Skip to main content

Protein arginine methylation: an emerging regulator of the cell cycle

Abstract

Protein arginine methylation is a common post-translational modification where a methyl group is added onto arginine residues of a protein to alter detection by its binding partners or regulate its activity. It is known to be involved in many biological processes, such as regulation of signal transduction, transcription, facilitation of protein–protein interactions, RNA splicing and transport. The enzymes responsible for arginine methylation, protein arginine methyltransferases (PRMTs), have been shown to methylate or associate with important regulatory proteins of the cell cycle and DNA damage repair pathways, such as cyclin D1, p53, p21 and the retinoblastoma protein. Overexpression of PRMTs resulting in aberrant methylation patterns in cancers often correlates with poor recovery prognosis. This indicates that protein arginine methylation is also an important regulator of the cell cycle, and consequently a target for cancer regulation. The effect of protein arginine methylation on the cell cycle and how this emerging key player of cell cycle regulation may be used in therapeutic strategies for cancer are the focus of this review.

Background

The cell cycle allows cells to divide and is characterised by the replication of DNA and the subsequent division of duplicated chromosomes into two daughter cells [1]. Regulatory proteins including the tumor suppressor protein p53 and the retinoblastoma protein (pRb) monitor the cell cycle and initiate pathways in response to DNA damage [2]. Cancer is considered to be a disease of the cell cycle [3] and as such these cell cycle regulatory proteins are often found to be deregulated in cancer. Protein arginine methylation is a post-translational modification often upregulated in cancer and other diseases [4], contributing to the deregulation of the cell cycle.

This review will discuss the current understanding of the role protein arginine methylation plays in cell cycle regulation and the implications for potential cancer treatment.

Cell division

The controlled regulation of cell division is essential in growth, repair and re-generation of healthy tissues and occurs via the cell cycle. The cell cycle is divided into non-overlapping stages or phases which are referred to as gap 1 (G1), synthesis (S), gap 2 (G2) and mitosis (M) phase [5] (Fig. 1). In the G1 phase, the cell is preparing for DNA replication which then occurs in the S phase, while cells in the G2 phase are preparing for the M phase, where cell division takes place [5]. M phase can be further divided into prophase, metaphase, anaphase, telophase and cytokinesis [6]. For a visual review of mitotic regulation see [7]. Cells that are not actively dividing are said to be in G0, a resting or quiescent phase. Once cells have committed to DNA replication they cannot return to G0 [1]. In order to guarantee successful cell division without errors, the cell cycle is stringently regulated. Deregulation of the cell cycle, resulting in uncontrolled cell proliferation is one of the hallmarks of cancer [8].

Fig. 1
figure 1

Overview of the regulation of the cell cycle. The cell cycle is depicted as a circle where each black arrow represents one phase of the cell cycle. Cells enter the cell cycle into the G1 phase, which is followed by the S phase, G2 phase and then mitosis (M). The cell cycle is regulated by CDKs (purple), their regulatory subunit cyclins (pink), CDK inhibitors such as p21 (light blue), and other regulatory kinases, such as the checkpoint kinases 1 and 2 (Chk1 and Chk2; yellow). The orange lines at the end of G1, G2 and during mitosis indicate cell cycle checkpoints where the cell is monitored for defects during replication and can respond by the p53 (green) pathway or the ATM/R (red) pathway, among others. Yellow circles with “M” indicate proteins known to be methylated on arginine residues and the dark blue circles with “P” indicate phosphorylation of pRB

Cell cycle regulation

Regulation of the cell cycle is mainly achieved by a family of serine/threonine kinases called cyclin-dependent kinases (CDKs). However, full CDK activation only occurs upon the association of a CDK and a cyclin subunit [9]. CDKs allow cell cycle phase transitions to occur [10]. Nine CDKs have been identified with five of them found to be active as complexes with various cyclin subunits during the cell cycle [11, 12]. CDK levels remain constant throughout the cell cycle, while the levels of cyclin proteins rise and fall during the cell cycle to activate CDKs when they are required [13]. Cyclin D1 is overexpressed in 50% of breast cancers [14], and at lower frequencies in many other cancers including prostate cancer [15]; while Cyclin E has been reported to be deregulated in malignant tumors of the lung, breast, gastrointestinal tract and in ovarian cancer [16].

Although CDK levels remain constant throughout the cell cycle, their activity is regulated by phosphorylation on specific residues to induce conformational changes and enhance cyclin binding. For example, full activation of CDK2 requires phosphorylation on threonine 160 by the CDK7-cyclin H complex also called the CDK activating kinase (CAK) [17]. Inhibitory kinases, such as Wee1 and Myt1 can function as inactivators of CDK1 by phosphorylating it. Dephosphorylation is then required for CDK1 reactivation [18]. Cyclin dependent kinase inhibitors (CDKIs) regulate CDK activity by binding to CDKs or the CDK-cyclin complex [1]. There are two families of CDKIs which have been discovered so far: Inhibitors of CDK4 (INK4) proteins and CDK interacting proteins/kinase inhibitory proteins (CIP/KIP) [1] (Fig. 1). The INK4 inhibitors (p15INK4b, p16INK4a, p18INK4c, p19INK4d) specifically inactivate the cyclin D-dependent CDKs, i.e. CDK4 and CDK6, by competitively binding these CDKs and preventing them from forming a complex with cyclin D [19]. The CIP/KIP inhibitors (p21CIP1, p27KIP1, p57KIP2) bind to and inactivate multiple G1 CDK-cyclin complexes [20]. p21 also binds to the proliferating cell nuclear antigen (PCNA), inhibiting DNA replication [21]. CDKIs are themselves regulated by internal and external signals. For example, the p21 gene promoter encodes a p53 binding site allowing p53 to transcriptionally activate the p21 gene [22]. The expression of p15 and activation of p27 both increase in response to transforming growth factor β [20]. CDK inhibitors, such as p15 and p18, are frequently deregulated in cancer [3], allowing uncontrolled progression into S phase [23].

The Phosphatidylinositol 3-kinase (PI3K) pathway regulates cellular processes through its association with the protein kinase AKT and is involved in the G1/S phase transition of the cell cycle [24]. AKT prevents cyclin D1 degradation by directly phosphorylating glycogen synthase kinase 3 β (GSK3β), blocking its kinase activity and allowing cyclin D1 to accumulate [25]. (PI3K)/AKT signaling is frequently deregulated in human cancers, including ovarian, breast, lung, thyroid and melanomas [26]. The tumor suppressor Phosphatase and tensin homolog (PTEN) gene is a negative regulator of the PI3K/AKT signaling pathway [27] and is one of the most mutated or deleted genes across different cancer types [28], leading to deregulation of the PI3K/AKT pathway.

Cell cycle checkpoints

In addition to the cell cycle regulation by CDKs and CDKIs, there are several checkpoints, such as the DNA repair checkpoints that regulate G1/S and G2/M phase transitions (see Fig. 1), where cell size, extracellular growth signals [3], and progression through the cell cycle are monitored to prevent defects or to repair any DNA damage which may have occurred during DNA synthesis [29]. It is these DNA defects, which if left unrepaired, are passed onto daughter cells during cell division and contribute to the deregulation of the cell cycle and may lead to the unrestrained cell proliferation characteristics of cancer [1].

DNA damage checkpoints occur before cells enter S phase (G1-S checkpoint) where cell cycle arrest induced by DNA damage is p53 dependent; or after DNA replication, with or without the tumour suppressor protein, p53 (G2-M checkpoint) [30]. The checkpoint kinase proteins, Chk1 and Chk2, phosphorylate the cell division cycle phosphatases, Cdc25, to regulate the phosphorylation of CDK1 and CDK2, and therefore regulate G1 into S phase and S phase into G2/M transitions, respectively. DNA damage repair is also monitored during S phase to block replication, if damaged DNA has escaped the G1/S phase checkpoint without undergoing repair [31].

The spindle assembly checkpoint (SAC) detects the improper alignment of chromosomes on the mitotic spindle and can stop the cell cycle in metaphase, if required, to prevent the formation of cells with the incorrect number of chromosomes [32]. This checkpoint is controlled by the mitotic-arrest deficient (mad) and the budding uninhibited by benzimidazole (bub1) families of genes [33, 34]. The mitotic checkpoint complex (MCC) consists of MAD2, BUBR1 and BUB3 proteins and negatively regulates the activity of cell division cycle protein 20 (CDC20) [35,36,37,38]. CDC20 binds to the ubiquitin ligase anaphase-promoting complex/cyclosome (APC/C) [38], preventing the APC/C from exerting its ubiquitinylation activity on securin and cyclin B [39]. Securin is an inhibitor of the protease separase which is required to cleave the cohesion complex which holds sister chromatids together [40], while cyclin B is required to activate CDK1 to promote exit of the cells from mitosis [41]. Thus, inhibition of the APC/C by the MCC prevents cells from entering anaphase. Other important SAC proteins include MAD1, BUB1, MAPK and Aurora B, which promote recruitment of the SAC proteins to the kinetochores [42]. Aurora B is also involved in correcting aberrant merotelic attachment of kinetochores to microtubules from opposite poles to prevent the division of cells with the improper number of chromosomes [43]. Ubiquitin-associated protein 2-like (UBAP2L) is also necessary for proper kinetochore/microtubule attachment and depletion of UBAP2L activates SAC signalling, delaying progression into anaphase [44]. For a comprehensive review on the SAC see [45].

DNA damage repair

DNA damage can be induced via a variety of triggers (including UV radiation, chemicals, and stress) and if left unrepaired can lead to cancer. Therefore, cells respond to unrepairable DNA damage by stopping cell cycle progression or by initiating programmed cell death via several pathways.

Depending on the type of DNA damage, five major pathways of DNA repair are utilised by cells including: direct repair, base excision repair, nucleotide excision repair, mismatch and recombinational repair, and double strand break repair [46]. These repair pathways will be briefly explained here but the reader is referred to other reviews [47,48,49,50,51] for a more comprehensive coverage.

In mammalian cells, alkylation damage is repaired by direct repair alkyltransferases. O6-methylguanine-DNA methyltransferase is the main enzyme facilitating direct repair by transferring a methyl group from the DNA backbone into the active site of the enzyme [52].

Three excision repair mechanisms remove specific types of damage: Firstly, base excision repair (BER) specifically recognizes base damage. It most commonly deals with base damage that occurs during normal aerobic metabolism via reactive oxygen species (most common base damage: dihydro-8-oxoguanine). BER requires removal of the damaged DNA base by a specific DNA glycosylase to initiate the process [51]. Short nucleotide gaps are then re-synthesized via APE1, DNA polymerase β [53] and the Ligase 3/XRCC1 complex; while longer breaks of 2–10 nucleotides are repaired by the RFC/PCNA-DNA polymerase δ/ε complex [54]. Secondly, nucleotide excision repair is the major pathway for the removal of bulky DNA lesions formed by exposure to environmental sources, ultraviolet radiation or chemicals [55]. In humans, the process of nucleotide excision repair involves damage recognition, incision around the lesion to release a 24–32 nucleotide oligomer and repair of the resulting gap by DNA polymerase δ/ε and ligation. For a review on nucleotide excision repair see [56]. In humans, nucleotide excision repair requires more than 30 proteins to successfully execute the repair [55]. Repair proteins including XPA, RPA, XPC, TFIIH, XPG and XPF-ERCC1 perform these functions [57,58,59]. Thirdly, mismatch repair functions to repair errors such as mismatches, insertions or deletions that can occur during replication [50].

Double strand breaks (DSBs) are the most lethal type of damage to the cell. DSB damage occurs when cells are exposed to ionizing radiation and during normal cellular recombination and immunoglobulin class-switching processes [60]. DSBs can be repaired via homologous recombination or non-homologous end-joining (NHEJ) [61]. Homologous recombination occurs at a homologous stretch of DNA on a sister chromatid serving as a template to guide repair of the broken strand, while NHEJ requires enzymes to capture the ends of the broken DNA, bringing them together in a DNA–protein complex, then the DNA break is repaired via ligation [62]. Both require the proteins NBS1 [63], MRE11 [64], and Rad50 [65]. Homologous recombination also requires the heterodimer MUS81–MMS4 [46], BRCA1 and BRCA2 [66], while NHEJ requires DNA-PKCs and the heterodimer Ligase 4-XRCC4 [67].

DNA strand crosslinks can be induced by many chemotherapeutic drugs and can stall DNA replication, leading to cell death [68]. Depending on the type of cross-link, they are processed either by nucleotide excision repair (intra-strand cross-link) or converted into a double-stranded DNA break and processed by homologous recombination (inter-strand crosslink) [69]. Hence, the same repair proteins are utilised, such as XPF-ERCC1 to degrade one of the cross-linked strands [70]; or the recombination proteins Rad51 and Rad52 to promote homologous DNA pairing and strand exchange [65].

DNA damage repair can also be initiated by tumor suppressor proteins which can induce cell cycle arrest to either initiate repair or mark the cell for apoptosis. Two such proteins that regulate the major pathways of apoptosis and cellular progression are p53 and pRb [10]. Between them, these two proteins can initiate differentiation, DNA repair, cell cycle arrest or progression and induction/inhibition of apoptosis. The p53 and pRb pathways are summarised in the following sections.

p53

p53 is normally expressed at low levels but rises within cells in response to various stimuli, such as DNA damage, hypoxia and oncogene activation, to activate various pathways that initiate differentiation, DNA repair, cell cycle arrest, inhibition of angiogenesis and apoptosis [71]. There are three main responses that p53 can initiate: (1) facilitation of cell cycle arrest by inducing the expression of p21, 14-3-3 proteins, Cdc25C and GADD45 [72]; (2) stimulation of DNA repair by inducing the expression of p21, GADD45 and p48 [10]; and (3) induction of apoptosis by upregulating the transcription of BAX and other apoptotic proteins [73]. Low levels of stress or DNA damage induce p53-mediated cell cycle arrest, while high levels of stress activate p53-mediated apoptotic pathways [74]. Further, different protein kinases such as ataxia-telangiectasia mutated (ATM) and ataxia-telangiectasia and Rad3 related (ATR) phosphorylate p53 in response to DNA damage, leading to p21 arresting the cell cycle at the G1-S checkpoint [47, 75] (see Fig. 1). Inactivation of the p53 gene results in the dysfunction of proteins that would normally inhibit cell cycle proliferation [1]. The inactivation of the p53 gene through mutations is the most commonly occurring loss of a pro-apoptotic regulator and is seen in more than 50% of cancers [76].

Retinoblastoma protein

pRb interacts with proteins involved in transcriptional control such as the E2F proteins that are regulators of gene expression and genes involved in DNA replication, DNA repair and G2/M progression [77] (Fig. 1). pRb regulates cell cycle arrest by binding to E2F. E2F and its heterodimer, the transcription factor DP, are mediators of the p16/pRb pathway of G1 cell cycle arrest [78]. However, only unphosphorylated forms of pRb can interact with E2F proteins [79] (Fig. 1). Thus, phosphorylation of pRb carried out by the CDK4-cyclin D1 complex regulates pRb function [80]. pRb is found to be unphosphorylated during G0 and G1 phase, and phosphorylated during the remainder of the cell cycle [81].

Phosphorylation clearly plays a pivotal role in cell cycle regulation. Protein methylation is a post-translational modification that may be just as important in cell cycle regulation as phosphorylation.

Protein arginine methylation

During protein methylation, a methyl group is added to a specific protein residue to alter detection by its binding partners or to regulate its activity [82]. This can occur on lysine (K), histidine (H) or arginine (R) residues of both histones and non-histone proteins [83]. Protein arginine methyltransferases were discovered almost 50 years ago and while histone and non-histone protein lysine methylation have been extensively studied over the past 60 years, protein arginine methylation has only gained more attention in the past 20–25 years [84].

Protein arginine methylation has many documented regulatory roles including in signal transduction [85], transcription [86, 87], protein–protein interactions such as facilitating the interactions of Tudor domains with glycine arginine rich (GAR) and proline glycine methionine (PGM) motifs in proteins [88], RNA transport [89], and RNA splicing [90, 91]. For a recent review that covers arginine methylation in different organ systems, see [92].

Protein arginine methyltransferases

In 1968, Paik and Kim attempted to purify the enzyme responsible for the methylation of histone lysine residues [93]. The enzyme they discovered, protein methylase I, was not a lysine methyltransferase but actually a protein arginine methyltransferase or PRMT [93], now known as PRMT1. There are currently eleven PRMTs which are divided into four enzyme types [94]. For a review see [82, 95, 96]. Types I to III exist in mammalian cells and all catalyse the addition of a single methyl group onto a terminal nitrogen atom of arginine residues forming ω-NG-monomethyl arginine (MMA). Type I enzymes are the most common [4] and catalyse the addition of another methyl group onto the same terminal nitrogen atom, forming ω-NG, NG-dimethylarginine or asymmetric dimethylarginine (ADMA) [97]. Type II enzymes catalyse the addition of another methyl group onto the other terminal nitrogen atom, producing ω-NG, NG-dimethylarginine or symmetric dimethylarginine (SDMA) [98]. Type III enzymes only catalyse the formation of MMA [99]. While Type IV enzymes also catalyse the formation of MMA, the methyl group is added to the internal (σ) guanidino nitrogen atom and this has so far only been described in yeast [100]. The structures of the three methylarginine types present in mammals (MMA, ADMA, and SDMA) are shown in Fig. 2.

Fig. 2
figure 2

Types of mammalian protein arginine methylation. Arginine residues in proteins can be monomethylated by type I, II and III PRMTs to form MMA, while type I and II PRMTs further methylate to produce ADMA and SDMA residues, respectively. Red arrows indicate known methylation locations in mammalian cells; red circles indicate methyl groups

Regulation of the cell cycle by arginine methylation

The interaction of many key regulatory proteins of the cell cycle with several PRMTs has been characterised and/or methylated residues have been identified (for summary see Table 1). It should be noted that these findings, while representative of the cell type and/or tissue types examined, may not be representative of all tissues especially as cancer is notorious for differing mutations between patients. While there had been earlier indications that arginine methylation may be involved in cell proliferation, a study by Kim et al. [101] was one of the first to describe the involvement of protein arginine methylation in the different phases of the cell cycle in HeLa cells.

Table 1 Key cell cycle regulators known to be substrates and interacting partners of PRMTs

PRMT1

As the most abundant PRMT [102], the majority of PRMT1 activity can be described as anti-proliferative and/or tumor-suppressor-inducing. Recent work has revealed that PRMT1 binds to the tumor suppressor, BGT2, to play a role in pre-B cell differentiation [103]. The BGT2-PRMT1 complex destabilizes the binding of CDK4 to cyclin D3 through the methylation of CDK4 at arginines 55, 73, 82 and 163, halting cell cycle progression and leading to pre-B cell differentiation [103]. E2F1 is competitively methylated by PRMT1 and PRMT5 [104]. Methylation by PRMT1 stabilizes E2F1 during DNA damage [105] and assists in E2F1-dependent apoptosis [104], while methylation by PRMT5 promotes cellular proliferation [104, 106]. Interestingly, recent publications have brought to light the interplay between arginine methylation and serine phosphorylation, adding an extra layer of complexity to the regulation of the cell cycle. For example, methylation of p16 at arginine 138 by PRMT1 increases as phosphorylation of p16 at serine 140 decreases, and vice versa in 293T cells [107]. These modifications work antagonistically to regulate the interaction of p16 with CDK4 [107], thus regulating cell proliferation and apoptosis.

PRMT2

It is unclear at this stage whether PRMT2 has pro-proliferative or anti-proliferative functions in cancer. Arginine methylation of pRb by PRMT2 contributes to E2F1 transcriptional regulation for progression into S phase [108] by impairing pRb/E2F1 binding [109]. Recently, PRMT2 levels were found to be decreased in breast cancer cells, while knockdown of PRMT2 correlated with increased expression of both p21 [110] and cyclin D1 [111] indicating a regulatory role for PRMT2 in cell cycle progression. Further studies are needed to clarify the exact functions of PRMT2 in cell cycle progression.

PRMT4

PRMT4 appears to have pro-proliferative functions and may have a pro-oncogenic role in some cancers. Expression of Cyclin E1 rises sharply leading into S phase to allow transition into the S phase [6], and levels of cyclin E1 are frequently deregulated in breast cancer [16]. PRMT4 is a positive regulator of the Cyclin E1 gene, by acting as a transcriptional co-activator of ACTR coinciding with histone methylation of the CCNE1 promoter region [112]. PRMT4 is essential for estrogen induced cell cycle proliferation in breast cancer by the positive regulation of E2F1 RNA and protein expression [113]. In addition, increased expression of PRMT4 may contribute to the development of prostate cancer, as it correlates with the androgen independence required for the progression of aggressive tumours [114]. Interestingly, a decrease in PRMT4-mediated methylation of pRb led to a decrease in its phosphorylation, suggesting arginine methylation can increase pRb-mediated cell proliferation [109].

PRMT5

The most studied type II PRMT, PRMT5, has pro-proliferative and pro-oncogenic roles, as described below, even though it appears to have the opposite effect in Fanconi anemia patients. This highlights that the effect of specific PRMTs and their inhibition may not be the same in all types of cancers and may differ in individual patients.

PRMT5 is essential for cell proliferation [115] and correlates with increased protein expression of the G1 phase regulators CDK4 and CDK6 [116]. Deficiency of PRMT5 triggers cell-cycle arrest in the G1 phase [117]. Overexpression of PRMT5 correlates with increased cell proliferation and knockdown of PRMT5 results in cell cycle arrest leading to apoptosis [118]. CDK4 interacts with PRMT5 in HepG2 cells which regulates phosphorylation of pRb, thus regulating pRb/E2F-mediated transcription [119]. Disruption of the PRMT5/CDK4 interaction revealed PRMT5 knockdown HepG2 cells to be more sensitive to the CDK4 inhibitor fascaplysin, marking the combination of PRMT5 and CDK4 inhibition as a potential cancer therapy [119]. PRMT5 deficiency also led to apoptosis in differentiated glioblastoma cells however, in glioblastoma neurospheres it led to G1 cell cycle arrest through increased protein expression of p27 and a decrease in phosphorylation of pRb [120]. PRMT5 expression correlated with cyclin D1 protein levels, while also inversely correlating with p16 levels [121]. Nuclear PRMT5/p16-negative tumors were associated with poor prognosis in oropharyngeal squamous cell carcinoma when compared to nuclear PRMT5-negative/p16- positive tumors [121]. PRMT5 is found to be overexpressed in many cancer types, including glioblastoma [118, 122], lung [123], mantle cell lymphoma [124], ovarian [125] and prostate cancer [126]. In all these studies, knockdown of PRMT5 slowed or inhibited cell proliferation, indicating the enzyme may be affecting key regulators of cell proliferation.

p53-dependent cell cycle arrest is enhanced by methylation of p53 by PRMT5 [127], while reduced expression of p53 during the DNA damage response is triggered by PRMT5-deficiency [117]. While the methylation of p53 at arginine 213 may be controversial [128], mutation of this residue resulted in a decrease of p21 RNA and protein expression directly affecting the S phase of the cell cycle [129]. This indicates that p21 activation by p53 may be mediated by arginine methylation of R213 [129].

Fanconi anemia is a genetic disorder characterized by a high risk of developing cancer, among other clinical characteristics, due to a cellular hypersensitivity to DNA cross-linking agents and faulty DNA damage repair pathways [130]. Fanconi anemia proteins, FANCA-C, E–G, L and M, form a “core complex” which activates DNA repair pathways. In Fanca/ and Fancc−/− mice, arginine methylation of p53 by PRMT5 was decreased in response to oncogenic stress [131]. Interestingly, forced expression of PRMT5 led to the delayed onset of leukemia in irradiated mice [131], indicating that PRMT5 may play a tumor suppressor role in Fanconi anemia patients.

PRMT6

PRMT6 may have pro-proliferative and pro-oncogenic functions in both colon and lung cancers. PRMT6 inhibits p21 along with p27, allowing cell cycle progression through the cyclin dependent kinases 1 and 2 (CDK1/2) [132,133,134]. Methylation at arginine 156 of the p21 protein by PRMT6 increased the cytoplasmic localisation of p21 and resulted in HCT116 colon cancer cells becoming more resistant to the chemotherapy drug, doxorubicin [135]. Methylation of p16 at arginine 138 by PRMT6 caused reduced binding of p16 to CDK4, leading to increased cellular proliferation in A549 cells [136]. Further studies are required to determine the function of PRMT6 in other types of cancers.

Regulation of the DNA damage repair by arginine methylation

Proteins involved in the DNA damage repair pathways are regulated by protein arginine methylation (see Table 2). Arginine methylation of Heterogeneous Nuclear Ribonucleoprotein U-Like 1 (hnRNPUL1) by PRMT1 regulates the interaction of hnRNPUL1 with Nijmegen Breakage Syndrome 1 (NBS1) [137], a component of the double strand DNA break repair complex with MRN (MRE11/Rad50/NBS1). Arginine methylation of hnRNPUL1 also regulates its recruitment to sites of DNA damage [137]. The DNA damage repair proteins, MRE11 and p53 binding protein (p53BP1), are methylated by PRMT1 which regulates their DNA exonuclease activity [138, 139] and localisation to DNA damage sites [140], respectively. PRMT1-deficient cells have an impaired ability to recruit Rad51 to DNA damage sites, causing chromosome instability and cell cycle arrest [141]. PRMT1 also regulates the activity of DNA polymerase β by methylating arginine 137 within the PCNA binding site [142]. This prevents the binding of DNA polymerase β and PCNA, implicating PRMT1 in regulation of BER specifically.

Table 2 Key DNA damage repair proteins known to be substrates and interacting partners of PRMTs

PRMT5 also regulates DNA damage repair through the methylation of Rad9 [143]. HEK293T cells without methylated Rad9 were more susceptible to DNA damage by hydroxyurea and led to increased S/M and G2/M cell cycle checkpoint activation [143].

The DNA binding affinity of DNA polymerase β and its ability to repair short single-stranded DNA breaks is enhanced by the methylation of DNA polymerase β by PRMT6 [144].

Methylation of DNA repair protein, Flap endonuclease (FEN1) at arginine 192, enhances its localization to DNA repair sites [145]. Interestingly, a disruption in arginine methylation causes reduced binding to PCNA allowing the phosphorylation of FEN1 at serine 187 by the CDK2/cyclin E complex, leading to decreased localization of FEN1 at DNA repair sites and a delay in cell cycle progression [145].

Regulation of indirect mediators of the cell cycle by arginine methylation

In addition to the key regulatory proteins of cell cycle and DNA damage repair pathways discussed in the previous sections, other cellular proteins that can indirectly affect cell cycle progression are regulated by protein arginine methylation (see Table 3).

Table 3 Indirect mediators of the cell cycle known to be substrates and interacting partners of PRMTs

PRMT1 regulates telomere length and stability by the methylation of the Telomere repeat binding factor 2 (TRF2) [146]. Depletion of PRMT1 results in telomere doublets and promotes telomere shortening [146]. Methylation at arginine 887 of the Inner centromere protein (INCENP) by PRMT1 enhances its binding affinity to Aurora B [147]. A decrease in INCENP methylation led to repression of Aurora B activity resulting in abnormal chromosome alignment and segregation [147]. Arginine methylation of Ubiquitin-associated protein 2-like (UBAP2L) by PRMT1 is necessary for progression of mitosis by regulating chromosome alignment and distribution [44].

The estrogen receptor α (ERα) is methylated by PRMT1 at arginine 260, resulting in cytoplasmic localisation of ERα and indirectly prevents the downstream phosphorylation of the protein kinase β/α serine/threonine-protein kinase, PKB/AKT [148]. PRMT2 is another co-activator for the ERα and has been implicated in tumour cell growth and progression [149].

The androgen receptor (AR) found in the prostate plays a role in regulating the G1/S phase transition of the cell cycle in prostate cancer [150]. Inhibition of arginine methylation resulted in reduced expression of the AR and reduced cell proliferation [151]. PRMT2 acts as a co-activator of the AR, allowing translocation of both the AR and PRMT2 from the cytoplasm into the nucleus [152]. The AR also associates with PRMT9/10 in the prostate cancer cell line, LNCaP, and knockdown of PRMT9/10 suppressed both cellular growth and expression of the prostate specific antigen (PSA) [153]. PRMT5 was recently found to function as an activator of the AR and regulator of AR-dependent proliferation in LNCaP cells and expression of PRMT5 also correlated with RNA and protein expression of the AR [154]. Although it is not yet known if the AR is methylated on arginine residues, it is clear that methylation plays an important role in AR-mediated cell cycle progression, through the association with PRMTs.

The transcriptional activity of the acetyltransferases CREB-binding protein (CBP)/p300 is regulated by PRMT4-mediated methylation of p300, which is important for G1/S phase transition [155, 156]. Methylation of p300 induces complex formation of PRMT4 and p300 with BRCA1 and p53 to induce expression of p21, hence inhibiting cell cycle progression [155, 157]. p300 also associates with PRMT5, Strap and JMY for arginine methylation of p53 which affects the promoter specificity of p53 and enhances p53-dependent cell cycle arrest [127].

Pathways involved in cell cycle regulation can also affect alternative splicing. The ATM/ATR pathways can regulate alternative splicing in response to DNA damage to promote pro-apoptotic genes [158]. Aurora kinase A inhibition caused downregulation of the splicing factor, SF2/ASF, to regulate bcl-x splicing to trigger apoptosis [159]. Interestingly, splicing factors and their products are often overexpressed in cancer, including ovarian [160], breast [161, 162], lung and colon [161]. hnRNPs are a prominent splicing factor family known to be methylated on arginine residues [163,164,165]. Serine–arginine rich (SR) splicing factors are another prominent splicing factor family [166] and it should come as no surprise that SR splicing factors are methylated due to the abundance of arginine residues they contain. The SR protein SF2A-p32 associates with PRMT1 and PRMT5 [167], while PRMT2 associates with multiple SR proteins and hnRNPs, including Sam68, to regulate alternative splicing of the mitochondrial protein Bcl-x [168]. Arginine methylation of Sam68 by PRMT1 localizes it to the cytoplasm [169] and reduces its RNA binding ability [170], thus indicating that Sam68 is highly regulated by arginine methylation. The localisation and therefore the function of SF2/ASF is also highly regulated by arginine methylation [171]. PRMT4 methylates the splicing factors CA150, SmB, U1C, and SAP49 and promotes exon skipping to regulate alternative pre-mRNA splicing [86]. Alternative splicing of the p53 regulator, MDM4, activates the p53 pathway in response to PRMT5 depletion [172]. The role of PRMTs in splicing has been reviewed by [173]. The effect of the methylation of splicing factors in cell cycle regulation and cancer needs to be further investigated.

From this review, it should have become evident, that the majority of proteins and pathways of the cell cycle reported to be deregulated in various cancers are also the ones that can be regulated via protein arginine methylation. Further, studies that suggest regulation of the DNA damage response by arginine methylation are accumulating [44, 137, 138, 140, 142, 144]. These studies also implicate arginine methylation in the deregulation of DNA damage repair pathways which, if not rectified by the cell, may lead to genomic instability and eventually to the progression of cancer. Altered expression levels of PRMTs have been found in many types of cancer, reviewed by [94, 174], implicating their importance in cell cycle and proliferation, cell cycle checkpoints and DNA replication [175]. Therefore, protein arginine methylation may represent a novel target in the development of anti-cancer drugs.

Implications for cancer treatment

PRMT1 [176, 177], PRMT2 [178], PRMT4 [179], PRMT5 [180], PRMT7 [181], and PRMT8 [182] have been reported to have isoforms with differing subcellular locations and are expressed in different cancer types. For a review see [183]. Although Baldwin et al. [183] speculate that the cancer-specific isoforms may have different substrates, this is based on the different localisations of the isoforms which would allow access to different substrates [183]. It is currently unknown if these isoforms behave differently in their substrate-specificity or their activity. These cancer-specific PRMT isoforms could also be targets for future anti-cancer drug development. Drugs could potentially target the aberrant isoforms in cancerous cells without affecting the other isoforms required for normal cell function.

The opportunity may further exist to use PRMT inhibitors in combination with classic chemotherapy drugs. The currently widely used chemotherapy drug Taxol (Paclitaxel) binds to tubulin-β, causing stabilisation of microtubules during mitosis and inhibiting cell division [184]. Recently, protein arginine methylation in the Taxol-binding region of tubulin-β was proposed to affect the binding ability of the drug [185]. This could pave the way for the use of methylation inhibitors as a parallel treatment with Taxol. Inhibiting arginine methylation of tubulin-β would allow Taxol full access to the tubulin-β binding site to inhibit cell division in a cancer- targeted therapy. Alternatively, Taxol derivatives that bind with higher affinity to methylated tubulin may be developed.

PRMT inhibitors could also be used in combination with chemotherapy drugs to combat chemoresistance. There are multiple known pathways of chemoresistance (for review see [186]). Activation of the NF-κB pro-inflammatory pathway can activate the production of anti-apoptotic proteins resulting in tumour growth [187]; polymorphisms in ATP-binding cassette (ABC) multidrug efflux pumps can prevent drugs from crossing the blood–brain barrier and even pump drugs back out of target cells [188]; or DNA repair pathways can be upregulated to repair damage caused by DNA-damaging drugs [189, 190]. As discussed previously, PRMT1 methylates or associates with DNA repair proteins and in many cases, this regulates or enhances the binding of DNA repair proteins to the damage sites [137, 138, 140, 142,143,144,145]. An inhibitor targeting PRMT1 could be used to disrupt the binding of DNA repair proteins to the DNA damage sites, thus preventing cancer cells from evading apoptosis. This treatment would need to be delivered in a tumor-targeting vector as PRMT1 knockdown is known to be embryonically lethal [141] and would also be damaging to non-tumorigenic cells.

Other drugs should be looked at with the possibility of combinatorial cancer therapies. As discussed previously, PRMT5 interacts with CDK4 to regulate pRb/E2F-mediated transcription. Knockdown of PRMT5 made HepG2 cells more sensitive to the CDK4 inhibitor fascaplysin [119]. While this is a promising study, fascaplysin has only been used on cell lines [191,192,193,194], and would require further study in order to be used for cancer therapy. PD0332991 (Palbociclib), was the first CDK4/6 inhibitor to be approved for cancer therapy. It is used in the treatment of pRb positive breast cancer [195], mantle cell lymphoma (MCL) [196], and liposarcoma [197]. This CDK4 inhibitor may also be a potential candidate for combination therapy with PRMTs. An earlier study showed that knockdown of PRMT7 sensitised HeLa cells to the DNA topoisomerase I inhibitor Camptothecin, although the authors did not investigate the mechanism of how this occurs [198]. PRMT7 does not have many known substrates and its PRMT type status has been a controversial topic in the past with some studies claiming it to be a type II PRMT [199], while others proposed it to be a type III PRMT [99]. Further research on the exact mechanism of methylation by PRMT7 is required before it can be considered as a target for combinatorial therapy.

As mentioned previously, PRMTs are often overexpressed in cancers, leading to aberrant methylation patterns. Inhibitors of arginine methylation may be useful to treat tumors by correcting increased levels of protein methylation which may deregulate the cell cycle, DNA damage repair and other important cellular functions. Arginine methyltransferase inhibitor-1 (AMI-1), is able to inhibit the coactivator function of PRMTs [200] while Adenosine dialdehyde (AdOx) inhibits S-adenosyl homocysteine hydrolase preventing methylation from occurring by a negative feedback mechanism [89]. Although AdOx does not inhibit PRMTs directly, several studies have shown that treatment of cell lines with high concentrations of AdOx induces a G2/M phase arrest of the cell cycle [201,202,203]. Inhibition of methylation with AdOx also showed a similar decreased growth rate and reduced migration activity when compared to PRMT1 knockdown cells [204]. This may be useful in treating tumors, if the inhibition of methylation by AdOx can be targeted towards tumor cells only. However, AdOx also causes a decrease in DNA, RNA and lysine methylation, thus further studies would be required to determine its suitability for treatment in vivo as well as the actual role of DNA, RNA and lysine methylation in cancer development.

Other non-specific PRMT inhibitors have been identified-MS023 binds to the active site of type I PRMTs [205]; while DS-437 acts on PRMTs 5 and 7 [206]. Inhibitors which are more specific to individual PRMTs would be preferential for combinatorial cancer therapy to minimise potential off-target effects. Promising specific PRMT inhibitors will be discussed below and have recently been reviewed [207].

PRMT1, being the most abundant PRMT [102], has been the target of the majority of PRMT-specific inhibitor development. Cyanine-derivative compounds have been synthesized based on the structure of AMI-1, such as E-84 which preferentially binds to PRMT1 over other PRMTs and has been demonstrated to decrease leukemia cell proliferation [208]. High-throughput screening was utilised to identify compounds to competitively bind to PRMT1 and PRMT8 [209]. Only PRMT1 and PRMT8 contain a hyper-reactive cysteine residue (C101) within the active site that comes into direct contact with S-adenosyl methionine during the methylation reaction [210]. CID5380390 and CID2818500 were found to produce the strongest inhibition with IC50 values of 23 and 11 μM, respectively [209]. CID5380390 was used to characterize PRMT activity in E. grandis roots [211]. However, no further studies have been published on the mechanism of action for either of these inhibitors.

The PRMT5-specific inhibitor, EPZ015666 (GSK3235025), inhibited growth in a panel of five MCL cell lines and inhibited tumor growth in a dose-dependent manner in MCL xenograft models [212]. The inhibitor structure was altered and renamed EPZ015938 and entered phase I clinical dose-escalation trials in 2016 as compound GSK3326595 for the treatment of solid tumors and non-Hodgkin’s lymphoma (NCT02783300) [213].

More specific inhibition of individual methylated residues with peptide or small molecule inhibitors will present fewer side effects and will be the most likely successful design of personalised cancer treatment of the future. In line with this, a recent review discussed first-generation inhibitors of arginine methylation currently in pre-clinical or phase I/II clinical trials [214].

Next steps

While it is clear that protein arginine methylation is emerging as a key regulator of the cell cycle and may offer suitable targets for novel cancer drug development in the future, the immediate research effort should focus on a more detailed and complete cataloguing of PRMTs and their substrates at different stages of the cell cycle and in various cancer types. Hand in hand with this rather mammoth effort would need to be the development of more specific PRMT inhibitors to not only identify which substrate is methylated on which residue by what PRMT, but to also facilitate the study of downstream functions of the methylation and molecular mechanisms of arginine methylation including effects on alternative splicing. It should be noted that these arginine methylation modifications or aberrant PRMT expression levels could differ between cell/tissue type, between cancer types and even between individual patients. Further emphasis should be placed on unravelling the complex interplay and cross regulation of phosphorylation and methylation which has been so far reported in a few proteins but may be a more widespread regulatory mechanism of protein function that may offer further targets of intervention.

Conclusions

In the coming age of personalised cancer treatment, targeting the specific mutations and anomalous proteins of each patient’s cancer will lead to increased recovery rates. Although our current knowledge of the role of arginine methylation in cell cycle control and cancer development is still in its infancy, it is clear that arginine methylation is an emerging key regulator of the cell cycle that rivals protein phosphorylation in its importance. Further studies are required to determine the exact role that protein arginine methylation plays within the cell cycle, and how this may be used to develop future cancer treatments to target aberrant protein arginine methylation.

Abbreviations

ABC:

ATP-binding cassette

ADMA:

asymmetric dimethylarginine

AdOx:

adenosine dialdehyde oxidised

AKT/PKB:

α serine/threonine-protein kinase/protein kinase B

AMI1:

arginine methyltransferase inhibitor 1

APC/C:

anaphase-promoting complex/cyclosome

APE1:

apurinic/apyrimidinic endonuclease

AR:

androgen receptor

ATM:

ataxia-telangiectasia mutated

ATR:

ataxia-telangiectasia and rad3 related

BER:

base excision repair

BRCA:

tumor suppressor gene

BUB:

budding uninhibited by benzimidazole

C:

cysteine

CAK:

CDK activating kinase

CBP:

CREB-binding protein

CCNE1:

cyclin E1 gene

CDC:

cell division cycle phosphatase

CDK:

cyclin-dependent kinase

CDKI:

cyclin-dependent kinase inhibitor

Chk:

checkpoint kinase

CIP:

CDK interacting proteins

DNA:

deoxyribonucleic acid

DP:

transcription factor family

DSB:

double strand break

E2F:

transcription factor family

ER:

estrogen receptor

ERCC1:

DNA repair endonuclease

FEN1:

flap endonuclease 1

G0 :

resting phase of cell cycle

G1 phase:

first gap phase of cell cycle

G2 phase:

second gap phase of cell cycle

GADD45:

growth arrest and DNA damage-inducible 45α

GAR:

glycine–arginine rich

GSK3β:

glycogen synthase kinase 3 β

H:

histidine

hnRNPUL1:

heterogenous nuclear ribonucleo protein

INCENP:

inner centromere protein

INK4:

inhibitors of CDK4

JMY:

junction mediating and regulatory protein

K:

lysine

KIP:

kinase inhibitory proteins

M phase:

mitosis phase of cell cycle

MAD:

mitotic-arrest deficient

MCC:

mitotic checkpoint complex

MDM4:

murine double minute 4

MMA:

monomethyl arginine

MMS4:

DNA repair endonuclease

MRE11:

meiotic recombination 11

MRN:

MRE11/Rad51/Nbs1 complex

MUS81:

DNA repair endonuclease

NBS1:

nibrin

NHEJ:

non-homologous end joining

p53:

tumor suppressor protein p53

p53BP1:

p53-binding protein 1

PCNA:

proliferating cell nuclear antigen

PGM:

proline glycine methionine motif

PI3K:

phosphatidyl inositol-3-kinase

PKC:

protein kinase C

pRb:

retinoblastoma protein

PRMT:

protein arginine methyltransferase

PSA:

prostate specific antigen

PTEN:

phosphatase and tensin homolog

R:

arginine

Rad51:

DNA recombination protein

RFC:

replication factor C

RNA:

ribonucleic acid

RPA:

DNA repair protein

S phase:

DNA synthesis phase of cell cycle

SAC:

spindle assembly checkpoint

SDMA:

symmetric dimethylarginine

SF:

splicing factor

SR:

serine–arginine rich

TFIIH:

transcription factor IIH

TRF:

telomere repeat binding factor

UBAP2L:

ubiquitin-associated protein 2-like

XPA:

DNA repair protein

XPC:

DNA repair gene

XPF:

DNA repair endonuclease

XPG:

DNA repair endonuclease

XRCC:

DNA base excision repair gene

References

  1. Vermeulen K, Van Bockstaele DR, Berneman ZN. The cell cycle: a review of regulation, deregulation and therapeutic targets in cancer. Cell Prolif. 2003;36:131–49.

    Article  CAS  PubMed  Google Scholar 

  2. Manchado E, Guillamot M, Malumbres M. Killing cells by targeting mitosis. Cell Death Differ. 2012;19:369.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Park M-T, Lee S-J. Cell cycle and cancer. J Biochem Mol Biol. 2003;36:60–5.

    CAS  PubMed  Google Scholar 

  4. Yang Y, Bedford MT. Protein arginine methyltransferases and cancer. Nat Rev Cancer. 2013;13:37–50.

    Article  CAS  PubMed  Google Scholar 

  5. Norbury C, Nurse P. Animal cell cycles and their control. Annu Rev Biochem. 1992;61:441–68.

    Article  CAS  PubMed  Google Scholar 

  6. Schafer K. The cell cycle: a review. Vet Pathol Online. 1998;35:461–78.

    Article  CAS  Google Scholar 

  7. Burgess A, Vuong J, Rogers S, Malumbres M, O’Donoghue SI. SnapShot: phosphoregulation of mitosis. Cell. 2017;169:1358.

    Article  CAS  PubMed  Google Scholar 

  8. Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. Cell. 2011;144:646–74.

    Article  CAS  PubMed  Google Scholar 

  9. Collins I, Garrett MD. Targeting the cell division cycle in cancer: CDK and cell cycle checkpoint kinase inhibitors. Curr Opin Pharmacol. 2005;5:366–73.

    Article  CAS  PubMed  Google Scholar 

  10. Maddika S, Ande SR, Panigrahi S, Paranjothy T, Weglarczyk K, Zuse A, Eshraghi M, Manda KD, Wiechec E, Los M. Cell survival, cell death and cell cycle pathways are interconnected: implications for cancer therapy. Drug Resist Updates. 2007;10:13–29.

    Article  CAS  Google Scholar 

  11. Morgan DO. Principles of CDK regulation. Nature. 1995;374:131–4.

    Article  CAS  PubMed  Google Scholar 

  12. Pines J. Cyclins and cyclin-dependent kinases: theme and variations. Adv Cancer Res. 1995;66:181–212.

    Article  CAS  PubMed  Google Scholar 

  13. Evans T, Rosenthal ET, Youngblom J, Distel D, Hunt T. Cyclin: a protein specified by maternal mRNA in sea urchin eggs that is destroyed at each cleavage division. Cell. 1983;33:389–96.

    Article  CAS  PubMed  Google Scholar 

  14. Massague J. G1 cell-cycle control and cancer. Nature. 2004;432:298–306.

    Article  CAS  PubMed  Google Scholar 

  15. Malumbres M, Carnero A. Cell cycle deregulation: a common motif in cancer. Prog Cell Cycle Res. 2003;5:5–18.

    PubMed  Google Scholar 

  16. Dobashi Y. Cell cycle regulation and its aberrations in human lung carcinoma. Pathol Int. 2005;55:95–105.

    Article  CAS  PubMed  Google Scholar 

  17. Jeffrey PD, Russo AA, Polyak K, Gibbs E, Hurwitz J, Massague J, Pavletich NP. Mechanism of CDK activation revealed by the structure of a cyclinA-CDK2 complex. Nature. 1995;376:313–20.

    Article  CAS  PubMed  Google Scholar 

  18. Lew DJ, Kornbluth S. Regulatory roles of cyclin dependent kinase phosphorylation in cell cycle control. Curr Opin Cell Biol. 1996;8:795–804.

    Article  CAS  PubMed  Google Scholar 

  19. Hirai H, Roussel MF, Kato JY, Ashmun RA, Sherr CJ. Novel INK4 proteins, p19 and p18, are specific inhibitors of the cyclin D-dependent kinases CDK4 and CDK6. Mol Cell Biol. 1995;15:2672–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Polyak K, Kato JY, Solomon MJ, Sherr CJ, Massague J, Roberts JM, Koff A. p27Kip1, a cyclin-Cdk inhibitor, links transforming growth factor-beta and contact inhibition to cell cycle arrest. Genes Dev. 1994;8:9–22.

    Article  CAS  PubMed  Google Scholar 

  21. Waga S, Li R, Stillman B. p53-induced p21 controls DNA replication. Leukemia. 1997;11(Suppl 3):321–3.

    PubMed  Google Scholar 

  22. el-Deiry WS, Tokino T, Velculescu VE, Levy DB, Parsons R, Trent JM, Lin D, Mercer WE, Kinzler KW, Vogelstein B. WAF1, a potential mediator of p53 tumor suppression. Cell. 1993;75:817–25.

    Article  CAS  PubMed  Google Scholar 

  23. Zhang Z, Rosen DG, Yao JL, Huang J, Liu J. Expression of p14ARF, p15INK4b, p16INK4a, and DCR2 increases during prostate cancer progression. Mod Pathol. 2006;19:1339.

    Article  CAS  PubMed  Google Scholar 

  24. Liang J, Slingerland JM. Multiple roles of the PI3K/PKB (Akt) pathway in cell cycle progression. Cell Cycle. 2003;2:339–45.

    Article  CAS  PubMed  Google Scholar 

  25. Diehl JA, Cheng M, Roussel MF, Sherr CJ. Glycogen synthase kinase-3beta regulates cyclin D1 proteolysis and subcellular localization. Genes Dev. 1998;12:3499–511.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Vivanco I, Sawyers CL. The phosphatidylinositol 3-kinase–AKT pathway in human cancer. Nat Rev Cancer. 2002;2:489–501.

    Article  CAS  PubMed  Google Scholar 

  27. Stambolic V, Suzuki A, De La Pompa JL, Brothers GM, Mirtsos C, Sasaki T, Ruland J, Penninger JM, Siderovski DP, Mak TW. Negative regulation of PKB/Akt-dependent cell survival by the tumor suppressor PTEN. Cell. 1998;95:29–39.

    Article  CAS  PubMed  Google Scholar 

  28. Ali IU, Schriml LM, Dean M. Mutational spectra of PTEN/MMAC1 gene: a tumor suppressor with lipid phosphatase activity. J Natl Cancer Inst. 1999;91:1922–32.

    Article  CAS  PubMed  Google Scholar 

  29. Malumbres M, Barbacid M. Cell cycle, CDKs and cancer: a changing paradigm. Nat Rev Cancer. 2009;9:153–66.

    Article  CAS  PubMed  Google Scholar 

  30. Levine AJ. p53, the cellular gatekeeper for growth and division. Cell. 1997;88:323–31.

    Article  CAS  PubMed  Google Scholar 

  31. Paulovich AG, Hartwell LH. A checkpoint regulates the rate of progression through S phase in S. cerevisiae in response to DNA damage. Cell. 1995;82:841–7.

    Article  CAS  PubMed  Google Scholar 

  32. Kops GJ, Weaver BA, Cleveland DW. On the road to cancer: aneuploidy and the mitotic checkpoint. Nat Rev Cancer. 2005;5:773–85.

    Article  CAS  PubMed  Google Scholar 

  33. Hoyt MA, Totis L, Roberts BTS. S. cerevisiae genes required for cell cycle arrest in response to loss of microtubule function. Cell. 1991;66:507–17.

    Article  CAS  PubMed  Google Scholar 

  34. Li R, Murray AW. Feedback control of mitosis in budding yeast. Cell. 1991;66:519–31.

    Article  CAS  PubMed  Google Scholar 

  35. Sudakin V, Chan GK, Yen TJ. Checkpoint inhibition of the APC/C in HeLa cells is mediated by a complex of BUBR1, BUB3, CDC20, and MAD2. J Cell Biol. 2001;154:925–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Fang G, Yu H, Kirschner MW. The checkpoint protein MAD2 and the mitotic regulator CDC20 form a ternary complex with the anaphase-promoting complex to control anaphase initiation. Genes Dev. 1998;12:1871–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Shannon KB, Canman JC, Salmon ED. Mad2 and BubR1 function in a single checkpoint pathway that responds to a loss of tension. Mol Biol Cell. 2002;13:3706–19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Malureanu LA, Jeganathan KB, Hamada M, Wasilewski L, Davenport J, van Deursen JM. BubR1N terminus acts as a soluble inhibitor of cyclin B degradation by APC/CCdc20 in interphase. Dev Cell. 2009;16:118–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Hagting A, den Elzen N, Vodermaier HC, Waizenegger IC, Peters J-M, Pines J. Human securin proteolysis is controlled by the spindle checkpoint and reveals when the APC/C switches from activation by Cdc20 to Cdh1. J Cell Biol. 2002;157:1125–37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Hauf S, Waizenegger IC, Peters JM. Cohesin cleavage by separase required for anaphase and cytokinesis in human cells. Science. 2001;293:1320–3.

    Article  CAS  PubMed  Google Scholar 

  41. D’Angiolella V, Mari C, Nocera D, Rametti L, Grieco D. The spindle checkpoint requires cyclin-dependent kinase activity. Genes Dev. 2003;17:2520–5.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  42. Ditchfield C, Johnson VL, Tighe A, Ellston R, Haworth C, Johnson T, Mortlock A, Keen N, Taylor SS. Aurora B couples chromosome alignment with anaphase by targeting BubR1, Mad2, and Cenp-E to kinetochores. J Cell Biol. 2003;161:267–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Hauf S, Cole RW, LaTerra S, Zimmer C, Schnapp G, Walter R, Heckel A, van Meel J, Rieder CL, Peters JM. The small molecule Hesperadin reveals a role for Aurora B in correcting kinetochore-microtubule attachment and in maintaining the spindle assembly checkpoint. J Cell Biol. 2003;161:281–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Maeda M, Hasegawa H, Sugiyama M, Hyodo T, Ito S, Chen D, Asano E, Masuda A, Hasegawa Y, Hamaguchi M, et al. Arginine methylation of ubiquitin-associated protein 2-like is required for the accurate distribution of chromosomes. FASEB. 2015;30:312–23.

    Article  CAS  Google Scholar 

  45. Musacchio A, Salmon ED. The spindle-assembly checkpoint in space and time. Nat Rev Mol Cell Biol. 2007;8:379–93.

    Article  CAS  PubMed  Google Scholar 

  46. Sancar A, Lindsey-Boltz LA, Unsal-Kacmaz K, Linn S. Molecular mechanisms of mammalian DNA repair and the DNA damage checkpoints. Annu Rev Biochem. 2004;73:39–85.

    Article  CAS  PubMed  Google Scholar 

  47. Abraham RT. Cell cycle checkpoint signaling through the ATM and ATR kinases. Genes Dev. 2001;15:2177–96.

    Article  CAS  PubMed  Google Scholar 

  48. Branzei D, Foiani M. Regulation of DNA repair throughout the cell cycle. Nat Rev Mol Cell Biol. 2008;9:297–308.

    Article  CAS  PubMed  Google Scholar 

  49. Kastan MB, Bartek J. Cell-cycle checkpoints and cancer. Nature. 2004;432:316–23.

    Article  CAS  PubMed  Google Scholar 

  50. Kunkel TA, Erie DA. DNA mismatch repair. Annu Rev Biochem. 2005;74:681–710.

    Article  CAS  PubMed  Google Scholar 

  51. Krokan H, Standal R, Slupphaug G. DNA glycosylases in the base excision repair of DNA. Biochem J. 1997;325:1–16.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Demple B, Jacobsson A, Olsson M, Robins P, Lindahl T. Repair of alkylated DNA in Escherichia coli. Physical properties of O6-methylguanine-DNA methyltransferase. J Biol Chem. 1982;257:13776–80.

    CAS  PubMed  Google Scholar 

  53. Kubota Y, Nash RA, Klungland A, Schär P, Barnes D, Lindahl T. Reconstitution of DNA base excision-repair with purified human proteins: interaction between DNA polymerase beta and the XRCC1 protein. EMBO J. 1996;15:6662.

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Mayanagi K, Kiyonari S, Saito M, Shirai T, Ishino Y, Morikawa K. Mechanism of replication machinery assembly as revealed by the DNA ligase–PCNA–DNA complex architecture. Proc Natl Acad Sci. 2009;106:4647–52.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. de Laat WL, Jaspers NG, Hoeijmakers JH. Molecular mechanism of nucleotide excision repair. Genes Dev. 1999;13:768–85.

    Article  PubMed  Google Scholar 

  56. Sancar A, Tang MS. Nucleotide excision repair. Photochem Photobiol. 1993;57:905–21.

    Article  CAS  PubMed  Google Scholar 

  57. O’Donovan A, Davies AA, Moggs JG, West SC, Wood RD. XPG endonuclease makes the 3′ incision in human DNA nucleotide excision repair. Nature. 1994;371:432.

    Article  PubMed  Google Scholar 

  58. He Z, Henricksen LA, Wold MS, Ingles CJ. RPA involvement in the damage-recognition and incision steps of nucleotide excision repair. Nature. 1995;374:566–9.

    Article  CAS  PubMed  Google Scholar 

  59. Araújo SJ, Nigg EA, Wood RD. Strong functional interactions of TFIIH with XPC and XPG in human DNA nucleotide excision repair, without a preassembled repairosome. Mol Cell Biol. 2001;21:2281–91.

    Article  PubMed  PubMed Central  Google Scholar 

  60. Chu G. Double strand break repair. J Biol Chem. 1997;272:24097–100.

    Article  CAS  PubMed  Google Scholar 

  61. Ferguson DO, Alt FW. DNA double strand break repair and chromosomal translocation: lessons from animal models. Oncogene. 2001;20:5572–9.

    Article  CAS  PubMed  Google Scholar 

  62. Weterings E, Chen DJ. The endless tale of non-homologous end-joining. Cell Res. 2008;18:114–24.

    Article  CAS  PubMed  Google Scholar 

  63. Tauchi H, Kobayashi J, Morishima K, van Gent DC, Shiraishi T, Verkaik NS, Ito E, Nakamura A, Sonoda E, Takata M. Nbs1 is essential for DNA repair by homologous recombination in higher vertebrate cells. Nature. 2002;420:93–8.

    Article  CAS  PubMed  Google Scholar 

  64. D’Amours D, Jackson SP. The Mre11 complex: at the crossroads of DNA repair and checkpoint signalling. Nat Rev Mol Cell Biol. 2002;3:317–27.

    Article  PubMed  CAS  Google Scholar 

  65. Baumann P, West SC. Role of the human RAD51 protein in homologous recombination and double-stranded-break repair. Trends Biochem Sci. 1998;23:247–51.

    Article  CAS  PubMed  Google Scholar 

  66. Davies AA, Masson J-Y, McIlwraith MJ, Stasiak AZ, Stasiak A, Venkitaraman AR, West SC. Role of BRCA2 in control of the RAD51 recombination and DNA repair protein. Mol Cell. 2001;7:273–82.

    Article  CAS  PubMed  Google Scholar 

  67. Lees-Miller S, Meek K. Repair of DNA double strand breaks by non-homologous end joining. Biochimie. 2003;85:1161–73.

    Article  CAS  PubMed  Google Scholar 

  68. Dronkert ML, Kanaar R. Repair of DNA interstrand cross-links. Mutat Res DNA Repair. 2001;486:217–47.

    Article  CAS  PubMed  Google Scholar 

  69. McHugh PJ, Spanswick VJ, Hartley JA. Repair of DNA interstrand crosslinks: molecular mechanisms and clinical relevance. Lancet Oncol. 2001;2:483–90.

    Article  CAS  PubMed  Google Scholar 

  70. Kuraoka I, Kobertz WR, Ariza RR, Biggerstaff M, Essigmann JM, Wood RD. Repair of an interstrand DNA cross-link initiated by ERCC1-XPF repair/recombination nuclease. J Biol Chem. 2000;275:26632–6.

    Article  CAS  PubMed  Google Scholar 

  71. Finlan LE, Hupp TR. The life cycle of p53: a key target in drug development. Apoptotic pathways as targets for novel therapies in cancer and other diseases. Berlin: Springer; 2005. p. 157–72.

    Book  Google Scholar 

  72. Peng CY, Graves PR, Thoma RS, Wu Z, Shaw AS, Piwnica-Worms H. Mitotic and G2 checkpoint control: regulation of 14-3-3 protein binding by phosphorylation of Cdc25C on serine-216. Science. 1997;277:1501–5.

    Article  CAS  PubMed  Google Scholar 

  73. Kim R, Emi M, Tanabe K. Role of mitochondria as the gardens of cell death. Cancer Chemother Pharmacol. 2006;57:545–53.

    Article  CAS  PubMed  Google Scholar 

  74. Giono LE, Manfredi JJ. The p53 tumor suppressor participates in multiple cell cycle checkpoints. J Cell Physiol. 2006;209:13–20.

    Article  CAS  PubMed  Google Scholar 

  75. Siliciano JD, Canman CE, Taya Y, Sakaguchi K, Appella E, Kastan MB. DNA damage induces phosphorylation of the amino terminus of p53. Genes Dev. 1997;11:3471–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Hanahan D, Weinberg RA. The hallmarks of cancer. Cell. 2000;100:57–70.

    Article  CAS  PubMed  Google Scholar 

  77. Knudsen ES, Knudsen KE. Retinoblastoma tumor suppressor: where cancer meets the cell cycle. Exp Biol Med. 2006;231:1271–81.

    Article  Google Scholar 

  78. Lukas J, Parry D, Aagaard L, Mann DJ, Bartkova J, Strauss M, Peters G, Bartek J. Retinoblastoma-protein-dependent cell-cycle inhibition by the tumour suppressor p16. Nature. 1995;375:503–6.

    Article  CAS  PubMed  Google Scholar 

  79. Chellappan SP, Hiebert S, Mudryj M, Horowitz JM, Nevins JR. The E2F transcription factor is a cellular target for the RB protein. Cell. 1991;65:1053–61.

    Article  CAS  PubMed  Google Scholar 

  80. Connell-Crowley L, Harper JW, Goodrich D. Cyclin D1/Cdk4 regulates retinoblastoma protein-mediated cell cycle arrest by site-specific phosphorylation. Mol Biol Cell. 1997;8:287–301.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Buchkovich K, Duffy LA, Harlow E. The retinoblastoma protein is phosphorylated during specific phases of the cell cycle. Cell. 1989;58:1097–105.

    Article  CAS  PubMed  Google Scholar 

  82. Bedford MT, Clarke S. Protein arginine methylation in mammals: who, what, and why. Mol Cell. 2009;33:1–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Lee YH, Stallcup MR. Minireview: protein arginine methylation of nonhistone proteins in transcriptional regulation. Mol Endocrinol. 2009;23:425.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Paik WK, Paik DC, Kim S. Historical review: the field of protein methylation. Trends Biochem Sci. 2007;32:146–52.

    Article  CAS  PubMed  Google Scholar 

  85. Altschuler L, Wook J-O, Gurari D, Chebath J, Revel M. Involvement of receptor-bound protein methyltransferase PRMT1 in antiviral and antiproliferative effects of type I interferons. J Interferon Cytokine Res. 1999;19:189–95.

    Article  CAS  PubMed  Google Scholar 

  86. Cheng D, Côté J, Shaaban S, Bedford MT. The arginine methyltransferase CARM1 regulates the coupling of transcription and mRNA processing. Mol Cell. 2007;25:71–83.

    Article  PubMed  CAS  Google Scholar 

  87. Carr SM, Poppy Roworth A, Chan C, La Thangue NB. Post-translational control of transcription factors: methylation ranks highly. FEBS J. 2015;282:4450–65.

    Article  CAS  PubMed  Google Scholar 

  88. Côté J, Richard S. Tudor domains bind symmetrical dimethylated arginines. J Biol Chem. 2005;280:28476–83.

    Article  PubMed  CAS  Google Scholar 

  89. Pahlich S, Zakaryan RP, Gehring H. Protein arginine methylation: cellular functions and methods of analysis. BBA Proteins Proteom. 2006;1764:1890–903.

    Article  CAS  Google Scholar 

  90. Friesen WJ, Paushkin S, Wyce A, Massenet S, Pesiridis GS, Van Duyne G, Rappsilber J, Mann M, Dreyfuss G. The methylosome, a 20S complex containing JBP1 and pICln, produces dimethylarginine-modified Sm proteins. Mol Cell Biol. 2001;21:8289–300.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Hadjikyriacou A, Yang Y, Espejo A, Bedford MT, Clarke SG. Unique features of human protein arginine methyltransferase 9 (PRMT9) and its substrate RNA splicing factor SF3B2. J Biol Chem. 2015;290:16723–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Blanc RS, Richard S. Arginine methylation: the coming of age. Mol Cell. 2017;65:8–24.

    Article  CAS  PubMed  Google Scholar 

  93. Paik WK, Kim S. Protein methylase I. Purification and properties of the enzyme. J Biol Chem. 1968;243:2108–14.

    CAS  PubMed  Google Scholar 

  94. Wei H, Mundade R, Lange KC, Lu T. Protein arginine methylation of non-histone proteins and its role in diseases. Cell Cycle. 2014;13:32–41.

    Article  PubMed  CAS  Google Scholar 

  95. Mirto L, Piller S. Protein arginine methylation: a pivotal factor and a target for novel HIV-therapeutics, other viruses and multiple diseases. HIV Ther. 2010;4:65–81.

    Article  CAS  Google Scholar 

  96. Bedford MT. Arginine methylation at a glance. J Cell Sci. 2007;120:4243–6.

    Article  CAS  PubMed  Google Scholar 

  97. Peng C, Wong CC. The story of protein arginine methylation: characterization, regulation, and function. Exp Rev Proteom. 2017;14:157–70.

    Article  CAS  Google Scholar 

  98. Paik WK, Kim S. Natural occurrence of various methylated amino acid derivatives. In: Meister A, editor. Protein Methylation. New York: Wiley; 1980. p. 8–25.

    Google Scholar 

  99. Zurita-Lopez CI, Sandberg T, Kelly R, Clarke SG. Human protein arginine methyltransferase 7 (PRMT7) is a type III enzyme forming ω-NG-monomethylated arginine residues. J Biol Chem. 2012;287:7859–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Fisk JC, Read LK. Protein arginine methylation in parasitic protozoa. Eukaryot Cell. 2011;10:1013–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Kim C, Lim Y, Yoo BC, Won NH, Kim S, Kim G. Regulation of post-translational protein arginine methylation during HeLa cell cycle. Biochem Biophys Acta. 2010;1800:977–85.

    Article  CAS  PubMed  Google Scholar 

  102. Tang J, Frankel A, Cook RJ, Kim S, Paik WK, Williams KR, Clarke S, Herschman HR. PRMT1 is the predominant type I protein arginine methyltransferase in mammalian cells. J Biol Chem. 2000;275:7723–30.

    Article  CAS  PubMed  Google Scholar 

  103. Dolezal E, Infantino S, Drepper F, Borsig T, Singh A, Wossning T, Fiala GJ, Minguet S, Warscheid B, Tarlinton DM, et al. The BTG2-PRMT1 module limits pre-B cell expansion by regulating the CDK4-Cyclin-D3 complex. Nat Immunol. 2017;18:911–20.

    Article  CAS  PubMed  Google Scholar 

  104. Zheng S, Moehlenbrink J, Lu Y-C, Zalmas L-P, Sagum CA, Carr S, McGouran JF, Alexander L, Fedorov O, Munro S, et al. Arginine methylation-dependent reader-writer interplay governs growth control by E2F-1. Mol Cell. 2013;52:37–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Glorian V, Allègre J, Berthelet J, Dumetier B, Boutanquoi P-M, Droin N, Kayaci C, Cartier J, Gemble S, Marcion G, et al. DNA damage and S phase-dependent E2F1 stabilization requires the cIAP1 E3-ubiquitin ligase and is associated with K63-poly-ubiquitination on lysine 161/164 residues. Cell Death Dis. 2017;8:e2816.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Cho EC, Zheng S, Munro S, Liu G, Carr SM, Moehlenbrink J, Lu YC, Stimson L, Khan O, Konietzny R. Arginine methylation controls growth regulation by E2F‐1. EMBO J. 2012;31:1785–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Lu Y, Ma W, Li Z, Lu J, Wang X. The interplay between p16 serine phosphorylation and arginine methylation determines its function in modulating cellular apoptosis and senescence. Sci Rep. 2017;7:41390.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Yoshimoto T, Boehm M, Olive M, Crook MF, San H, Langenickel T, Nabel EG. The arginine methyltransferase PRMT2 binds RB and regulates E2F function. Exp Cell Res. 2006;312:2040–53.

    Article  CAS  PubMed  Google Scholar 

  109. Kim KY, Wang D-H, Campbell M, Huerta SB, Shevchenko B, Izumiya C, Izumiya Y. PRMT4-mediated arginine methylation negatively regulates retinoblastoma tumor suppressor protein and promotes E2F-1 dissociation. Mol Cell Biol. 2015;35:238–48.

    Article  PubMed  CAS  Google Scholar 

  110. Oh TG, Bailey P, Dray E, Smith AG, Goode J, Eriksson N, Funder JW, Fuller PJ, Simpson ER, Tilley WD. PRMT2 and RORγ expression are associated with breast cancer survival outcomes. Mol Endocrinol. 2014;28:1166–85.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  111. Zhong J, Cao RX, Liu JH, Liu YB, Wang J, Liu LP, Chen YJ, Yang J, Zhang QH, Wu Y, et al. Nuclear loss of protein arginine N-methyltransferase 2 in breast carcinoma is associated with tumor grade and overexpression of cyclin D1 protein. Oncogene. 2014;33:5546.

    Article  CAS  PubMed  Google Scholar 

  112. El Messaoudi S, Fabbrizio E, Rodriguez C, Chuchana P, Fauquier L, Cheng D, Theillet C, Vandel L, Bedford MT, Sardet C. Coactivator-associated arginine methyltransferase 1 (CARM1) is a positive regulator of the Cyclin E1 gene. Proc Natl Acad Sci. 2006;103:13351–6.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  113. Frietze S, Lupien M, Silver PA, Brown M. CARM1 regulates estrogen-stimulated breast cancer growth through up-regulation of E2F1. Cancer Res. 2008;68:301–6.

    Article  CAS  PubMed  Google Scholar 

  114. Hong H, Kao C, Jeng MH, Eble JN, Koch MO, Gardner TA, Zhang S, Li L, Pan CX, Hu Z. Aberrant expression of CARM1, a transcriptional coactivator of androgen receptor, in the development of prostate carcinoma and androgen-independent status. Cancer. 2004;101:83–9.

    Article  CAS  PubMed  Google Scholar 

  115. Pal S, Vishwanath SN, Erdjument-Bromage H, Tempst P, Sif S. Human SWI/SNF-associated PRMT5 methylates histone H3 arginine 8 and negatively regulates expression of ST7 and NM23 tumor suppressor genes. Mol Cell Biol. 2004;24:9630–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Wei TY, Juan CC, Hisa JY, Su LJ, Lee YC, Chou HY, Chen JM, Wu YC, Chiu SC, Hsu CP, et al. Protein arginine methyltransferase 5 is a potential oncoprotein that upregulates G1 cyclins/cyclin-dependent kinases and the phosphoinositide 3-kinase/AKT signaling cascade. Cancer Sci. 2012;103:1640–50.

    Article  CAS  PubMed  Google Scholar 

  117. Scoumanne A, Zhang J, Chen X. PRMT5 is required for cell-cycle progression and p53 tumor suppressor function. Nucleic Acids Res. 2009;37:4965–76.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Yan F, Alinari L, Lustberg ME, Martin LK, Cordero-Nieves HM, Banasavadi-Siddegowda Y, Virk S, Barnholtz-Sloan J, Bell EH, Wojton J, et al. Genetic validation of the protein arginine methyltransferase PRMT5 as a candidate therapeutic target in glioblastoma. Cancer Res. 2014;74:1752–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Yang H, Zhao X, Zhao L, Liu L, Li J, Jia W, Liu J, Huang G. PRMT5 competitively binds to CDK4 to promote G1-S transition upon glucose induction in hepatocellular carcinoma. Oncotarget. 2016;7:72131–47.

    PubMed  PubMed Central  Google Scholar 

  120. Banasavadi-Siddegowda Y, Russell L, Frair E, Karkhanis V, Relation T, Yoo J, Zhang J, Sif S, Imitola J, Baiocchi R. PRMT5–PTEN molecular pathway regulates senescence and self-renewal of primary glioblastoma neurosphere cells. Oncogene. 2017;36:263.

    Article  CAS  PubMed  Google Scholar 

  121. Kumar B, Yadav A, Brown NV, Zhao S, Cipolla MJ, Wakely PE, Schmitt AC, Baiocchi RA, Teknos TN, Old M, et al. Nuclear PRMT5, cyclin D1 and IL-6 are associated with poor outcome in oropharyngeal squamous cell carcinoma patients and is inversely associated with p16-status. Oncotarget. 2017;8:14847–59.

    PubMed  PubMed Central  Google Scholar 

  122. Han X, Li R, Zhang W, Yang X, Wheeler CG, Friedman GK, Province P, Ding Q, You Z, Fathallah-Shaykh HM, et al. Expression of PRMT5 correlates with malignant grade in gliomas and plays a pivotal role in tumor growth in vitro. J Neurooncol. 2014;118:61–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Gu Z, Gao S, Zhang F, Wang Z, Ma W, Davis RE, Wang Z. Protein arginine methyltransferase 5 is essential for growth of lung cancer cells. Biochem J. 2012;446:235.

    Article  CAS  PubMed  Google Scholar 

  124. Pal S, Baiocchi RA, Byrd JC, Grever MR, Jacob ST, Sif S. Low levels of miR-92b/96 induce PRMT5 translation and H3R8/H4R3 methylation in mantle cell lymphoma. EMBO J. 2007;26:3558–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Bao X, Zhao S, Liu T, Liu Y, Liu Y, Yang X. Overexpression of PRMT5 promotes tumor cell growth and is associated with poor disease prognosis in epithelial ovarian cancer. J Histochem Cytochem. 2013;61:206–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Gu Z, Li Y, Lee P, Liu T, Wan C, Wang Z. Protein arginine methyltransferase 5 functions in opposite ways in the cytoplasm and nucleus of prostate cancer cells. PLoS ONE. 2012;7:e44033.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Jansson M, Durant ST, Cho E-C, Sheahan S, Edelmann M, Kessler B, La Thangue NB. Arginine methylation regulates the p53 response. Nat Cell Biol. 2008;10:1431–9.

    Article  CAS  PubMed  Google Scholar 

  128. Jung SY, Li Y, Wang Y, Chen Y, Zhao Y, Qin J. Complications in the assignment of 14 and 28 Da mass shift detected by mass spectrometry as in vivo methylation from endogenous proteins. Anal Chem. 2008;80:1721–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Zhang Y, Zhang YJ, Zhao HY, Zhai QL, Zhang Y, Shen YF. The impact of R213 mutation on p53-mediated p21 activity. Biochimie. 2014;99:215–8.

    Article  CAS  PubMed  Google Scholar 

  130. Cheung RS, Taniguchi T. Recent insights into the molecular basis of Fanconi anemia: genes, modifiers, and drivers. Int J Hematol. 2017;106:335–44.

    Article  CAS  PubMed  Google Scholar 

  131. Du W, Amarachintha S, Erden O, Wilson A, Pang Q. The Fanconi anemia pathway controls oncogenic response in hematopoietic stem and progenitor cells by regulating PRMT5-mediated p53 arginine methylation. Oncotarget. 2016;7:60005–20.

    PubMed  PubMed Central  Google Scholar 

  132. Kleinschmidt MA, De Graaf P, Van Teeffelen HA, Timmers HTM. Cell cycle regulation by the PRMT6 arginine methyltransferase through repression of cyclin-dependent kinase inhibitors. PLoS ONE. 2012;7:e41446.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Stein C, Riedl S, Rüthnick D, Nötzold RR, Bauer U-M. The arginine methyltransferase PRMT6 regulates cell proliferation and senescence through transcriptional repression of tumor suppressor genes. Nucleic Acids Res. 2012;40:9522–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Phalke S, Mzoughi S, Bezzi M, Jennifer N, Mok WC, Low DH, Thike AA, Kuznetsov VA, Tan PH, Voorhoeve PM. p53-Independent regulation of p21Waf1/Cip1 expression and senescence by PRMT6. Nucleic Acids Res. 2012;40:9534–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Nakakido M, Deng Z, Suzuki T, Dohmae N, Nakamura Y, Hamamoto R. PRMT6 increases cytoplasmic localization of p21CDKN1A in cancer cells through arginine methylation and makes more resistant to cytotoxic agents. Oncotarget. 2015;6:30957–67.

    Article  PubMed  PubMed Central  Google Scholar 

  136. Wang X, Huang Y, Zhao J, Zhang Y, Lu J, Huang B. Suppression of PRMT6-mediated arginine methylation of p16 protein potentiates its ability to arrest A549 cell proliferation. Int J Biochem Cell Biol. 2012;44:2333–41.

    Article  CAS  PubMed  Google Scholar 

  137. Gurunathan G, Yu Z, Coulombe Y, Masson JY, Richard S. Arginine methylation of hnRNPUL1 regulates interaction with NBS1 and recruitment to sites of DNA damage. Sci Rep. 2015;5:10475.

    Article  PubMed  PubMed Central  Google Scholar 

  138. Boisvert FM, Hendzel MJ, Masson JY, Richard S. Methylation of MRE11 regulates its nuclear compartmentalization. Cell Cycle. 2005;4:981–9.

    Article  CAS  PubMed  Google Scholar 

  139. Dery U, Coulombe Y, Rodrigue A, Stasiak A, Richard S, Masson JY. A glycine–arginine domain in control of the human MRE11 DNA repair protein. Mol Cell Biol. 2008;28:3058–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Boisvert F-M, Rhie A, Richard S, Doherty AJ. The GAR motif of 53BP1 is arginine methylated by PRMT1 and is necessary for 53BP1 DNA binding activity. Cell Cycle. 2005;4:1834–41.

    Article  CAS  PubMed  Google Scholar 

  141. Yu Z, Chen T, Hébert J, Li E, Richard S. A mouse PRMT1 null allele defines an essential role for arginine methylation in genome maintenance and cell proliferation. Mol Cell Biol. 2009;29:2982–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. El-Andaloussi N, Valovka T, Toueille M, Hassa PO, Gehrig P, Covic M, Hübscher U, Hottiger MO. Methylation of DNA polymerase β by protein arginine methyltransferase 1 regulates its binding to proliferating cell nuclear antigen. FASEB J. 2007;21:26–34.

    Article  CAS  PubMed  Google Scholar 

  143. He W, Ma X, Yang X, Zhao Y, Qiu J, Hang H. A role for the arginine methylation of Rad9 in checkpoint control and cellular sensitivity to DNA damage. Nucleic Acids Res. 2011;39:4719–27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. El-Andaloussi N, Valovka T, Toueille M, Steinacher R, Focke F, Gehrig P, Covic M, Hassa PO, Schär P, Hübscher U. Arginine methylation regulates DNA polymerase β. Mol Cell. 2006;22:51–62.

    Article  CAS  PubMed  Google Scholar 

  145. Guo Z, Zheng L, Xu H, Dai H, Zhou M, Pascua MR, Chen QM, Shen B. Methylation of FEN1 suppresses nearby phosphorylation and facilitates PCNA binding. Nat Chem Biol. 2010;6:766–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Mitchell TR, Glenfield K, Jeyanthan K, Zhu X-D. Arginine methylation regulates telomere length and stability. Mol Cell Biol. 2009;29:4918–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Deng X, Von Keudell G, Suzuki T, Dohmae N, Nakakido M, Piao L, Yoshioka Y, Nakamura Y, Hamamoto R. PRMT1 promotes mitosis of cancer cells through arginine methylation of INCENP. Oncotarget. 2015;6:35173–82.

    PubMed  PubMed Central  Google Scholar 

  148. Poulard C, Rambaud J, Hussein N, Corbo L, Le Romancer M. JMJD6 regulates ERα methylation on arginine. PLoS ONE. 2014;9:e87982.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  149. Qi C, Chang J, Zhu Y, Yeldandi AV, Rao SM, Zhu Y-J. Identification of protein arginine methyltransferase 2 as a coactivator for estrogen receptor α. J Biol Chem. 2002;277:28624–30.

    Article  CAS  PubMed  Google Scholar 

  150. Knudsen KE, Arden KC, Cavenee WK. Multiple G1 regulatory elements control the androgen-dependent proliferation of prostatic carcinoma cells. J Biol Chem. 1998;273:20213–22.

    Article  CAS  PubMed  Google Scholar 

  151. Shiota M, Takeuchi A, Yokomizo A, Kashiwagi E, Tatsugami K, Naito S. Methyltransferase inhibitor adenosine dialdehyde suppresses androgen receptor expression and prostate cancer growth. J Urol. 2012;188:300.

    Article  CAS  PubMed  Google Scholar 

  152. Meyer R, Wolf SS, Obendorf M. PRMT2, a member of the protein arginine methyltransferase family, is a coactivator of the androgen receptor. J Steroid Biochem Mol Biol. 2007;107:1–14.

    Article  CAS  PubMed  Google Scholar 

  153. Harada N, Takagi T, Nakano Y, Yamaji R, Inui H. Protein arginine methyltransferase 10 is required for androgen-dependent proliferation of LNCaP prostate cancer cells. Biosci Biotechnol Biochem. 2015;79:1–8.

    Article  CAS  Google Scholar 

  154. Deng X, Shao G, Zhang HT, Li C, Zhang D, Cheng L, Elzey BD, Pili R, Ratliff TL, Huang J, et al. Protein arginine methyltransferase 5 functions as an epigenetic activator of the androgen receptor to promote prostate cancer cell growth. Oncogene. 2017;36:1223–31.

    Article  CAS  PubMed  Google Scholar 

  155. Xu W, Chen H, Du K, Asahara H, Tini M, Emerson BM, Montminy M, Evans RM. A transcriptional switch mediated by cofactor methylation. Science. 2001;294:2507–11.

    Article  CAS  PubMed  Google Scholar 

  156. Chevillard-Briet M, Trouche D, Vandel L. Control of CBP co-activating activity by arginine methylation. EMBO J. 2002;21:5457–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Lee YH, Bedford MT, Stallcup MR. Regulated recruitment of tumor suppressor BRCA1 to the p21 gene by coactivator methylation. Genes Dev. 2011;25:176–88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Katzenberger RJ, Marengo MS, Wassarman DA. ATM and ATR pathways signal alternative splicing of Drosophila TAF1 pre-mRNA in response to DNA damage. Mol Cell Biol. 2006;26:9256–67.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Moore MJ, Wang Q, Kennedy CJ, Silver PA. An alternative splicing network links cell-cycle control to apoptosis. Cell. 2010;142:625–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Fischer D-C, Noack K, Runnebaum IB, Watermann DO, Kieback DG, Stamm S, Stickeler E. Expression of splicing factors in human ovarian cancer. Oncol Rep. 2004;11:1085–90.

    CAS  PubMed  Google Scholar 

  161. Karni R, de Stanchina E, Lowe SW, Sinha R, Mu D, Krainer AR. The gene encoding the splicing factor SF2/ASF is a proto-oncogene. Nat Struct Mol Biol. 2007;14:185–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  162. Stickeler E, Kittrell F, Medina D, Berget SM. Stage-specific changes in SR splicing factors and alternative splicing in mammary tumorigenesis. Oncogene. 1999;18:3574–82.

    Article  CAS  PubMed  Google Scholar 

  163. Liu Q, Dreyfuss G. In vivo and in vitro arginine methylation of RNA-binding proteins. Mol Cell Biol. 1995;15:2800–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Shen EC, Henry MF, Weiss VH, Valentini SR, Silver PA, Lee MS. Arginine methylation facilitates the nuclear export of hnRNP proteins. Genes Dev. 1998;12:679–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Chen Y, Zhou X, Liu N, Wang C, Zhang L, Mo W, Hu G. Arginine methylation of hnRNP K enhances p53 transcriptional activity. FEBS Lett. 2008;582:1761–5.

    Article  CAS  PubMed  Google Scholar 

  166. Ratnadiwakara M, Mohenska M, Anko ML. Splicing factors as regulators of miRNA biogenesis—links to human disease. Semin Cell Dev Biol. 2017. https://doi.org/10.1016/j.semcdb.2017.10.008

    PubMed  Google Scholar 

  167. Yanagida M, Hayano T, Yamauchi Y, Shinkawa T, Natsume T, Isobe T, Takahashi N. Human fibrillarin forms a sub-complex with splicing factor 2-associated p32, protein arginine methyltransferases, and tubulins alpha 3 and beta 1 that is independent of its association with preribosomal ribonucleoprotein complexes. J Biol Chem. 2004;279:1607–14.

    Article  CAS  PubMed  Google Scholar 

  168. Vhuiyan MI, Pak ML, Park MA, Thomas D, Lakowski TM, Chalfant CE, Frankel A. PRMT2 interacts with splicing factors and regulates the alternative splicing of BCL-X. J Biochem. 2017;162:17–25.

    PubMed  Google Scholar 

  169. Coˆté J, Boulanger M-C, Bedford MT, Richard S. Sam68 RNA binding protein is an in vivo substrate for protein arginine N-methyltransferase 1. Mol Biol Cell. 2003;14:274–87.

    Article  CAS  Google Scholar 

  170. Rho J, Choi S, Jung C-R, Im D-S. Arginine methylation of Sam68 and SLM proteins negatively regulates their poly (U) RNA binding activity. Arch Biochem Biophys. 2007;466:49–57.

    Article  CAS  PubMed  Google Scholar 

  171. Sinha R, Allemand E, Zhang Z, Karni R, Myers MP, Krainer AR. Arginine methylation controls the subcellular localization and functions of the oncoprotein splicing factor SF2/ASF. Mol Cell Biol. 2010;30:2762–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  172. Bezzi M, Teo SX, Muller J, Mok WC, Sahu SK, Vardy LA, Bonday ZQ, Guccione E. Regulation of constitutive and alternative splicing by PRMT5 reveals a role for Mdm4 pre-mRNA in sensing defects in the spliceosomal machinery. Genes Dev. 2013;27:1903–16.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Yu MC. The role of protein arginine methylation in mRNP dynamics. Mol Biol Int. 2011;2011:163827.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  174. Poulard C, Corbo L, Le Romancer M. Protein arginine methylation/demethylation and cancer. Oncotarget. 2016;7:67532–50.

    Article  PubMed  PubMed Central  Google Scholar 

  175. Yoshimatsu M, Toyokawa G, Hayami S, Unoki M, Tsunoda T, Field HI, Kelly JD, Neal DE, Maehara Y, Ponder BAJ, et al. Dysregulation of PRMT1 and PRMT6, type I arginine methyltransferases, is involved in various types of human cancers. Int J Cancer. 2011;128:562.

    Article  CAS  PubMed  Google Scholar 

  176. Mathioudaki K, Papadokostopoulou A, Scorilas A, Xynopoulos D, Agnanti N, Talieri M. The PRMT1 gene expression pattern in colon cancer. Br J Cancer. 2008;99:2094–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Baldwin RM, Bejide M, Trinkle-Mulcahy L, Cote J. Identification of the PRMT1v1 and PRMT1v2 specific interactomes by quantitative mass spectrometry in breast cancer cells. Proteomics. 2015;15:2187–97.

    Article  CAS  PubMed  Google Scholar 

  178. Zhong J, Cao RX, Zu XY, Hong T, Yang J, Liu L, Xiao XH, Ding WJ, Zhao Q, Liu JH, et al. Identification and characterization of novel spliced variants of PRMT2 in breast carcinoma. FEBS J. 2012;279:316–35.

    Article  CAS  PubMed  Google Scholar 

  179. Ohkura N, Takahashi M, Yaguchi H, Nagamura Y, Tsukada T. Coactivator-associated arginine methyltransferase 1, CARM1, affects pre-mRNA splicing in an isoform-specific manner. J Biol Chem. 2005;280:28927–35.

    Article  CAS  PubMed  Google Scholar 

  180. Sohail M, Zhang M, Litchfield D, Wang L, Kung S, Xie J. Differential expression, distinct localization and opposite effect on Golgi structure and cell differentiation by a novel splice variant of human PRMT5. BBA Mol Cell Res. 2015;1853:2444–52.

    CAS  Google Scholar 

  181. Gros L, Renodon-Corniere A, de Saint Vincent BR, Feder M, Bujnicki JM, Jacquemin-Sablon A. Characterization of prmt7alpha and beta isozymes from Chinese hamster cells sensitive and resistant to topoisomerase II inhibitors. Biochem Biophys Acta. 2006;1760:1646–56.

    Article  CAS  PubMed  Google Scholar 

  182. Hernandez S, Dominko T. Novel protein arginine methyltransferase 8 isoform is essential for cell proliferation. J Cell Biochem. 2016;117:2056–66.

    Article  CAS  PubMed  Google Scholar 

  183. Baldwin RM, Morettin A, Cote J. Role of PRMTs in cancer: could minor isoforms be leaving a mark? World J Biol Chem. 2014;5:115–29.

    PubMed  PubMed Central  Google Scholar 

  184. Parness J, Horwitz SB. Taxol binds to polymerized tubulin in vitro. J Cell Biol. 1981;91:479–87.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Piller S, Jwad N, Hejazi L, Gamsjaeger R, Sucher N. Protein arginine methylation of tubulin beta decreases binding of taxol in Neuro2a cells. FASEB J. 2015;29(717):16.

    Google Scholar 

  186. Pan ST, Li ZL, He ZX, Qiu JX, Zhou SF. Molecular mechanisms for tumour resistance to chemotherapy. Clin Exp Pharmacol Physiol. 2016;43:723–37.

    Article  CAS  PubMed  Google Scholar 

  187. Chen R, Alvero AB, Silasi DA, Mor G. Inflammation, cancer and chemoresistance: taking advantage of the toll-like receptor signaling pathway. Am J Reprod Immunol. 2007;57:93–107.

    Article  CAS  PubMed  Google Scholar 

  188. Sharom FJ. ABC multidrug transporters: structure, function and role in chemoresistance. Pharmacogenomics. 2008;9:105–27.

    Article  CAS  PubMed  Google Scholar 

  189. Soengas MS, Lowe SW. Apoptosis and melanoma chemoresistance. Oncogene. 2003;22:3138–51.

    Article  CAS  PubMed  Google Scholar 

  190. Itamochi H, Kigawa J, Terakawa N. Mechanisms of chemoresistance and poor prognosis in ovarian clear cell carcinoma. Cancer Sci. 2008;99:653–8.

    Article  CAS  PubMed  Google Scholar 

  191. Chen S, Guan X, Wang L-L, Li B, Sang X-B, Liu Y, Zhao Y. Fascaplysin inhibit ovarian cancer cell proliferation and metastasis through inhibiting CDK4. Gene. 2017;635:3–8.

    Article  CAS  PubMed  Google Scholar 

  192. Segraves NL, Robinson SJ, Garcia D, Said SA, Fu X, Schmitz FJ, Pietraszkiewicz H, Valeriote FA, Crews P. Comparison of fascaplysin and related alkaloids: a study of structures, cytotoxicities, and sources. J Nat Prod. 2004;67:783–92.

    Article  CAS  PubMed  Google Scholar 

  193. Lu X-L, Zheng Y-L, Chen H-M, Yan X-J, Wang F, Xu W-F. Anti-proliferation of human cervical cancer HeLa cell line by fascaplysin through apoptosis induction. Yao Xue Xue Bao. 2009;44:980–6.

    CAS  PubMed  Google Scholar 

  194. Hamilton G. Cytotoxic effects of fascaplysin against small cell lung cancer cell lines. Marine Drugs. 2014;12:1377–89.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  195. DeMichele A, Clark AS, Tan KS, Heitjan DF, Gramlich K, Gallagher M, Lal P, Feldman M, Zhang P, Colameco C, et al. CDK 4/6 inhibitor palbociclib (PD0332991) in Rb + advanced breast cancer: phase II activity, safety, and predictive biomarker assessment. Clin Cancer Res. 2015;21:995.

    Article  CAS  PubMed  Google Scholar 

  196. Leonard JP, LaCasce AS, Smith MR, Noy A, Chirieac LR, Rodig SJ, Yu JQ, Vallabhajosula S, Schoder H, English P, et al. Selective CDK4/6 inhibition with tumor responses by PD0332991 in patients with mantle cell lymphoma. Blood. 2012;119:4597.

    Article  CAS  PubMed  Google Scholar 

  197. Dickson MA, Tap WD, Keohan ML, D’Angelo SP, Gounder MM, Antonescu CR, Landa J, Qin L-X, Rathbone DD, Condy MM, et al. Phase II trial of the CDK4 inhibitor PD0332991 in patients with advanced CDK4-Amplified well-differentiated or dedifferentiated liposarcoma. J Clin Oncol. 2013;31:2024–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  198. Verbiest V, Montaudon D, Tautu MT, Moukarzel J, Portail J-P, Markovits J, Robert J, Ichas F, Pourquier P. Protein arginine (N)-methyl transferase 7 (PRMT7) as a potential target for the sensitization of tumor cells to camptothecins. FEBS Lett. 2008;582:1483–9.

    Article  CAS  PubMed  Google Scholar 

  199. Lee J-H, Cook JR, Yang Z-H, Mirochnitchenko O, Gunderson SI, Felix AM, Herth N, Hoffmann R, Pestka S. PRMT7, a new protein arginine methyltransferase that synthesizes symmetric dimethylarginine. J Biol Chem. 2005;280:3656–64.

    Article  CAS  PubMed  Google Scholar 

  200. Bedford MT, Richard S. Arginine methylation: an emerging regulator of protein function. Mol Cell. 2005;18:263–72.

    Article  CAS  PubMed  Google Scholar 

  201. Dasgupta A, Jung KJ, Jeong SJ, Brady JN. Inhibition of methyltransferases results in induction of G2/M checkpoint and programmed cell death in human T-lymphotropic virus type 1-transformed cells. J Virol. 2008;82:49–59.

    Article  CAS  PubMed  Google Scholar 

  202. Heit R, Rattner JB, Chan GK, Hendzel MJ. G2 histone methylation is required for the proper segregation of chromosomes. J Cell Sci. 2009;122:2957–68.

    Article  CAS  PubMed  Google Scholar 

  203. Schwerk C, Schulze-Osthoff K. Methyltransferase inhibition induces p53-dependent apoptosis and a novel form of cell death. Oncogene. 2005;24:7002–11.

    Article  CAS  PubMed  Google Scholar 

  204. Chuang CY, Chang CP, Lee YJ, Lin WL, Chang WW, Wu JS, Cheng YW, Lee H, Li C. PRMT1 expression is elevated in head and neck cancer and inhibition of protein arginine methylation by adenosine dialdehyde or PRMT1 knockdown downregulates proliferation and migration of oral cancer cells. Oncol Rep. 2017;38:1115–23.

    Article  PubMed  Google Scholar 

  205. Eram MS, Shen Y, Szewczyk M, Wu H, Senisterra G, Li F, Butler KV, Kaniskan HU, Speed BA, Dela Sena C, et al. A potent, selective, and cell-active inhibitor of human type i protein arginine methyltransferases. ACS Chem Biol. 2016;11:772–81.

    Article  CAS  PubMed  Google Scholar 

  206. Smil D, Eram MS, Li F, Kennedy S, Szewczyk MM, Brown PJ, Barsyte-Lovejoy D, Arrowsmith CH, Vedadi M, Schapira M. Discovery of a dual PRMT5–PRMT7 inhibitor. ACS Med Chem Lett. 2015;6:408–12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  207. Hu H, Qian K, Ho MC, Zheng YG. Small molecule inhibitors of protein arginine methyltransferases. Expert Opin Investig Drugs. 2016;25:335–58.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  208. Hu H, Owens EA, Su H, Yan L, Levitz A, Zhao X, Henary M, Zheng YG. Exploration of cyanine compounds as selective inhibitors of protein arginine methyltransferases: synthesis and biological evaluation. J Med Chem. 2015;58:1228–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  209. Dillon MB, Bachovchin DA, Brown SJ, Finn MG, Rosen H, Cravatt BF, Mowen KA. Novel inhibitors for PRMT1 discovered by high-throughput screening using activity-based fluorescence polarization. ACS Chem Biol. 2012;7:1198–204.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  210. Weerapana E, Wang C, Simon GM, Richter F, Khare S, Dillon MBD, Bachovchin DA, Mowen K, Baker D, Cravatt BF. Quantitative reactivity profiling predicts functional cysteines in proteomes. Nature. 2010;468:790–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  211. Plett KL, Raposo AE, Bullivant S, Anderson IC, Piller SC, Plett JM. Root morphogenic pathways in Eucalyptus grandis are modified by the activity of protein arginine methyltransferases. BMC Plant Biol. 2017;17:62.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  212. Chan-Penebre E, Kuplast KG, Majer CR, Boriack-Sjodin PA, Wigle TJ, Johnston LD, Rioux N, Munchhof MJ, Jin L, Jacques SL, et al. A selective inhibitor of PRMT5 with in vivo and in vitro potency in MCL models. Nat Chem Biol. 2015;11:432–7.

    Article  CAS  PubMed  Google Scholar 

  213. Castillo-Aguilera O, Depreux P, Halby L, Arimondo PB, Goossens L. DNA Methylation Targeting: the DNMT/HMT crosstalk challenge. Biomolecules. 2017;7:3.

    Article  PubMed Central  CAS  Google Scholar 

  214. Hamamoto R, Nakamura Y. Dysregulation of protein methyltransferases in human cancer: an emerging target class for anticancer therapy. Cancer Sci. 2016;107:377–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Authors’ contributions

AER and SCP contributed to writing the manuscript. Both authors read and approved the final manuscript.

Acknowledgements

The authors would like to acknowledge Dr. Liza Cubeddu for critically reviewing this manuscript.

Competing interests

The authors declare that they have no competing interests.

Availability of data and materials

Not applicable.

Consent for publication

Not applicable.

Ethics approval and consent to participate

Not applicable.

Funding

This work was supported by a Postgraduate Research Award from Western Sydney University (AER).

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sabine C. Piller.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Raposo, A.E., Piller, S.C. Protein arginine methylation: an emerging regulator of the cell cycle. Cell Div 13, 3 (2018). https://doi.org/10.1186/s13008-018-0036-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13008-018-0036-2

Keywords