of HIV-1 and SIVmac. Retrovirology. 9:492012. View Article : Google Scholar

3 

Wei W, Guo H, Han X, Liu X, Zhou X, Zhang W and Yu XF: A novel DCAF1-binding motif required for Vpx-mediated degradation of nuclear SAMHD1 and Vpr-induced G2 arrest. Cell Microbiol. 14:1745–1756. 2012. View Article : Google Scholar : PubMed/NCBI

4 

Rice GI, Bond J, Asipu A, Brunette RL, Manfield IW, Carr IM, Fuller JC, Jackson RM, Lamb T, Briggs TA, et al: Mutations involved in Aicardi-Goutieres syndrome implicate SAMHD1 as regulator of the innate immune response. Nat Genet. 41:829–832. 2009. View Article : Google Scholar : PubMed/NCBI

5 

Laguette N, Sobhian B, Casartelli N, Ringeard M, Chable-Bessia C, Ségéral E, Yatim A, Emiliani S, Schwartz O and Benkirane M: SAMHD1 is the dendritic- and myeloid-cell-specific HIV-1 restriction factor counteracted by Vpx. Nature. 474:654–657. 2011. View Article : Google Scholar : PubMed/NCBI

6 

Goldstone DC, Ennis-Adeniran V, Hedden JJ, Groom HC, Rice GI, Christodoulou E, Walker PA, Kelly G, Haire LF, Yap MW, et al: HIV-1 restriction factor SAMHD1 is a deoxynucleoside triphosphate triphosphohydrolase. Nature. 480:379–382. 2011. View Article : Google Scholar : PubMed/NCBI

7 

Chen Z, Zhu M, Pan X, Zhu Y, Yan H, Jiang T, Shen Y, Dong X, Zheng N, Lu J, et al: Inhibition of Hepatitis B virus replication by SAMHD1. Biochem Biophys Res Commun. 450:1462–1468. 2014. View Article : Google Scholar : PubMed/NCBI

8 

Jeong GU, Park IH, Ahn K and Ahn BY: Inhibition of hepatitis B virus replication by a dNTPase-dependent function of the host restriction factor SAMHD1. Virology. 495:71–78. 2016. View Article : Google Scholar : PubMed/NCBI

9 

Kim ET, White TE, Brandariz-Núñez A, Diaz-Griffero F and Weitzman MD: SAMHD1 restricts herpes simplex virus 1 in macrophages by limiting DNA replication. J Virol. 87:12949–12956. 2013. View Article : Google Scholar : PubMed/NCBI

10 

Buj R and Aird KM: Deoxyribonucleotide Triphosphate Metabolism in Cancer and Metabolic Disease. Front Endocrinol (Lausanne). 9:1772018. View Article : Google Scholar : PubMed/NCBI

11 

Ji X, Wu Y, Yan J, Mehrens J, Yang H, DeLucia M, Hao C, Gronenborn AM, Skowronski J, Ahn J and Xiong Y: Mechanism of allosteric activation of SAMHD1 by dGTP. Nat Struct Mol Biol. 20:1304–1309. 2013. View Article : Google Scholar : PubMed/NCBI

12 

Franzolin E, Pontarin G, Rampazzo C, Miazzi C, Ferraro P, Palumbo E, Reichard P and Bianchi V: The deoxynucleotide triphosphohydrolase SAMHD1 is a major regulator of DNA precursor pools in mammalian cells. Proc Natl Acad Sci USA. 110:14272–14277. 2013. View Article : Google Scholar : PubMed/NCBI

13 

Clifford R, Louis T, Robbe P, Ackroyd S, Burns A, Timbs AT, Wright Colopy G, Dreau H, Sigaux F, et al: SAMHD1 is mutated recurrently in chronic lymphocytic leukemia and is involved in response to DNA damage. Blood. 123:1021–1031. 2014. View Article : Google Scholar : PubMed/NCBI

14 

Wang JL, Lu FZ, Shen XY, Wu Y and Zhao LT: SAMHD1 is down regulated in lung cancer by methylation and inhibits tumor cell proliferation. Biochem Biophys Res Commun. 455:229–233. 2014. View Article : Google Scholar : PubMed/NCBI

15 

Rentoft M, Lindell K, Tran P, Chabes AL, Buckland RJ, Watt DL, Marjavaara L, Nilsson AK, Melin B, Trygg J, et al: Heterozygous colon cancer-associated mutations of SAMHD1 have functional significance. Proc Natl Acad Sci USA. 113:4723–4728. 2016. View Article : Google Scholar : PubMed/NCBI

16 

Lee EJ, Seo JH, Park JH, Vo TTL, An S, Bae SJ, Le H, Lee HS, Wee HJ, Lee D, et al: SAMHD1 acetylation enhances its deoxynucleotide triphosphohydrolase activity and promotes cancer cell proliferation. Oncotarget. 8:68517–68529. 2017. View Article : Google Scholar : PubMed/NCBI

17 

Tramentozzi E, Ferraro P, Hossain M, Stillman B, Bianchi V and Pontarin G: The dNTP triphosphohydrolase activity of SAMHD1 persists during S-phase when the enzyme is phosphorylated at T592. Cell Cycle. 17:1102–1114. 2018. View Article : Google Scholar : PubMed/NCBI

18 

Coquel F, Silva MJ, Técher H, Zadorozhny K, Sharma S, Nieminuszczy J, Mettling C, Dardillac E, Barthe A, Schmitz AL, et al: SAMHD1 acts at stalled replication forks to prevent interferon induction. Nature. 557:57–61. 2018. View Article : Google Scholar : PubMed/NCBI

19 

Zhao K, Du J, Han X, Goodier JL, Li P, Zhou X, Wei W, Evans SL, Li L, Zhang W, et al: Modulation of LINE-1 and Alu/SVA retrotransposition by Aicardi-Goutieres syndrome-related SAMHD1. Cell Rep. 4:1108–1115. 2013. View Article : Google Scholar : PubMed/NCBI

20 

Gao W, Li G, Bian X, Rui Y, Zhai C, Liu P, Su J, Wang H, Zhu C, Du Y, et al: Defective modulation of LINE-1 retrotransposition by cancer-associated SAMHD1 mutants. Biochem Biophys Res Commun. 519:213–219. 2019. View Article : Google Scholar : PubMed/NCBI

21 

Herold N, Rudd SG, Ljungblad L, Sanjiv K, Myrberg IH, Paulin CB, Heshmati Y, Hagenkort A, Kutzner J, Page BD, et al: Targeting SAMHD1 with the Vpx protein to improve cytarabine therapy for hematological malignancies. Nat Med. 23:256–263. 2017. View Article : Google Scholar : PubMed/NCBI

22 

Herold N, Rudd SG, Sanjiv K, Kutzner J, Bladh J, Paulin CBJ, Helleday T, Henter JI and Schaller T: SAMHD1 protects cancer cells from various nucleoside-based antimetabolites. Cell Cycle. 16:1029–1038. 2017. View Article : Google Scholar : PubMed/NCBI

23 

de Silva S, Hoy H, Hake TS, Wong HK, Porcu P and Wu L: Promoter methylation regulates SAMHD1 gene expression in human CD4+ T cells. J Biol Chem. 288:9284–9292. 2013. View Article : Google Scholar : PubMed/NCBI

24 

de Silva S, Wang F, Hake TS, Porcu P, Wong HK and Wu L: Downregulation of SAMHD1 expression correlates with promoter DNA methylation in Sezary syndrome patients. J Invest Dermatol. 134:562–565. 2014. View Article : Google Scholar : PubMed/NCBI

25 

Chen W, Cheng P, Jiang J, Ren Y, Wu D and Xue D: Epigenomic and genomic analysis of transcriptome modulation in skin cutaneous melanoma. Aging (Albany NY). 12:12703–12725. 2020. View Article : Google Scholar : PubMed/NCBI

26 

Oellerich T, Schneider C, Thomas D, Knecht KM, Buzovetsky O, Kaderali L, Schliemann C, Bohnenberger H, Angenendt L, Hartmann W, et al: Selective inactivation of hypomethylating agents by SAMHD1 provides a rationale for therapeutic stratification in AML. Nat Commun. 10:34752019. View Article : Google Scholar : PubMed/NCBI

27 

White TE, Brandariz-Nuñez A, Valle-Casuso JC, Amie S, Nguyen LA, Kim B, Tuzova M and Diaz-Griffero F: The retroviral restriction ability of SAMHD1, but not its deoxynucleotide triphosphohydrolase activity, is regulated by phosphorylation. Cell Host Microbe. 13:441–451. 2013. View Article : Google Scholar : PubMed/NCBI

28 

Welbourn S, Dutta SM, Semmes OJ and Strebel K: Restriction of virus infection but not catalytic dNTPase activity is regulated by phosphorylation of SAMHD1. J Virol. 87:11516–11524. 2013. View Article : Google Scholar : PubMed/NCBI

29 

Hochegger H, Takeda S and Hunt T: Cyclin-dependent kinases and cell-cycle transitions: Does one fit all? Nat Rev Mol Cell Biol. 9:910–916. 2008. View Article : Google Scholar : PubMed/NCBI

30 

Takeda DY, Wohlschlegel JA and Dutta A: A bipartite substrate recognition motif for cyclin-dependent kinases. J Biol Chem. 276:1993–1997. 2001. View Article : Google Scholar : PubMed/NCBI

31 

Cribier A, Descours B, Valadão AL, Laguette N and Benkirane M: Phosphorylation of SAMHD1 by cyclin A2/CDK1 regulates its restriction activity toward HIV-1. Cell Rep. 3:1036–1043. 2013. View Article : Google Scholar : PubMed/NCBI

32 

Kretschmer S, Wolf C, König N, Staroske W, Guck J, Häusler M, Luksch H, Nguyen LA, Kim B, Alexopoulou D, et al: SAMHD1 prevents autoimmunity by maintaining genome stability. Ann Rheum Dis. 74:e172015. View Article : Google Scholar : PubMed/NCBI

33 

Yan J, Hao C, DeLucia M, Swanson S, Florens L, Washburn MP, Ahn J and Skowronski J: CyclinA2-cyclin-dependent kinase regulates SAMHD1 protein phosphohydrolase domain. J Biol Chem. 290:13279–13292. 2015. View Article : Google Scholar : PubMed/NCBI

34 

Hu J, Qiao M, Chen Y, Tang H, Zhang W, Tang D, Pi S, Dai J, Tang N, Huang A and Hu Y: Cyclin E2-CDK2 mediates SAMHD1 phosphorylation to abrogate its restriction of HBV replication in hepatoma cells. FEBS Lett. 592:1893–1904. 2018. View Article : Google Scholar : PubMed/NCBI

35 

Pauls E, Badia R, Torres-Torronteras J, Ruiz A, Permanyer M, Riveira-Muñoz E, Clotet B, Marti R, Ballana E and Esté JA: Palbociclib, a selective inhibitor of cyclin-dependent kinase4/6, blocks HIV-1 reverse transcription through the control of sterile alpha motif and HD domain-containing protein-1 (SAMHD1) activity. AIDS. 28:2213–2222. 2014. View Article : Google Scholar : PubMed/NCBI

36 

Badia R, Angulo G, Riveira-Muñoz E, Pujantell M, Puig T, Ramirez C, Torres-Torronteras J, Martí R, Pauls E, Clotet B, et al: Inhibition of herpes simplex virus type 1 by the CDK6 inhibitor PD-0332991 (palbociclib) through the control of SAMHD1. J Antimicrob Chemother. 71:387–394. 2016. View Article : Google Scholar : PubMed/NCBI

37 

Allouch A, David A, Amie SM, Lahouassa H, Chartier L, Margottin-Goguet F, Barré-Sinoussi F, Kim B, Sáez-Cirión A and Pancino G: p21-mediated RNR2 repression restricts HIV-1 replication in macrophages by inhibiting dNTP biosynthesis pathway. Proc Natl Acad Sci USA. 110:E3997–E4006. 2013. View Article : Google Scholar

38 

Pauls E, Ruiz A, Riveira-Muñoz E, Permanyer M, Badia R, Clotet B, Keppler OT, Ballana E and Este JA: p21 regulates the HIV-1 restriction factor SAMHD1. Proc Natl Acad Sci USA. 111:E1322–E1324. 2014. View Article : Google Scholar

39 

Osei Kuffour E, Schott K, Jaguva Vasudevan AA, Holler J, Schulz WA, Lang PA, Lang KS, Kim B, Häussinger D, König R and Münk C: USP18 (UBP43) abrogates p21-mediated inhibition of HIV-1. J Virol. 92:e00592–18. 2018. View Article : Google Scholar

40 

Schott K, Fuchs NV, Derua R, Mahboubi B, Schnellbächer E, Seifried J, Tondera C, Schmitz H, Shepard C, Brandariz-Nuñez A, et al: Dephosphorylation of the HIV-1 restriction factor SAMHD1 is mediated by PP2A-B55alpha holoenzymes during mitotic exit. Nat Commun. 9:22272018. View Article : Google Scholar : PubMed/NCBI

41 

Kueck T, Cassella E, Holler J, Kim B and Bieniasz PD: The aryl hydrocarbon receptor and interferon gamma generate antiviral states via transcriptional repression. Elife. 7:e388672018. View Article : Google Scholar : PubMed/NCBI

42 

Bonifati S, Daly MB, St Gelais C, Kim SH, Hollenbaugh JA, Shepard C, Kennedy EM, Kim DH, Schinazi RF, Kim B and Wu L: SAMHD1 controls cell cycle status, apoptosis and HIV-1 infection in monocytic THP-1 cells. Virology. 495:92–100. 2016. View Article : Google Scholar : PubMed/NCBI

43 

Kodigepalli KM, Li M, Liu SL and Wu L: Exogenous expression of SAMHD1 inhibits proliferation and induces apoptosis in cutaneous T-cell lymphoma-derived HuT78 cells. Cell Cycle. 16:179–188. 2017. View Article : Google Scholar : PubMed/NCBI

44 

Welbourn S, Miyagi E, White TE, Diaz-Griffero F and Strebel K: Identification and characterization of naturally occurring splice variants of SAMHD1. Retrovirology. 9:862012. View Article : Google Scholar : PubMed/NCBI

45 

Kodigepalli KM, Bonifati S, Tirumuru N and Wu L: SAMHD1 modulates in vitro proliferation of acute myeloid leukemia-derived THP-1 cells through the PI3K-Akt-p27 axis. Cell Cycle. 17:1124–1137. 2018. View Article : Google Scholar : PubMed/NCBI

46 

Feng Q, Moran JV, Kazazian HH Jr and Boeke JD: Human L1 retrotransposon encodes a conserved endonuclease required for retrotransposition. Cell. 87:905–916. 1996. View Article : Google Scholar : PubMed/NCBI

47 

Burns KH: Transposable elements in cancer. Nat Rev Cancer. 17:415–424. 2017. View Article : Google Scholar : PubMed/NCBI

48 

White TE, Brandariz-Nuñez A, Valle-Casuso JC, Knowlton C, Kim B, Sawyer SL and Diaz-Griffero F: Effects of human SAMHD1 polymorphisms on HIV-1 susceptibility. Virology. 460-461:34–44. 2014. View Article : Google Scholar : PubMed/NCBI

49 

White TE, Brandariz-Nuñez A, Valle-Casuso JC, Amie S, Nguyen L, Kim B, Brojatsch J and Diaz-Griffero F: Contribution of SAM and HD domains to retroviral restriction mediated by human SAMHD1. Virology. 436:81–90. 2013. View Article : Google Scholar : PubMed/NCBI

50 

Du J, Peng Y, Wang S, Hou J, Wang Y, Sun T and Zhao K: Nucleocytoplasmic shuttling of SAMHD1 is important for LINE-1 suppression. Biochem Biophys Res Commun. 510:551–557. 2019. View Article : Google Scholar : PubMed/NCBI

51 

Herrmann A, Wittmann S, Thomas D, Shepard CN, Kim B, Ferreirós N and Gramberg T: The SAMHD1-mediated block of LINE-1 retroelements is regulated by phosphorylation. Mobile DNA. 9:112018. View Article : Google Scholar : PubMed/NCBI

52 

Hu S, Li J, Xu F, Mei S, Le Duff Y, Yin L, Pang X, Cen S, Jin Q, Liang C and Guo F: SAMHD1 inhibits LINE-1 retrotransposition by promoting stress granule formation. PLoS Genetics. 11:e10053672015. View Article : Google Scholar : PubMed/NCBI

53 

Ha°kansson P, Hofer A and Thelander L: Regulation of mammalian ribonucleotide reduction and dNTP pools after DNA damage and in resting cells. J Biol Chem. 281:7834–7841. 2006. View Article : Google Scholar

54 

Jauregui P and Landau NR: DNA damage induces a SAMHD1-mediated block to the infection of macrophages by HIV-1. Sci Rep. 8:41532018. View Article : Google Scholar : PubMed/NCBI

55 

Daddacha W, Koyen AE, Bastien AJ, Head PE, Dhere VR, Nabeta GN, Connolly EC, Werner E, Madden MZ, Daly MB, et al: SAMHD1 promotes DNA end resection to facilitate DNA repair by homologous recombination. Cell Rep. 20:1921–1935. 2017. View Article : Google Scholar : PubMed/NCBI

56 

Ballana E, Badia R, Terradas G, Torres-Torronteras J, Ruiz A, Pauls E, Riveira-Muñoz E, Clotet B, Martí R and Esté JA: SAMHD1 specifically affects the antiviral potency of thymidine analog HIV reverse transcriptase inhibitors. Antimicrob Agents Chemother. 58:4804–4813. 2014. View Article : Google Scholar : PubMed/NCBI

57 

Guièze R, Robbe P, Clifford R, de Guibert S, Pereira B, Timbs A, Dilhuydy MS, Cabes M, Ysebaert L, Burns A, et al: Presence of multiple recurrent mutations confers poor trial outcome of relapsed/refractory CLL. Blood. 126:2110–2117. 2015. View Article : Google Scholar

58 

Grant S: Ara-C: Cellular and molecular pharmacology. Adv Cancer Res. 72:197–233. 1998. View Article : Google Scholar : PubMed/NCBI

59 

Schneider C, Oellerich T, Baldauf HM, Schwarz SM, Thomas D, Flick R, Bohnenberger H, Kaderali L, Stegmann L, Cremer A, et al: SAMHD1 is a biomarker for cytarabine response and a therapeutic target in acute myeloid leukemia. Nat Med. 23:250–255. 2017. View Article : Google Scholar : PubMed/NCBI

60 

Arnold LH, Kunzelmann S, Webb MR and Taylor IA: A continuous enzyme-coupled assay for triphosphohydrolase activity of HIV-1 restriction factor SAMHD1. Antimicrob Agents Chemother. 59:186–192. 2015. View Article : Google Scholar : PubMed/NCBI

61 

Knecht KM, Buzovetsky O, Schneider C, Thomas D, Srikanth V, Kaderali L, Tofoleanu F, Reiss K, Ferreirós N, Geisslinger G, et al: The structural basis for cancer drug interactions with the catalytic and allosteric sites of SAMHD1. Proc Natl Acad Sci USA. 115:E10022–E10031. 2018. View Article : Google Scholar

62 

Rassidakis GZ, Herold N, Myrberg IH, Tsesmetzis N, Rudd SG, Henter JI, Schaller T, Ng SB, Chng WJ, Yan B, et al: Low-level expression of SAMHD1 in acute myeloid leukemia (AML) blasts correlates with improved outcome upon consolidation chemotherapy with high-dose cytarabine-based regimens. Blood Cancer J. 8:982018. View Article : Google Scholar : PubMed/NCBI

63 

Wang YT, Yuan B, Chen HD, Xu L, Tian YN, Zhang A, He JX and Miao ZH: Acquired resistance of phosphatase and tensin homolog-deficient cells to poly(ADP-ribose) polymerase inhibitor and Ara-C mediated by 53BP1 loss and SAMHD1 overexpression. Cancer Sci. 109:821–831. 2017. View Article : Google Scholar

64 

Rudd SG, Tsesmetzis N, Sanjiv K, Paulin CB, Sandhow L, Kutzner J, Hed Myrberg I, Bunten SS, Axelsson H, Zhang SM, et al: Ribonucleotide reductase inhibitors suppress SAMHD1 ara-CTPase activity enhancing cytarabine efficacy. EMBO Mol Med. 12:e104192020. View Article : Google Scholar : PubMed/NCBI

65 

Hollenbaugh JA, Shelton J, Tao S, Amiralaei S, Liu P, Lu X, Goetze RW, Zhou L, Nettles JH, Schinazi RF and Kim B: Substrates and Inhibitors of SAMHD1. PLoS One. 12:e01690522017. View Article : Google Scholar : PubMed/NCBI

66 

Jones S, Zhang X, Parsons DW, Lin JC, Leary RJ, Angenendt P, Mankoo P, Carter H, Kamiyama H, Jimeno A, et al: Core signaling pathways in human pancreatic cancers revealed by global genomic analyses. Science. 321:1801–1806. 2008. View Article : Google Scholar : PubMed/NCBI

67 

Kohnken R, Kodigepalli KM and Wu L: Regulation of deoxynucleotide metabolism in cancer: Novel mechanisms and therapeutic implications. Mol Cancer. 14:1762015. View Article : Google Scholar : PubMed/NCBI

68 

James CD, Prabhakar AT, Otoa R, Evans MR, Wang X, Bristol ML, Zhang K, Li R and Morgan IM: SAMHD1 regulates human papillomavirus 16-induced cell proliferation and viral replication during differentiation of keratinocytes. mSphere. 4:e00448–19. 2019. View Article : Google Scholar

69 

Stanland LJ and Luftig MA: The role of EBV-induced hypermethylation in gastric cancer tumorigenesis. Viruses. 12:12222020. View Article : Google Scholar : PubMed/NCBI

70 

Cruz-Gregorio A, Aranda-Rivera AK and Pedraza-Chaverri J: Human papillomavirus-related cancers and mitochondria. Virus Res. 286:1980162020. View Article : Google Scholar : PubMed/NCBI

71 

Cabello-Lobato MJ, Wang S and Schmidt CK: SAMHD1 sheds moonlight on DNA double-strand break repair. Trends Genet. 33:895–897. 2017. View Article : Google Scholar : PubMed/NCBI

72 

Seamon KJ and Stivers JT: A high-throughput enzyme-coupled assay for SAMHD1 dNTPase. J Biomol Screen. 20:801–809. 2015. View Article : Google Scholar : PubMed/NCBI

73 

Mauney CH, Perrino FW and Hollis T: Identification of inhibitors of the dNTP triphosphohydrolase SAMHD1 using a novel and direct high-throughput assay. Biochemistry. 57:6624–6636. 2018. View Article : Google Scholar : PubMed/NCBI

74 

Jiang H, Li C, Liu Z and Shengjing Hospital; Hu: Expression and relationship of SAMHD1 with other apoptotic and autophagic genes in acute myeloid leukemia patients. Acta Haematol. 143:51–59. 2020. View Article : Google Scholar : PubMed/NCBI

75 

Kodigepalli KM, Li M, Bonifati S, Panfil AR, Green PL, Liu SL and Wu L: SAMHD1 inhibits epithelial cell transformation in vitro and affects leukemia development in xenograft mice. Cell Cycle. 17:2564–2576. 2018. View Article : Google Scholar : PubMed/NCBI

76 

Herold N, Rudd SG, Sanjiv K, Kutzner J, Myrberg IH, Paulin CBJ, Olsen TK, Helleday T, Henter JI and Schaller T: With me or against me: Tumor suppressor and drug resistance activities of SAMHD1. Exp Hematol. 52:32–39. 2017. View Article : Google Scholar : PubMed/NCBI

77 

Rothenburger T, McLaughlin KM, Herold T, Schneider C, Oellerich T, Rothweiler F, Feber A, Fenton TR, Wass MN, Keppler OT, et al: SAMHD1 is a key regulator of the lineage-specific response of acute lymphoblastic leukaemias to nelarabine. Commun Biol. 3:3242020. View Article : Google Scholar : PubMed/NCBI

78 

Yang CA, Huang HY, Chang YS, Lin CL, Lai IL and Chang JG: DNA-sensing and nuclease gene expressions as markers for colorectal cancer progression. Oncology. 92:115–124. 2017. View Article : Google Scholar : PubMed/NCBI

79 

Shang Z, Qian L, Liu S, Niu X, Qiao Z, Sun Y, Zhang Y, Fan LY, Guan X, Cao CX and Xiao H: Graphene oxide-facilitated comprehensive analysis of cellular nucleic acid binding proteins for lung cancer. ACS Appl Mater Interfaces. 10:17756–17770. 2018. View Article : Google Scholar : PubMed/NCBI

80 

Shi Y, Lv G, Chu Z, Piao L, Liu X, Wang T, Jiang Y and Zhang P: Identification of natural splice variants of SAMHD1 in virus-infected HCC. Oncol Rep. 31:687–692. 2014. View Article : Google Scholar : PubMed/NCBI

81 

Merati M, Buethe DJ, Cooper KD, Honda KS, Wang H and Gerstenblith MR: Aggressive CD8(+) epidermotropic cutaneous T-cell lymphoma associated with homozygous mutation in SAMHD1. JAAD Case Rep. 1:227–229. 2015. View Article : Google Scholar : PubMed/NCBI

82 

Amin NA, Seymour E, Saiya-Cork K, Parkin B, Shedden K and Malek SN: A quantitative analysis of subclonal and clonal gene mutations before and after therapy in chronic lymphocytic leukemia. Clin Cancer Res. 22:4525–4535. 2016. View Article : Google Scholar : PubMed/NCBI

83 

Zhu KW, Chen P, Zhang DY, Yan H, Liu H, Cen LN, Liu YL, Cao S, Zhou G, Zeng H, et al: Association of genetic polymorphisms in genes involved in Ara-C and dNTP metabolism pathway with chemosensitivity and prognosis of adult acute myeloid leukemia (AML). J Transl Med. 16:902018. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

June-2021
Volume 21 Issue 6

Print ISSN: 1792-1074
Online ISSN:1792-1082

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Chen Z, Hu J, Ying S and Xu A: Dual roles of SAMHD1 in tumor development and chemoresistance to anticancer drugs (Review). Oncol Lett 21: 451, 2021
APA
Chen, Z., Hu, J., Ying, S., & Xu, A. (2021). Dual roles of SAMHD1 in tumor development and chemoresistance to anticancer drugs (Review). Oncology Letters, 21, 451. https://doi.org/10.3892/ol.2021.12712
MLA
Chen, Z., Hu, J., Ying, S., Xu, A."Dual roles of SAMHD1 in tumor development and chemoresistance to anticancer drugs (Review)". Oncology Letters 21.6 (2021): 451.
Chicago
Chen, Z., Hu, J., Ying, S., Xu, A."Dual roles of SAMHD1 in tumor development and chemoresistance to anticancer drugs (Review)". Oncology Letters 21, no. 6 (2021): 451. https://doi.org/10.3892/ol.2021.12712