Abstract

Breast cancer is the most commonly diagnosed female cancer in the world. Though therapeutic treatments are available to treat breast cancer and in some instances are successful, the occurrence of unsuccessful treatment, or the rate of tumour recurrence, still remains strikingly high. Therefore, novel therapeutic treatment targets need to be discovered and tested. The protein arginine methyltransferases (PRMTs) are a family of enzymes that catalyse arginine methylation and are implicated in a myriad of cellular pathways including transcription, DNA repair, RNA metabolism, signal transduction, protein–protein interactions and subcellular localisation. In breast cancer, the expression levels and enzymatic activity of a number of PRMTs is dysregulated; significantly altering the regulation of many cellular pathways that are implicated in breast cancer development and progression. Here, we review the current knowledge on PRMTs in breast cancer and provide a rationale for how PRMTs may provide novel therapeutic targets for the treatment of breast cancer.

Introduction

Worldwide, breast cancer is the most common cancer observed in women accounting for 25% of all cancers, with 1.7 million new cases diagnosed in 2012 (1). Like all cancers, breast cancer arises through the dysregulation of cellular pathways regulating sensitivity to growth stimuli, cellular proliferation, replicative potential, apoptosis, angiogenesis and tissue invasiveness and metastasis (2). Though therapeutic options are available for the treatment of breast cancer including surgical resection, radiation and chemotherapy, in many instances tumours become refractory to these treatments or the tumours may recur. Therefore, understanding the biological events that lead to the progression and therapeutic resistance of breast cancer is essential for the development of novel treatment options for this disease.

Arginine methylation, a common post-translation modification catalysed by a family of enzymes called the protein arginine methyltransferases (PRMTs) is implicated in a number of cellular processes including transcription, DNA repair, RNA metabolism, signal transduction, protein–protein interactions and subcellular localisation (3). PRMTs catalyse the transfer of a methyl group from S-adenosylmethionine (AdoMet) to the guanidine nitrogen of an arginine residue, resulting in the formation of a methylarginine and S-adenosylhomocysteine (AdoHcy) as a by-product (3). The addition of methyl groups to a substrate promotes increased bulkiness and local hydrophobicity which can affect protein–protein interactions, albeit without changing the cationic charge of the arginine residue (4).

Members of the PRMT family are characterised by four conserved sequence motifs (I, post-I, II and III), and ‘double E’ and ‘THW’ loops (5, 6). There are currently nine identified human PRMTs (PRMT 1–9) in mammals. PRMTs are classified based on the type of arginine methylation that they catalyse. Both type I and type II PRMTs produce monomethylarginines as an intermediate, with type I PRMTs (PRMT1, 2, 3, 4(CARM1), 6, 8) catalyzing the formation of asymmetric dimethylarginines (ADMA), while type II PRMTs (PRMT5) form symmetric dimethylarginines (SDMA). Type III PRMTs (PRMT7) are capable of forming stable monomethylarginine (MMA) products (7), while the enzymatic activity of PRMT9 has yet to be fully characterised (8) (Figure 1). Most PRMTs methylate arginine residues within a glycine- and arginine-rich (GAR) motif (9). However, CARM1and PRMT7 display unique substrate specificity, as CARM1 methylates arginine residues within a proline-, glycine- and methionine-rich (PGM) motif (10), and PRMT7 methylates arginine residues in a RXR motif (11). PRMT5 can methylate arginine residues in both a GAR and PGM motif (12, 13), while PRMT6 is efficient at methylating arginine residues both within the GAR motif, in addition to arginine residues independent of this sequence (14–17).

Figure 1.

Mammalian PRMTs. There are currently nine identified members of the PRMT family. Type I, II, and III PRMTs catalyse the formation of monomethylarginines (MMA). Type I and II PRMTs form MMA as an intermediate before the second methyl group is placed on the substrate. Type I PRMTs catalyse ADMA, while type II PRMTs form SDMA. Type I PRMTs include PRMTs 1, 2, 3, 4, 6 and 8. PRMT5 is a Type II PRMT and PRMT7 is a Type III PRMT. The enzymatic activity of PRMT9 has yet to be elucidated.

PRMTs and breast cancer

As PRMTs operate in a plethora of central cell regulatory pathways, it would be hypothesised that dysregulation of their expression might have adverse effects on cellular and tissue homeostasis. Indeed, a growing body of evidence has linked PRMTs to the development and progression of cancer. Aberrant expression of PRMTs has been characterised in human tumours including, lung, colorectal, bladder, prostate, leukaemia and breast (18–24). PRMTs 1–7 have been linked with breast cancer development and progression and this review will outline the current knowledge regarding the proposed role of these PRMTs in breast cancer and elaborate on their potential as novel therapeutic targets for breast cancer.

PRMT1

PRMT1 is the main arginine methyltransferase in cells, accounting for ~85% of all arginine methylation activity (25, 26). Complete depletion of PRMT1 expression results in a loss of cell viability and embryonic lethality in mice (27, 28). Functionally, PRMT1 has been implicated in transcription activation, signal transduction, RNA splicing and DNA repair (29–33). Alternative splicing of the 5′ end of the PRMT1 pre-mRNA yields seven distinct isoforms (v1–v7) with each isoform having a unique N-terminal sequence and a slightly different molecular weight. Each PRMT1 isoform has distinct enzymatic properties, substrate specificity and subcellular localisation, and thus it would be hypothesised that they should be functionally distinct (34).

In breast cancer tumour samples, PRMT1 mRNA expression is upregulated compared to adjacent normal tissue with PRMT1 expression correlating with patient age, menopausal status, tumour grade and status of the progesterone receptor (22, 35). Examination of the expression of PRMT1v1, v2 and v3 in breast cancer samples identified a strong correlation between PRMT1v1 mRNA expression and poor patient prognosis whereas PRMT1v2 and PRMT1v3 expression did not correlate with any clinical or pathological indicators at least at the RNA level (22). Immunohistochemical examination of total PRMT1 protein expression, not differentiating between specific isoforms, identified increased PRMT1 expression within the cytoplasm (22). Since PRMT1v2 localises predominantly to the cytoplasm, while PRMT1v1 displays nuclear localisation (34, 36), it is interesting to speculate that the observed increased cytoplasmic expression of PRMT1 may actually represent increased PRMT1v2 expression. In a panel of breast cancer cell lines, increased mRNA and protein expression of all PRMT1 isoforms was observed relative to normal mammary epithelial cells, though expression of PRMT1v2 was increased to an extent greater than the other isoforms. Additionally, increased PRMT1v1 and v2 mRNA expression was observed in breast tumour tissue compared to normal tissue (34). Interestingly, contradictory results were obtained pertaining to PRMT1v2 mRNA expression as Mathioudaki et al. (22) saw no increase in PRMT1v2 expression in breast tumour samples, while Goulet et al. (34) describe increased expression in breast cancer cell lines and in a tissue sample. Further assessment of PRMT1v2 mRNA expression in a greater number of breast tumour samples from different types and disease stages will be required to address this potential discrepancy.

PRMT1v2 is the only PRMT1 isoform containing exon 2 within its coding sequence. Exon 2 encodes a CRM1-dependent nuclear export sequence which distinguishes PRMT1v2 as the only PRMT1 isoform with predominantly cytoplasmic localisation. Examination of the contribution of PRMT1v2 to breast cancer indicated an involvement for PRMT1v2 in promoting cell survival and cellular invasion (37). Specific depletion of PRMT1v2, without affecting the expression of PRMT1v1, the predominant PRMT1 isoform resulted in apoptotic cell death. Furthermore, PRMT1v2 depletion caused a significant decrease in cell motility and invasion in highly aggressive MDA MB 231 breast cancer cells; a decrease significantly greater than what could be accounted for from the observed apoptosis. Conversely, overexpression of PRMT1v2 increased motility and invasion in mildly aggressive MCF7 cells; a phenotype dependent on PRMT1v2’s cytoplasmic localisation and methyltransferase activity. While overexpression of PRMT1v1 was also able to enhance cell motility in MCF7 cells, augmented invasion was a phenotype specific to PRMT1v2 overexpression. PRMT1v2 overexpression in MCF7 cells resulted in disruption of adherens junctions, with dissociation of F-actin, E-cadherin and β-catenin from these junctions (37). Disruption of adherens junction stability is associated with increased invasive potential and metastasis in numerous epithelial-derived cancers, including breast cancer (38–43). PRMT1v2-mediated cell invasion occurs through phosphorylation-dependent downregulation of β-catenin (37). β-catenin protein stability is regulated through phosphorylation of serine residues 33 and 37 and threonine residue 41 by a degradation complex composed of axin, adenomatous polyposis coli (APC), casein kinase 1 (CK1) and glycogen synthase kinase 3β (GSK3β), which negatively regulates WNT signalling through the degradation of β-catenin (44–46). Cha et al. have shown that PRMT1 functions to negatively regulate β-catenin protein expression through the methylation of axin. Both PRMT1v1 and PRMT1v2 were capable of methylating axin, in vitro. Axin methylation by PRMT1 results in increased axin protein stability; therefore enabling an enhanced interaction with GSK3β, thus promoting β-catenin degradation. Conversely, PRMT1 depletion results in increased cytoplasmic β-catenin and β-catenin-dependent transcriptional activity (47).

Additionally, PRMT1 influences signalling pathways aberrantly regulated in breast cancer pathology. Presence of the estrogen receptor (ER) in a breast tumour is an important prognostic and predictive biomarker (48). Upon rapid estrogen stimulation, PRMT1 methylates estrogen receptor α (ERα) on arginine residue 260, resulting in the assembly of a complex containing ERα, the p85 subunit of phosphoinositide 3-kinase (PI3K), Src and focal adhesion kinase (FAK). Assembly of this complex results in an activating phosphorylation on serine 473 on Akt, a downstream target of this pathway whose activation promotes cellular proliferation and anti-apoptotic signalling. Mutation of arginine residue 260 or PRMT1 depletion attenuates ERα-mediated activation of Akt. Moreover, in breast cancer patient samples, an increase in ERα methylation is observed, in comparison to adjacent normal mammary epithelium (49–51). Alterations in gene expression of several nodes of the PI3K/Akt signalling pathway are frequent in ER positive breast cancer (52). In accordance with results observed by Le Romancer et al. (49), Simoncini et al. (53) show that estrogen stimulation of ERα promotes receptor binding to the p85α subunit of PI3K resulting in downstream activation of Akt. Thus, it appears that the therapeutic targeting of PRMT1 may be beneficial in attenuating ER-mediated activation of the PI3K/Akt signalling pathway. Furthermore, methylated ERα is present exclusively in the cytoplasm (49); therefore, it could be hypothesised that PRMT1v2 is the isoform responsible for depositing this methyl mark as it is the only isoform which displays predominant cytoplasmic localisation (34, 36). Further experimentation will be required to determine if this is true.

Interestingly, a number of studies have illustrated conflicting roles for PRMT1 in regulating apoptosis. PRMT1 methylation of forkhead box O1 (FOXO1) at arginine residues 248 and 250 attenuates Akt phosphorylation of FOXO1 at serine 253. Inhibition of Akt phosphorylation facilitates nuclear localisation of FOXO1 which promotes oxidative-stress induced apoptosis (54). In accordance with the above study, in MCF7 cells, PRMT1 methylates the pro-apoptotic protein, BCL-2 antagonist of cell death (BAD) promoting apoptosis. PRMT1 methylation of BAD prevents its phosphorylation by Akt, which would result in BAD’s mitochondrial sequestration and binding to the anti-apoptotic protein BCL-XL, and thus inhibit apoptosis (55). However, in contrast to these two studies, an inhibitory methylation mark by PRMT1 on the apoptotic signal-regulating kinase 1 (ASK1) prevents its activation, protecting cells from stress induced apoptosis in MCF7 and MDA MB 231 cells (56). Although discrepancies have arisen regarding PRMT1’s function in regulating apoptosis, these discrepancies could potentially be due to PRMT1 isoform specific functions in apoptosis, as the above studies focused on the role of PRMT1 in general. Further experimentation will be required to delineate if besides PRMT1v2 other PRMT1 isoforms function in apoptotic or cell survival pathways, and if their functions are contradictory or redundant.

PRMT1 null mouse embryonic fibroblasts (MEFs) exhibit spontaneous DNA damage and aberrant cell cycle regulation leading to chromosomal irregularities that eventually lead to cell death (57). Therefore, it is not surprising that PRMT1 functions in pathways responsible for maintaining genomic stability through the regulation of specific target proteins. PRMT1 methylates breast cancer 1, early onset (BRCA1), a tumour suppressor protein involved in DNA damage pathways. Genetic mutation of the BRCA1 gene results in increased predisposition to the development of breast cancer (58). BRCA1 was shown to be methylated in both breast cancer cell lines and patient breast tumour samples. PRMT1 methylation of BRCA1 occurs within the DNA-binding region, mediating not only BRCA1’s ability to bind its target genes but also its ability to interact with other proteins (59).

PRMT1 regulates additional components of the DNA damage pathway thereby regulating their function. PRMT1 methylation of the DNA damage sensing protein p53-binding protein 1 (53BP1) regulates its recruitment to sites of DNA double-strand breaks (32). Furthermore, PRMT1 methylates MRE11, a component of the MRE11/RAD50/NBS1 (MRN) complex responsible for repairing DNA double-stranded breaks, mediating its cellular localisation and exonuclease activity (60, 61). Additionally, PRMT1 has also been implicated in maintaining telomere length. PRMT1 methylates TRF2, a component of the shelterin complex which functions to control telomere length and stability, with depletion of PRMT1 in cancer cells resulting in telomere shortening (62). Finally, upon ultraviolet irradiation, PRMT1 functions synergistically with CARM1 and the acetyltransferase p300 as transcriptional co-activators of p53, regulating its target gene expression (63). These studies exhibit that PRMT1 exudes an integral role in the maintenance of genomic stability.

The transforming growth factor β (TGFβ)/Smad signalling pathway has been implicated in breast cancer through stimulating the de-differentiation of epithelial cells to malignant fibroblastic cells capable of invasion and metastasis (64, 65). PRMT1 was recently identified as a component of the TGFβ signalling pathway in response to bone morphogenetic protein (BMP) binding to TGFβ receptors, RI and RII (66). Activation of the TGFβ receptor occurs through ligand binding and dimerization of the RI and RII receptors (67). Smad6 typically sequesters the TGFβ RI receptor holding it in a dormant state, with PRMT1 binding the TGFβ RII receptor. Upon stimulation of the TGFβ receptor though BMP ligand binding, PRMT1 methylates Smad6 resulting in dissociation of Smad6 from the RII receptor, promoting RI and RII dimerization and activation of the receptor. TGFβ receptor activation through BMP stimulation has been shown to promote cancer cell invasion (68).

Since PRMT1 accounts for the majority of the identified methyltransferase activity in the cell (25, 26), it is not surprising that abnormal regulation of PRMT1 has been associated with cancer development and progression in mammary epithelial cells. PRMT1 mRNA and protein expression are increased in breast cancer tumours (22, 35), and PRMT1 was shown to contribute to pathways promoting increased cellular proliferation, invasion and survival in breast cancer cell lines (35, 37, 49, 56, 66). The majority of the published data pertaining to PRMT1 has focused on the contribution of total PRMT1 to breast cancer, not differentiating between the contributions of specific isoforms. It was recently shown that PRMT1v2 promotes cell survival, migration and invasion while PRMT1v1 stimulates cell migration in breast cancer cells (37). Therefore, future work should focus on the contribution of the specific isoforms in breast cancer and elucidating each of their roles in breast cancer, as all were shown to be upregulated in breast cancer cell lines (34).

Similarly, future work should additionally focus on how PRMT1 expression and its enzymatic activity are regulated. PRMT1 interacts with the mammalian-early gene TIS21/BTG2/PC3, the leukaemia-associated B-cell translocation gene (BTG1) (69), and CCR4-associated factor 1 (hCAF1) (70), with both BTG2 (71) and hCAF1 (70) capable of modulating PRMT1 methyltransferase activity in a substrate-dependent manner. Furthermore, hCAF1 was found to interact with both BTG1 and BTG2 to mediate transcription of ERα target genes (72). BTG1 and BTG2 typically function to inhibit cell proliferation and both have been implicated in the development of breast cancer. Decreased BTG1 expression correlates with poor patient prognosis in breast cancer patient samples (73, 74), while BTG2 maps to chromosome 1q32, a region frequently altered in breast adenocarcinomas (75, 76). Therefore, it is plausible that under normal growth conditions BTG1, BTG2 and hCAF1 control PRMT1 enzymatic activity, and in breast cancer cells, loss of BTG1 and/or BTG2 removes this regulation on PRMT1 activity allowing PRMT1 to function promiscuously. Further investigation will be required to prove if this hypothesis is true. The published data suggests that dysregulation of PRMT1 contributes significantly to breast oncogenesis, and although further experimentation will elaborate on PRMT1’s role in breast cancer, investigating PRMT1 as a novel therapeutic target in breast cancer may prove beneficial.

PRMT2

PRMT2 was identified through its sequence homology (~50%) with the catalytic domain of PRMT1 (77). Unlike PRMT1 null mice which display embryonic lethality, PRMT2 null mice are viable; although PRMT2 null MEFs exhibit increased NF-κB transcriptional activity and are more resistant to apoptosis than wild-type MEFs (78,79). PRMT2 exhibits weak methyltransferase activity relative to PRMT1 (80). With the exception of arginine residue 8 on histone 3, identified substrates for PRMT2 are lacking (81). However, PRMT2 contains an N-terminal Src homology 3 (SH3) binding domain, which has identified functions in cell proliferation, migration and cytoskeletal modifications (82). Therefore, this does not exclude the possibility that additional yet unidentified PRMT2 substrates exist, or in some cellular processes, PRMT2 may function independent of its catalytic activity.

PRMT2 undergoes alternative splicing to produce five distinct isoforms: full length PRMT2 and four truncated isoforms PRMT2L2, PRMT2α, β, and γ that occur through alternative splicing in exons 7–10 of the mRNA. PRMT2L2, PRMT2α, β and γ have lost conserved motifs III and the THW loop which forms part of the AdoMet binding pocket, and thus may exhibit no methyltransferase activity (83, 84). Expression of each isoform (PRMT2, PRMT2L2, PRMT2α, β and γ) was investigated in breast cancer cell lines, and it was discovered that PRMT2 mRNA and protein expression of each isoform is increased in a panel of ER positive breast cancer cell lines in comparison to a panel of ER negative breast cancer cell lines. In breast tumour samples, PRMT2 mRNA expression of all of the isoforms is increased; consistent with breast cancer cell lines, PRMT2 expression in ER+ tumours is increased relative to ER− tumours (83, 84). Furthermore, PRMT2 protein expression, not differentiating between isoforms, was also elevated in breast tumours with a similar increase in expression observed in ER+ tumours relative to ER− tumours (84).

PRMT2 interacts with and enhances the activity of the estrogen and progesterone receptors (PR) in a ligand-dependent manner (85). Binding of the ERα receptor by all PRMT2 isoforms was shown to enhance estrogen-mediated transactivation of the receptor augmenting promoter activity of the transcription factor Snail (83, 84). Increased Snail expression in cancer cells is consistent with enhanced cellular invasion (86). In contrast to PRMT1 which methylates ERα (49), to date methylation of the ER by PRMT2 has not been documented. Intriguingly, in response to estrogen stimulation, PRMT2 negatively regulates cell proliferation and colony formation in ER+ breast cancer cells, while PRMT2 depletion increases tumour growth in a MCF7 xenograft mouse model (84, 87). PRMT2-mediated growth suppression in ER+ breast cancer cell lines was discovered to occur through modulation of the E2F/cyclin D1 pathways. Depletion of all PRMT2 splice variants coincides with an enhancement in E2F expression and activity (84). This observation is consistent with PRMT2 binding the retinoblastoma (Rb) protein causing E2F repression (78). PRMT2 regulates cyclin D1 by indirectly binding to the AP-1 site on the cyclin D1 promoter, and through suppression of the Akt/GSK-3β signalling pathways (87). PRMT2 displays predominantly nuclear localisation (83, 84). Loss of PRMT2 nuclear expression in breast tumour samples correlates with increased cyclin D1 expression, thus promoting breast tumour cell proliferation (87).

These results suggest that in contrast to PRMT1 which promotes breast tumour development and progression, PRMT2, at least in ER+ breast tumours suppresses cell proliferation, with loss of nuclear PRMT2 expression a potential driving force to tumour development. Interestingly, in ER+ breast cancer cell lines, PRMT2L2, a predominantly cytoplasmic variant of PRMT2 displays equal or increased mRNA expression compared to full length PRMT2 which localises predominantly to the nucleus (83). This suggests that abnormal alternative splicing of PRMT2 could at least in part be one mechanism that can facilitate ER+ breast tumour development. PRMT2 alternatively spliced variants PRMT2L2, PRMT2α, β and γ have lost conserved motifs III and the THW loop which form part of the AdoMet-binding motif, and thus should be enzymatically inactive (83, 84). Therefore, these variants may function to promote breast tumorigenesis through co-transactivation of the ER or other unknown transcription factors, or through other functions potentially mediated through PRMT2’s N-terminal SH3 binding domain.

PRMT3

PRMT3 shares 67% sequence similarity with PRMT1 in the catalytic methyltransferase domain. PRMT3 distinguishes itself from other members of the PRMT family as it contains a zinc-finger domain at its N-terminus, which is thought to function as a substrate recognition module (88). PRMT3 null mice are viable. Although mutant embryos are small, the mice attain normal size in adulthood (89). Though, to date a limited number of PRMT3 substrates have been identified, PRMT3 does methylate the ribosomal protein RPS2 to regulate ribosome homeostasis (89, 90), as well as p53 in an ARF and VHL30 dependent manner (91).

Though no studies have alluded to aberrant expression of PRMT3 in breast cancer, the tumour suppressor protein DAL-1/4.1B, whose expression is frequently lost in breast tumour samples (92, 93), binds to PRMT3, negatively regulating its methyltransferase activity (94). These results suggest that breast tumours may exhibit higher PRMT3 methyltransferase activity in tumours where loss of DAL-1/4.1B is observed. Further experimentation will be required to determine if PRMT3 expression and/or enzymatic activity is altered in breast tumours and if loss of DAL-1/4.1B binding to PRMT3 were to mediate any potential effects.

CARM1 (PRMT4)

PRMT4, best known as co-activator associated methyltransferase 1 (CARM1) exhibits a unique substrate specificity, where it methylates arginine residues within a proline-, glycine- and methionine-rich (PGM) motif instead of a GAR motif (10, 12). CARM1 null mice survive through gestation, though they are small and die shortly after birth; a phenotype mimicked in transgenic mice expressing catalytically inactive CARM1. CARM1 null embryos display differentiation defects in T cells, adipose tissue, chondrocytes, muscles and lungs (95–102). CARM1 functions in a myriad of cellular processes including transcription regulation (103), mRNA splicing (12, 104), cell cycle progression (105) and the DNA damage response (106, 107), through methylation of identified substrates including histones (108, 109), transcription factors (12, 110), co-activators (111–114), splicing factors (12, 109) and RNA polymerase II (115).

A number of studies have described a role for CARM1 in breast cancer aetiology and progression. CARM1 acts as a co-activator for ERα-mediated transcription in a ligand-independent manner. This association with ERα is dependent on phosphorylation by cAMP-associated protein kinase A at CARM1 serine residue 448 and CARM1 auto-methylation at arginine residue 551 (116, 117). In addition to a ligand-independent interaction with ERα, CARM1 is necessary for estrogen-stimulated proliferation of breast cancer cells which occurs through CARM1 methylation of H3R17, resulting in expression of the cell cycle regulator E2F1. Upon estrogen stimulation, CARM1 also regulates the protein stability and promotes the activity of the transcription factor AIB1, a protein frequently overexpressed in breast tumours (118). It has been observed that CARM1 functions synergistically with the proto-oncogene proline-, glutamic acid- and leucine-rich protein 1 (PELP1) to enhance ERα activity through PELP1 modulation of CARM1’s H3 arginine methylation at the promoters of ERα target genes (119). PELP1 promotes breast cancer invasion and metastasis, and functions as a co-activator of ERα. In breast tumours, PELP1 expression correlates with poor patient prognosis (120–124). Pharmacologic inhibition or depletion of CARM1 in ER+ breast cancer cells inhibits PELP1-mediated breast cancer cell proliferation, migration and anchorage-independent growth (119). Furthermore, CARM1 expression correlated with PELP1 expression in ER+ tumours, and was elevated in metastatic breast tumours compared to normal breast tissue (119). This data suggests that CARM1 functions synergistically with at least PELP1 to promote tumourigenesis in ER+ breast cancers.

Additional studies have further corroborated this detected increase in CARM1 expression in breast cancer samples. CARM1 expression was observed to be upregulated in breast tumours with CARM1 expression correlating with tumour size and grade, as well as Ki-67, TK1, CD71 and cyclin E expression (122, 123). Consistent with CARM1 expression correlating with cyclin E expression, CARM1 was shown to be recruited to the promoter of cyclin E1 where it functions as a transcriptional co-activator of cyclin E1 expression. This regulation coincided with a correlative increase in CARM1 and cyclin E1 expression in grade 3 breast tumours (105). Furthermore, CARM1 expression is increased in non-luminal breast tumours displaying HER2 expression (125, 126). Davis et al. (127) further elaborate on CARM1 expression in breast cancer through classification of CARM1 cellular localisation in correlation to distinct histo-pathological and molecular tumour subtypes, as well as expression of prognostic markers. The authors reasoned that since CARM1 functions as a transcription regulator in steroid pathways (118, 128), the majority of CARM1 in a breast cancer cells should be localised within the nucleus. However, an equal distribution of nuclear and cytoplasmic CARM1 was observed in ductal carcinoma in situ, and invasive and metastatic histological subtypes; although, in basal-like and HER2+ breast tumours a significant increase in cytoplasmic CARM1 localisation occurred (127). Interestingly, no correlation was detected between both nuclear or cytoplasmic CARM1 and the ER, as CARM1 was previously shown to interact and stimulate the activity of the ER in a ligand-dependent and independent manner (116, 118). No correlation was discerned between the PR and CARM1, although CARM1 cytoplasmic localisation did correlate with expression of both the HER2 and EGFR receptors. Lastly, cytoplasmic CARM1 expression was found to correlate with poor patient prognosis (127). Altogether, these results suggest that in addition to functioning as a co-activator of ERα, CARM1 may perform additional functions in promoting tumorigenesis in HER2+ and basal-like tumours that require CARM1 cytoplasmic expression.

In contrast to these studies, Al-Dhaheri et al. show CARM1 overexpression in ER+ breast cancer cells inhibiting estrogen-stimulated cell proliferation through repression of estrogen-activated target genes. Depletion of CARM1 in MCF7 cells resulted in an increase in tumour growth in a xenograft mouse model. CARM1 overexpression and the diminution in estrogen-stimulated cell proliferation coincided with an increase in expression of the cyclin-dependent kinase inhibitors p21 and p27, and a change in cell morphology towards a more differentiated phenotype. Lastly, CARM1 protein expression correlated with ERα expression with higher CARM1 expression detected in lower grade ER+ breast tumours (129).

CARM1 also methylates arginine residue 1064 of BAF155, the core subunit of the SWI/SNF chromatin remodeling complex. Methylated BAF155 expression was increased in breast patient samples, with methylated BAF155 promoting colony formation and cell migration, in vitro, and metastasis in an in vivo lung metastatic mouse model (130).

Though contradictory data has arisen for the role of CARM1 in breast cancer, the majority of data suggests that CARM1 functions to promote breast cancer development and progression. CARM1 is required for the ligand dependent and independent activation of ERα promoting transcription of ERα target genes (116, 118); however whether CARM1 expression is changed in ER+ breast tumours will require further investigation. CARM1 cellular localisation was previously observed to be cell type dependent (131). This could explain why increased cytoplasmic CARM1 expression was detected in only HER2+ and basal-like breast tumours (127) and not in other subtypes of breast cancer due to dysregulation of CARM1 expression in these particular tumours. CARM1 appears to promote breast cancer regardless of the specific subtype; although the molecular mechanisms regulated by CARM1 in the development and/or progression of breast cancer have yet to be completely elucidated. Nevertheless, the published data suggests that CARM1 may represent a novel therapeutic target.

PRMT5

PRMT5 is, thus far, the only characterised symmetric methyltransferase. Complete loss of PRMT5 results in loss of cell viability and mouse embryonic lethality, with PRMT5 required for embryonic epiblast cell differentiation (132). PRMT5 participates in a wide variety of cellular processes including transcription repression (133), alternative splicing (134–136), signal transduction (137–140), ribosome biogenesis (141), assembly of the Golgi apparatus (142), cellular differentiation (102, 143, 144), and germ cell specification (145, 146).

In breast cancer cells, PRMT5 methylates programmed cell death 4 (PDCD4) on arginine residue 110 (147). PDCD4 has been characterised as a tumour suppressor protein with lower expression observed in invasive breast carcinoma compared to normal mammary epithelium (148). In addition, PDCD4 expression is capable of suppressing anchorage-independent cell growth (149). Surprisingly, xenografts in severe combined immunodeficiency (SCID) mice of MCF7 cells co-overexpressing PDCD4 and PRMT5 display enhanced tumour growth which is dependent on PRMT5 methyltransferase activity. PRMT5 expression in breast cancer patient samples was observed to vary between tumour samples, though high expression of both PDCD4 and PRMT5 correlates with poor patient outcome (147).

Like PRMT3, DAL-1/4.1B can regulate PRMT5 methyltransferase activity (150). However, in contrast to PRMT3 where repression of PRMT3 methyltransferase activity was observed, DAL-1/4.1B modulates PRMT5 methyltransferase activity in a substrate-dependent manner. Interestingly, Jiang et al. (151) show DAL-1/4.1B co-operates with protein methylation to induce caspase 8-dependent apoptosis in MCF7 cells. Though, in this study, methylation was inhibited with adenosine-2′-3′-dialdehyde (Adox), a general methylation inhibitor, it is intriguing to speculate based on DAL-1/4.1B modulation of PRMT3 and PRMT5 activity that DAL-1/4.1B collaborates with one or both of these arginine methyltransferases to induce apoptosis. Furthermore, in breast cancer, could DAL-1/4.1B loss coincide with altered PRMT3 and/or PRMT5 enzymatic activity possibly correlating with a propensity for these enzymes to promote oncogenesis. Subsequently, does DAL-1/4.1B influence PRMT5 enzymatic activity towards PDCD4? Further investigations will be required to discover if this is the case.

PRMT6

PRMT6, a predominantly nuclear protein was identified as the first PRMT possessing the capacity for auto-methylation (14). Subsequently, it has been shown that this auto-methylation which occurs on arginine residue 35 is necessary to maintain PRMT6 protein stability (152). PRMT6 shares a high sequence similarity with the other PRMTs within the main AdoMet-catalytic domain. Similar to PRMT1, PRMT6 contains only the core PRMT catalytic domain with no additional domains present within its sequence (14). PRMT6 null mice are viable, although PRMT6 null MEFs undergo rapid senescence (153). PRMT6 displays distinct substrate specificity from other PRMTs and primarily functions as a transcriptional regulator through methylation of H3R2 (153–157), H3R42 (158) and H2AR29 (159).

Contradictory evidence has arisen pertaining to PRMT6 expression levels in breast cancer. Yoshimatsu et al. (35) identified increased PRMT6 expression in about 33% of breast cancer patient samples with increased PRMT6 expression effectuating poor patient prognosis. In support of this study, Phalke et al. (157) show increased PRMT6 expression in both breast cancer cell lines and patient tumour samples. However, Dowhan et al. (160) demonstrate that PRMT6 expression was significantly down-regulated in invasive ductal carcinomas. Therefore, further research will be required to clarify the precise impact PRMT6 expression exacts on breast tumours with a more thorough examination of PRMT6 expression in different histo-pathological and molecular subsets of breast cancer providing a more concise view as to its contribution.

PRMT6 functions in numerous cellular pathways including transcription regulation, alternative splicing, DNA repair and cell cycle regulation, all of which have been linked to breast cancer aetiology. PRMT6 transcriptionally represses the anti-angiogenic factor thrombospondin 1 (TSP-1) concurring with an increase in cell migration and invasion (155); however, overexpression of PRMT6 was also shown to induce TSP-1 expression, coinciding with a decrease in cellular migration and invasion due to downregulation of MMP2 and MMP9 (161). Furthermore, PRMT6 methylation of H3R2 promotes the transcriptional repression of HoxA10 (154), a protein whose upregulation in breast cancer promotes increased p53 expression and reduced invasive potential (162). Interestingly, PRMT6 transcriptionally represses the expression of TERT, Golph3 and prothymosin α through H3R2 methylation, all genes implicated in promoting breast cancer (154). Additionally, H2AR29 methylation represses MMP9 transcription (159); a matrix metalloproteinase involved in breast cancer pathology (163–168). Contrary to its function as a transcriptional repressor, PRMT6 can also function as a transcription activator by mediating the transcriptional activity of the PR, as well as working synergistically with CARM1 to promote expression of the ER (169).

PRMT6 methylates the high mobility group A1a protein (HMGA1a) within its second AT hook domain on arginine residues 57 and 59; a region critical for protein-DNA and protein–protein interactions (15, 170, 171). HMGA1a belongs to the high mobility group family of proteins which function as master switches of the chromatin structure through DNA binding, and promoting or repressing the formation of transcription factor complexes on the promoter regions of target genes (172, 173). HMGA1a expression is increased in both breast cancer cell lines and patient tumour samples. Elevated expression positively correlates with histological grade and HER2, and negatively with BRCA2 expression (174–177). Furthermore, HMGA1a overexpression in normal mammary epithelial cells promotes oncogenic transformation (174, 178).

Like CARM1, PRMT6 associates with PELP1, modulating PRMT6 enrichment of H3R2 methylation as well as PRMT6-mediated alternative splicing (160). Previously, PRMT6 was shown to mediate the alternative splicing of two genes whose expression is altered in breast cancer: vascular endothelial growth factor (VEGF) and spleen tyrosine kinase (Syk) (169). PRMT6 promotes the alternative splicing of VEGF isoform 165, the VEGF splice variant with the highest expression in breast tumour samples (179), while PRMT6 depletion inhibits splicing of a shorter form of Syk which displays increased expression in breast tumour samples compared to normal mammary epithelial tissue (180). Therefore, it is likely that PRMT6 regulation of alternative splicing, potentially through PELP1 modulation can stimulate that splicing of variants that may promote or inhibit breast cancer development.

PRMT6 methylates DNA polymerase β (Polβ), an enzyme involved in DNA base excision repair. PRMT6 methylation of Pol β arginine residues 83 and 152 is necessary to stimulate Pol β’s DNA binding capacity and facilitate its processivity (16). Low Polβ mRNA and protein expression correlates with breast cancer incidence (181, 182). This decrease in Polβ expression is associated with high tumour grade, positive lymph node status, increased ER+ tumour aggressiveness and poor patient survival (182). This suggests that PRMT6 methylation of Polβ serves as a mechanism to promote genomic stability, thus potentially inhibiting breast cancer development.

Dysregulation of the cell cycle through altered expression of critical regulatory mechanisms is often associated with oncogenesis. Inhibition of the cyclin-dependent kinase inhibitors (CDKI) p16, p21 and p27 has been shown to contribute to the development and progression of breast cancer (183–187). PRMT6 regulates the expression of these CDKIs transcriptionally or through methylation. PRMT6 depletion leads to cell cycle arrest resulting in increased expression of p16, p21 and p27 (153, 157, 188, 189), with PRMT6 methylation of H3R2 leading to transcriptional regulation of both p21 and p27 (157, 189). Furthermore, PRMT6 knockdown causes senescence in MEFs and breast cancer cells through increased p21 expression, albeit whether this occurs in a p53-dependent or independent manner remains to be determined (153, 157). Contrary to these results, Wang et al. (190) demonstrate that PRMT6 methylation of p16 impedes the propensity of the p16-CDK4 interaction that promotes cell proliferation. As discrepancies have arisen as to the precise role PRMT6 facilitates in cell cycle progression, further experimentation will be required to delineate and determine its function in the cell cycle.

Published data suggests that PRMT6 plays a significant role in breast cancer development; however, whether PRMT6 functions as a promoter or inhibitor of breast cancer remains to be determined. Discrepancies in PRMT6 expression may arise due to altered PRMT6 regulation in different breast cancer subtypes. Therefore, PRMT6 could potentially serve as a therapeutic marker in a subset of breast cancers, although further examination is required to determine which breast cancers display increased PRMT6 expression and how increased PRMT6 expression affects specific cellular pathways.

PRMT7

PRMT7 differs from the other PRMTs as it contains both N- and C-terminal putative AdoMet-binding domains. The presence of both AdoMet-binding domains is required for PRMT7 enzymatic activity; however, only the N-terminal domain binds AdoMet (191, 192). PRMT7 was initially characterised as a Type II PRMT (193), although recent data suggests that PRMT7 catalyses the formation of stable MMA; being the only PRMT proficient in catalyzing this reaction (7, 11). Though little functional data about PRMT7 is available, PRMT7 is capable of auto-methylation, and methylates substrates in a RXR motif with histone H2B an identified substrate (11). Previous characterisation of H2AR3 and H4R3 as substrates for PRMT7 (194) was likely due to contamination of the PRMT7 preparation with PRMT5 as PRMT7 is unable to methylate these arginines residues (11).

In several genome-wide studies in human patients, a strong correlation between over-expression of PRMT7 and breast cancer aggressiveness and metastasis exists. Specifically, expression of chromosome 16q22 (which contains the PRMT7 gene) is constantly upregulated in metastatic breast cancer (195). Moreover, PRMT7 interacts with CTCFL, a known oncogene which increases the activity of PRMT7 that is highly expressed in breast cancer tumours (196, 197). These studies suggest that PRMT7 may contribute to breast cancer pathogenesis, though specific experimentation still needs to be performed to identify a specific contribution of PRMT7 to breast cancer.

Small molecular inhibitors of PRMTs

Numerous groups have attempted to develop PRMT inhibitors for potential therapeutic treatment in cancer (reviewed in references 198, 199). In 2004, nine compounds named arginine methyltransferase inhibitors (AMIs) were the first specific inhibitors of PRMTs discovered. AMI-1 was shown to inhibit arginine but not lysine methyltransferase activity in vitro, prevent in vivo methylation of cellular proteins, and modulate oestrogen and androgen-mediated transcription (200). Several groups have used AMIs as analogs to synthesise additional inhibitors; however, issues pertaining to selectivity, potency, and cellular permeability have made these inhibitors less than ideal for in vivo analysis (201–204).

Pertaining to cancer, Spannhoff et al. (205) identified two compounds, allantodapsone and stilbamidine which specifically inhibit PRMT1. Both compounds were shown to inhibit PRMT1 methylation of H4R3 in a dose-dependent manner while having minimal inhibitory effects on lysine methylation of H3K4. Furthermore, both compounds could constrain estradiol-mediated activation of the ER. Bis-chloroacetyl compound 2e, a derivative of allantodapsone that specifically inhibits PRMT1 is capable of impeding proliferation in MCF7 breast cancer and LNCaP prostate cancer cells. Moreover, this compound was capable of inhibiting androgen-mediated transcription (206). Additionally, Wang et al. (207) describe inhibition of LNCaP cell growth by compound 9, a PRMT1 specific inhibitor. Lastly, Yan et al. (208) demonstrate that 2,5-bis(4-amidinophenyl)furan, a competitive inhibitor for PRMT1’s AdoMet binding site is cell permeable and can inhibit intracellular PRMT1 activity, resulting in a decrease in cell proliferation in leukaemia cells. In addition to PRMT1, specific inhibitors of CARM1 have also been identified. Cheng et al. (209) show that xenoestrogens licochalcone A, kepone, benzyl 4-hydroxybenzoate, and tamoxifen can inhibit CARM1 enzymatic activity.

Recent advances in drug development have identified compounds that can inhibit the enzymatic activity of specific PRMTs, while impeding cancer cell growth and hormone-mediated transcription. Further experimentation will be needed to determine the efficacy of these compounds in the context of breast cancer. Although the majority of compounds have focused on inhibition of PRMT1, development of additional high potency PRMT1 inhibitors, as well as inhibitors targeting other PRMTs will be needed and their efficacy in breast cancer will need to be examined.

Conclusion

The importance of PRMTs in the aetiology and progression of breast cancer is now only beginning to be examined. Altered expression and enzymatic activity of the PRMTs impact crucial cellular pathways including growth stimuli, cellular proliferation, replicative potential, apoptosis, angiogenesis and tissue invasiveness and metastasis whose differential regulation causes breast cancer development and progression (Figure 2). Thus, further experimentation should identify and elucidate the specific impact each PRMT imparts mechanistically on distinct subsets of breast tumours. In contrast to the other PRMTs, PRMT2 appears to inhibit breast tumour growth, at least in ER+ breast tumours. While limited data pertaining to PRMT3, PRMT5 and PRMT7 in breast cancer suggests that they do function in breast cell oncogenesis, further experimentation is required to determine their precise role. PRMT1, CARM1 and PRMT6 appear to promote tumorigenesis in breast cancer cells; although PRMT6 may function in only a distinct subset of tumour, these three PRMTs may serve as novel therapeutic targets in the treatment of breast cancer.

Figure 2.

PRMT involvement in cellular pathways linked to oncogenesis. PRMTs are involved in cellular pathways linked to growth stimuli, cellular proliferation, replicative potential, apoptosis, angiogenesis and tissue invasiveness and metastasis whose differential regulation causes breast cancer development and progression.

Funding

This work was supported by grants from the Cancer Research Society (JC #16135) and the Canadian Breast Cancer Foundation (JC #2013-16). Baldwin RM is a Postdoctoral Fellow of the Canadian Institute of Health Research.

Conflict of interest statement: None declared.

References

1.

Ferlay
J.
Soerjomataram
I.
Ervik
M
et al.  .
(2013)
Cancer Incidence and Mortality Worldwide: IARC CancerBase No. 11
.

2.

Hanahan
D.
Weinberg
R. A
.
(2000)
The hallmarks of cancer
.
Cell
,
100
,
57
70
.

3.

Bedford
M. T.
Richard
S
.
(2005)
Arginine methylation an emerging regulator of protein function
.
Mol. Cell
,
18
,
263
272
.

4.

Tripsianes
K.
Madl
T.
Machyna
M.
Fessas
D.
Englbrecht
C.
Fischer
U.
Neugebauer
K. M.
Sattler
M
.
(2011)
Structural basis for dimethylarginine recognition by the Tudor domains of human SMN and SPF30 proteins
.
Nat. Struct. Mol. Biol.
,
18
,
1414
1420
.

5.

Katz
J. E.
Dlakić
M.
Clarke
S
.
(2003)
Automated identification of putative methyltransferases from genomic open reading frames
.
Mol. Cell. Proteomics
,
2
,
525
540
.

6.

Cheng
X.
Collins
R. E.
Zhang
X
.
(2005)
Structural and sequence motifs of protein (histone) methylation enzymes
.
Annu. Rev. Biophys. Biomol. Struct.
,
34
,
267
294
.

7.

Zurita-Lopez
C. I.
Sandberg
T.
Kelly
R.
Clarke
S. G
.
(2012)
Human protein arginine methyltransferase 7 (PRMT7) is a type III enzyme forming ω-NG-monomethylated arginine residues
.
J. Biol. Chem.
,
287
,
7859
7870
.

8.

Lee
J.
Sayegh
J.
Daniel
J.
Clarke
S.
Bedford
M. T
.
(2005)
PRMT8, a new membrane-bound tissue-specific member of the protein arginine methyltransferase family
.
J. Biol. Chem.
,
280
,
32890
32896
.

9.

Boffa
L. C.
Karn
J.
Vidali
G.
Allfrey
V. G
.
(1977)
Distribution of NG, NG,-dimethylarginine in nuclear protein fractions
.
Biochem. Biophys. Res. Commun.
,
74
,
969
976
.

10.

Lee
J.
Bedford
M. T
.
(2002)
PABP1 identified as an arginine methyltransferase substrate using high-density protein arrays
.
EMBO Rep.
,
3
,
268
273
.

11.

Feng
Y.
Maity
R.
Whitelegge
J. P.
et al. 
(2013)
Mammalian protein arginine methyltransferase 7 (PRMT7) specifically targets RXR sites in lysine- and arginine-rich regions
.
J. Biol. Chem.
,
288
,
37010
37025
.

12.

Cheng
D.
Côté
J.
Shaaban
S.
Bedford
M. T
.
(2007)
The arginine methyltransferase CARM1 regulates the coupling of transcription and mRNA processing
.
Mol. Cell
,
25
,
71
83
.

13.

Branscombe
T. L.
Frankel
A.
Lee
J. H.
Cook
J. R.
Yang
Z.
Pestka
S.
Clarke
S
.
(2001)
PRMT5 (Janus kinase-binding protein 1) catalyzes the formation of symmetric dimethylarginine residues in proteins
.
J. Biol. Chem.
,
276
,
32971
32976
.

14.

Frankel
A.
Yadav
N.
Lee
J.
Branscombe
T. L.
Clarke
S.
Bedford
M. T
.
(2002)
The novel human protein arginine N-methyltransferase PRMT6 is a nuclear enzyme displaying unique substrate specificity
.
J. Biol. Chem.
,
277
,
3537
3543
.

15.

Sgarra
R.
Lee
J.
Tessari
M. A.
Altamura
S.
Spolaore
B.
Giancotti
V.
Bedford
M. T.
Manfioletti
G
.
(2006)
The AT-hook of the chromatin architectural transcription factor high mobility group A1a is arginine-methylated by protein arginine methyltransferase 6
.
J. Biol. Chem.
,
281
,
3764
3772
.

16.

El-Andaloussi
N.
Valovka
T.
Toueille
M.
et al. 
(2006)
Arginine methylation regulates DNA polymerase beta
.
Mol. Cell
,
22
,
51
62
.

17.

Boulanger
M. C.
Liang
C.
Russell
R. S.
Lin
R.
Bedford
M. T.
Wainberg
M. A.
Richard
S
.
(2005)
Methylation of Tat by PRMT6 regulates human immunodeficiency virus type 1 gene expression
.
J. Virol.
,
79
,
124
131
.

18.

Hong
H.
Kao
C.
Jeng
M. H.
et al. 
(2004)
Aberrant expression of CARM1, a transcriptional coactivator of androgen receptor, in the development of prostate carcinoma and androgen-independent status
.
Cancer
,
101
,
83
89
.

19.

Cheung
N.
Chan
L. C.
Thompson
A.
Cleary
M. L.
So
C. W
.
(2007)
Protein arginine-methyltransferase-dependent oncogenesis
.
Nat. Cell Biol.
,
9
,
1208
1215
.

20.

Kim
Y. R.
Lee
B. K.
Park
R. Y.
Nguyen
N. T.
Bae
J. A.
Kwon
D. D.
Jung
C
.
(2010)
Differential CARM1 expression in prostate and colorectal cancers
.
BMC Cancer
,
10
,
197
.

21.

Mathioudaki
K.
Papadokostopoulou
A.
Scorilas
A.
Xynopoulos
D.
Agnanti
N.
Talieri
M
.
(2008)
The PRMT1 gene expression pattern in colon cancer
.
Br. J. Cancer
,
99
,
2094
2099
.

22.

Mathioudaki
K.
Scorilas
A.
Ardavanis
A.
Lymberi
P.
Tsiambas
E.
Devetzi
M.
Apostolaki
A.
Talieri
M
.
(2011)
Clinical evaluation of PRMT1 gene expression in breast cancer
.
Tumour Biol.
,
32
,
575
582
.

23.

Papadokostopoulou
A.
Mathioudaki
K.
Scorilas
A.
Xynopoulos
D.
Ardavanis
A.
Kouroumalis
E.
Talieri
M
.
(2009)
Colon cancer and protein arginine methyltransferase 1 gene expression
.
Anticancer Res.
,
29
,
1361
1366
.

24.

Pal
S.
Baiocchi
R. A.
Byrd
J. C.
Grever
M. R.
Jacob
S. T.
Sif
S
.
(2007)
Low levels of miR-92b/96 induce PRMT5 translation and H3R8/H4R3 methylation in mantle cell lymphoma
.
EMBO J.
,
26
,
3558
3569
.

25.

Pahlich
S.
Zakaryan
R. P.
Gehring
H
.
(2006)
Protein arginine methylation: Cellular functions and methods of analysis
.
Biochim. Biophys. Acta
,
1764
,
1890
1903
.

26.

Tang
J.
Frankel
A.
Cook
R. J.
Kim
S.
Paik
W. K.
Williams
K. R.
Clarke
S.
Herschman
H. R
.
(2000)
PRMT1 is the predominant type I protein arginine methyltransferase in mammalian cells
.
J. Biol. Chem.
,
275
,
7723
7730
.

27.

Pawlak
M. R.
Scherer
C. A.
Chen
J.
Roshon
M. J.
Ruley
H. E
.
(2000)
Arginine N-methyltransferase 1 is required for early postimplantation mouse development, but cells deficient in the enzyme are viable
.
Mol. Cell. Biol.
,
20
,
4859
4869
.

28.

Nicholson
T. B.
Chen
T.
Richard
S
.
(2009)
The physiological and pathophysiological role of PRMT1-mediated protein arginine methylation
.
Pharmacol. Res.
,
60
,
466
474
.

29.

Strahl
B. D.
Briggs
S. D.
Brame
C. J.
et al. 
(2001)
Methylation of histone H4 at arginine 3 occurs in vivo and is mediated by the nuclear receptor coactivator PRMT1
.
Curr. Biol.
,
11
,
996
1000
.

30.

Wang
H.
Huang
Z. Q.
Xia
L.
et al. 
(2001)
Methylation of histone H4 at arginine 3 facilitating transcriptional activation by nuclear hormone receptor
.
Science
,
293
,
853
857
.

31.

Yu
Z.
Vogel
G.
Coulombe
Y.
Dubeau
D.
Spehalski
E.
Hébert
J.
Ferguson
D. O.
Masson
J. Y.
Richard
S
.
(2012)
The MRE11 GAR motif regulates DNA double-strand break processing and ATR activation
.
Cell Res.
,
22
,
305
320
.

32.

Boisvert
F. M.
Rhie
A.
Richard
S.
Doherty
A. J
.
(2005)
The GAR motif of 53BP1 is arginine methylated by PRMT1 and is necessary for 53BP1 DNA binding activity
.
Cell Cycle
,
4
,
1834
1841
.

33.

Bedford
M. T.
Frankel
A.
Yaffe
M. B.
Clarke
S.
Leder
P.
Richard
S
.
(2000)
Arginine methylation inhibits the binding of proline-rich ligands to Src homology 3, but not WW, domains
.
J. Biol. Chem.
,
275
,
16030
16036
.

34.

Goulet
I.
Gauvin
G.
Boisvenue
S.
Côté
J
.
(2007)
Alternative splicing yields protein arginine methyltransferase 1 isoforms with distinct activity, substrate specificity, and subcellular localization
.
J. Biol. Chem.
,
282
,
33009
33021
.

35.

Yoshimatsu
M.
Toyokawa
G.
Hayami
S.
et al. 
(2011)
Dysregulation of PRMT1 and PRMT6, Type I arginine methyltransferases, is involved in various types of human cancers
.
Int. J. Cancer
,
128
,
562
573
.

36.

Herrmann
F.
Fackelmayer
F. O
.
(2009)
Nucleo-cytoplasmic shuttling of protein arginine methyltransferase 1 (PRMT1) requires enzymatic activity
.
Genes Cells
,
14
,
309
317
.

37.

Baldwin
R. M.
Morettin
A.
Paris
G.
Goulet
I.
Côté
J
.
(2012)
Alternatively spliced protein arginine methyltransferase 1 isoform PRMT1v2 promotes the survival and invasiveness of breast cancer cells
.
Cell Cycle
,
11
,
4597
4612
.

38.

Shtutman
M.
Levina
E.
Ohouo
P.
Baig
M.
Roninson
I. B
.
(2006)
Cell adhesion molecule L1 disrupts E-cadherin-containing adherens junctions and increases scattering and motility of MCF7 breast carcinoma cells
.
Cancer Res.
,
66
,
11370
11380
.

39.

Tomaskovic-Crook
E.
Thompson
E. W.
Thiery
J. P
.
(2009)
Epithelial to mesenchymal transition and breast cancer
.
Breast Cancer Res.
,
11
,
213
.

40.

Hiscox
S.
Jiang
W. G.
Obermeier
K.
Taylor
K.
Morgan
L.
Burmi
R.
Barrow
D.
Nicholson
R. I
.
(2006)
Tamoxifen resistance in MCF7 cells promotes EMT-like behaviour and involves modulation of beta-catenin phosphorylation
.
Int. J. Cancer
,
118
,
290
301
.

41.

Van Aken
E.
De Wever
O.
Correia da Rocha
A. S.
Mareel
M
.
(2001)
Defective E-cadherin/catenin complexes in human cancer
.
Virchows Arch.
,
439
,
725
751
.

42.

Morrogh
M.
Andrade
V. P.
Giri
D.
et al. 
(2012)
Cadherin-catenin complex dissociation in lobular neoplasia of the breast
.
Breast Cancer Res. Treat.
,
132
,
641
652
.

43.

Birchmeier
W.
Behrens
J
.
(1994)
Cadherin expression in carcinomas: role in the formation of cell junctions and the prevention of invasiveness
.
Biochim. Biophys. Acta
,
1198
,
11
26
.

44.

Hart
M.
Concordet
J. P.
Lassot
I.
et al. 
(1999)
The F-box protein beta-TrCP associates with phosphorylated beta-catenin and regulates its activity in the cell
.
Curr. Biol.
,
9
,
207
210
.

45.

Latres
E.
Chiaur
D. S.
Pagano
M
.
(1999)
The human F box protein beta-Trcp associates with the Cul1/Skp1 complex and regulates the stability of beta-catenin
.
Oncogene
,
18
,
849
854
.

46.

Winston
J. T.
Strack
P.
Beer-Romero
P.
Chu
C. Y.
Elledge
S. J.
Harper
J. W
.
(1999)
The SCFbeta-TRCP-ubiquitin ligase complex associates specifically with phosphorylated destruction motifs in IkappaBalpha and beta-catenin and stimulates IkappaBalpha ubiquitination in vitro
.
Genes Dev.
,
13
,
270
283
.

47.

Cha
B.
Kim
W.
Kim
Y. K.
et al. 
(2011)
Methylation by protein arginine methyltransferase 1 increases stability of Axin, a negative regulator of Wnt signaling
.
Oncogene
,
30
,
2379
2389
.

48.

Harvey
J. M.
Clark
G. M.
Osborne
C. K.
Allred
D. C
.
(1999)
Estrogen receptor status by immunohistochemistry is superior to the ligand-binding assay for predicting response to adjuvant endocrine therapy in breast cancer
.
J. Clin. Oncol.
,
17
,
1474
1481
.

49.

Le Romancer
M.
Treilleux
I.
Leconte
N.
Robin-Lespinasse
Y.
Sentis
S.
Bouchekioua-Bouzaghou
K.
Goddard
S.
Gobert-Gosse
S.
Corbo
L
.
(2008)
Regulation of estrogen rapid signaling through arginine methylation by PRMT1
.
Mol. Cell
,
31
,
212
221
.

50.

Le Romancer
M.
Treilleux
I.
Bouchekioua-Bouzaghou
K.
Sentis
S.
Corbo
L
.
(2010)
Methylation, a key step for nongenomic estrogen signaling in breast tumors
.
Steroids
,
75
,
560
564
.

51.

Poulard
C.
Treilleux
I.
Lavergne
E.
Bouchekioua-Bouzaghou
K.
Goddard-Léon
S.
Chabaud
S.
Trédan
O.
Corbo
L.
Le Romancer
M
.
(2012)
Activation of rapid oestrogen signalling in aggressive human breast cancers
.
EMBO Mol. Med.
,
4
,
1200
1213
.

52.

Ciruelos Gil
E. M
.
(2014)
Targeting the PI3K/AKT/mTOR pathway in estrogen receptor-positive breast cancer
.
Cancer Treat. Rev.
,
40
,
862
871
.

53.

Simoncini
T.
Hafezi-Moghadam
A.
Brazil
D. P.
Ley
K.
Chin
W. W.
Liao
J. K
.
(2000)
Interaction of oestrogen receptor with the regulatory subunit of phosphatidylinositol-3-OH kinase
.
Nature
,
407
,
538
541
.

54.

Yamagata
K.
Daitoku
H.
Takahashi
Y.
Namiki
K.
Hisatake
K.
Kako
K.
Mukai
H.
Kasuya
Y.
Fukamizu
A
.
(2008)
Arginine methylation of FOXO transcription factors inhibits their phosphorylation by Akt
.
Mol. Cell
,
32
,
221
231
.

55.

Sakamaki
J.
Daitoku
H.
Ueno
K.
Hagiwara
A.
Yamagata
K.
Fukamizu
A
.
(2011)
Arginine methylation of BCL-2 antagonist of cell death (BAD) counteracts its phosphorylation and inactivation by Akt
.
Proc. Natl. Acad. Sci. U. S. A.
,
108
,
6085
6090
.

56.

Cho
J. H.
Lee
M. K.
Yoon
K. W.
Lee
J.
Cho
S. G.
Choi
E. J
.
(2012)
Arginine methylation-dependent regulation of ASK1 signaling by PRMT1
.
Cell Death Differ.
,
19
,
859
870
.

57.

Yu
Z.
Chen
T.
Hébert
J.
Li
E.
Richard
S
.
(2009)
A mouse PRMT1 null allele defines an essential role for arginine methylation in genome maintenance and cell proliferation
.
Mol. Cell. Biol.
,
29
,
2982
2996
.

58.

Shattuck-Eidens
D.
McClure
M.
Simard
J.
et al. 
(1995)
A collaborative survey of 80 mutations in the BRCA1 breast and ovarian cancer susceptibility gene. Implications for presymptomatic testing and screening
.
JAMA
,
273
,
535
541
.

59.

Guendel
I.
Carpio
L.
Pedati
C.
Schwartz
A.
Teal
C.
Kashanchi
F.
Kehn-Hall
K
.
(2010)
Methylation of the tumor suppressor protein, BRCA1, influences its transcriptional cofactor function
.
PLoS One
,
5
,
e11379
.

60.

Boisvert
F. M.
Déry
U.
Masson
J. Y.
Richard
S
.
(2005)
Arginine methylation of MRE11 by PRMT1 is required for DNA damage checkpoint control
.
Genes Dev.
,
19
,
671
676
.

61.

Boisvert
F. M.
Hendzel
M. J.
Masson
J. Y.
Richard
S
.
(2005)
Methylation of MRE11 regulates its nuclear compartmentalization
.
Cell Cycle
,
4
,
981
989
.

62.

Mitchell
T. R.
Glenfield
K.
Jeyanthan
K.
Zhu
X. D
.
(2009)
Arginine methylation regulates telomere length and stability
.
Mol. Cell. Biol.
,
29
,
4918
4934
.

63.

An
W.
Kim
J.
Roeder
R. G
.
(2004)
Ordered cooperative functions of PRMT1, p300, and CARM1 in transcriptional activation by p53
.
Cell
,
117
,
735
748
.

64.

Massagué
J
.
(2008)
TGFbeta in Cancer
.
Cell
,
134
,
215
230
.

65.

Ikushima
H.
Miyazono
K
.
(2012)
TGF-β signal transduction spreading to a wider field: a broad variety of mechanisms for context-dependent effects of TGF-β
.
Cell Tissue Res.
,
347
,
37
49
.

66.

Xu
J.
Wang
A. H.
Oses-Prieto
J.
et al. 
(2013)
Arginine Methylation Initiates BMP-Induced Smad Signaling
.
Mol. Cell
,
51
,
5
19
.

67.

ten Dijke
P.
Arthur
H. M
.
(2007)
Extracellular control of TGFbeta signalling in vascular development and disease
.
Nat. Rev. Mol. Cell Biol.
,
8
,
857
869
.

68.

Blanco Calvo
M.
Bolós Fernández
V.
Medina Villaamil
V.
Aparicio Gallego
G.
Díaz Prado
S.
Grande Pulido
E
.
(2009)
Biology of BMP signalling and cancer
.
Clin. Transl. Oncol.
,
11
,
126
137
.

69.

Lin
W. J.
Gary
J. D.
Yang
M. C.
Clarke
S.
Herschman
H. R
.
(1996)
The mammalian immediate-early TIS21 protein and the leukemia-associated BTG1 protein interact with a protein-arginine N-methyltransferase
.
J. Biol. Chem.
,
271
,
15034
15044
.

70.

Robin-Lespinasse
Y.
Sentis
S.
Kolytcheff
C.
Rostan
M. C.
Corbo
L.
Le Romancer
M
.
(2007)
hCAF1, a new regulator of PRMT1-dependent arginine methylation
.
J. Cell Sci.
,
120
,
638
647
.

71.

Choi
K. S.
Kim
J. Y.
Lim
S. K.
Choi
Y. W.
Kim
Y. H.
Kang
S. Y.
Park
T. J.
Lim
I. K
.
(2012)
TIS21(/BTG2/PC3) accelerates the repair of DNA double strand breaks by enhancing Mre11 methylation and blocking damage signal transfer to the Chk2(T68)-p53(S20) pathway
.
DNA Repair (Amst).
,
11
,
965
975
.

72.

Prévôt
D.
Morel
A. P.
Voeltzel
T.
Rostan
M. C.
Rimokh
R.
Magaud
J. P.
Corbo
L
.
(2001)
Relationships of the antiproliferative proteins BTG1 and BTG2 with CAF1, the human homolog of a component of the yeast CCR4 transcriptional complex: involvement in estrogen receptor alpha signaling pathway
.
J. Biol. Chem.
,
276
,
9640
9648
.

73.

Sheng
S. H.
Zhao
C. M.
Sun
G. G
.
(2014)
BTG1 expression correlates with the pathogenesis and progression of breast carcinomas
.
Tumour Biol.
,
35
,
3317
3326
.

74.

Zhu
R.
Zou
S. T.
Wan
J. M.
Li
W.
Li
X. L.
Zhu
W
.
(2013)
BTG1 inhibits breast cancer cell growth through induction of cell cycle arrest and apoptosis
.
Oncol. Rep.
,
30
,
2137
2144
.

75.

Duriez
C.
Falette
N.
Audoynaud
C.
Moyret-Lalle
C.
Bensaad
K.
Courtois
S.
Wang
Q.
Soussi
T.
Puisieux
A
.
(2002)
The human BTG2/TIS21/PC3 gene: genomic structure, transcriptional regulation and evaluation as a candidate tumor suppressor gene
.
Gene
,
282
,
207
214
.

76.

Kawakubo
H.
Carey
J. L.
Brachtel
E.
Gupta
V.
Green
J. E.
Walden
P. D.
Maheswaran
S
.
(2004)
Expression of the NF-kappaB-responsive gene BTG2 is aberrantly regulated in breast cancer
.
Oncogene
,
23
,
8310
8319
.

77.

Scott
H. S.
Antonarakis
S. E.
Lalioti
M. D.
Rossier
C.
Silver
P. A.
Henry
M. F
.
(1998)
Identification and characterization of two putative human arginine methyltransferases (HRMT1L1 and HRMT1L2)
.
Genomics
,
48
,
330
340
.

78.

Yoshimoto
T.
Boehm
M.
Olive
M.
Crook
M. F.
San
H.
Langenickel
T.
Nabel
E. G
.
(2006)
The arginine methyltransferase PRMT2 binds RB and regulates E2F function
.
Exp. Cell Res.
,
312
,
2040
2053
.

79.

Ganesh
L.
Yoshimoto
T.
Moorthy
N. C.
Akahata
W.
Boehm
M.
Nabel
E. G.
Nabel
G. J
.
(2006)
Protein methyltransferase 2 inhibits NF-kappaB function and promotes apoptosis
.
Mol. Cell. Biol.
,
26
,
3864
3874
.

80.

Lakowski
T. M.
Frankel
A
.
(2009)
Kinetic analysis of human protein arginine N-methyltransferase 2: formation of monomethyl- and asymmetric dimethyl-arginine residues on histone H4
.
Biochem. J.
,
421
,
253
261
.

81.

Blythe
S. A.
Cha
S. W.
Tadjuidje
E.
Heasman
J.
Klein
P. S
.
(2010)
beta-Catenin primes organizer gene expression by recruiting a histone H3 arginine 8 methyltransferase, Prmt2
.
Dev. Cell
,
19
,
220
231
.

82.

Lu
X. L.
Cao
X.
Liu
X. Y.
Jiao
B. H
.
(2010)
Recent progress of Src SH2 and SH3 inhibitors as anticancer agents
.
Curr. Med. Chem.
,
17
,
1117
1124
.

83.

Zhong
J.
Cao
R. X.
Hong
T.
Yang
J.
Zu
X. Y.
Xiao
X. H.
Liu
J. H.
Wen
G. B
.
(2011)
Identification and expression analysis of a novel transcript of the human PRMT2 gene resulted from alternative polyadenylation in breast cancer
.
Gene
,
487
,
1
9
.

84.

Zhong
J.
Cao
R. X.
Zu
X. Y.
et al. 
(2012)
Identification and characterization of novel spliced variants of PRMT2 in breast carcinoma
.
FEBS J.
,
279
,
316
335
.

85.

Qi
C.
Chang
J.
Zhu
Y.
Yeldandi
A. V.
Rao
S. M.
Zhu
Y. J
.
(2002)
Identification of protein arginine methyltransferase 2 as a coactivator for estrogen receptor alpha
.
J. Biol. Chem.
,
277
,
28624
28630
.

86.

Barrallo-Gimeno
A.
Nieto
M. A
.
(2005)
The Snail genes as inducers of cell movement and survival: implications in development and cancer
.
Development
,
132
,
3151
3161
.

87.

Zhong
J.
Cao
R.X.
Liu
J.H.
Liu
Y.B.
Wang
J.
Liu
L.P.
Chen
Y.J.
Yang
J.
Zhang
Q.H.
Wu
Y.
Ding
W.J.
Hong
T.
Xiao
X.H.
Zu
X.Y.
Wen
G.B
.
(2013)
Nuclear loss of protein arginine N-methyltransferase 2 in breast carcinoma is associated with tumor grade and overexpression of cyclin D1 protein
.
Oncogene
, doi:
10.1038/onc.2013.500
.

88.

Tang
J.
Gary
J. D.
Clarke
S.
Herschman
H. R
.
(1998)
PRMT 3, a type I protein arginine N-methyltransferase that differs from PRMT1 in its oligomerization, subcellular localization, substrate specificity, and regulation
.
J. Biol. Chem.
,
273
,
16935
16945
.

89.

Swiercz
R.
Cheng
D.
Kim
D.
Bedford
M. T
.
(2007)
Ribosomal protein rpS2 is hypomethylated in PRMT3-deficient mice
.
J. Biol. Chem.
,
282
,
16917
16923
.

90.

Choi
S.
Jung
C. R.
Kim
J. Y.
Im
D. S
.
(2008)
PRMT3 inhibits ubiquitination of ribosomal protein S2 and together forms an active enzyme complex
.
Biochim. Biophys. Acta
,
1780
,
1062
1069
.

91.

Lai
Y.
Song
M.
Hakala
K.
Weintraub
S. T.
Shiio
Y
.
(2011)
Proteomic dissection of the von Hippel-Lindau (VHL) interactome
.
J. Proteome Res.
,
10
,
5175
5182
.

92.

Heller
G.
Geradts
J.
Ziegler
B.
et al. 
(2007)
Downregulation of TSLC1 and DAL-1 expression occurs frequently in breast cancer
.
Breast Cancer Res. Treat.
,
103
,
283
291
.

93.

Takahashi
Y.
Iwai
M.
Kawai
T.
et al. 
(2012)
Aberrant expression of tumor suppressors CADM1 and 4.1B in invasive lesions of primary breast cancer
.
Breast Cancer
,
19
,
242
252
.

94.

Singh
V.
Miranda
T. B.
Jiang
W.
et al. 
(2004)
DAL-1/4.1B tumor suppressor interacts with protein arginine N-methyltransferase 3 (PRMT3) and inhibits its ability to methylate substrates in vitro and in vivo
.
Oncogene
,
23
,
7761
7771
.

95.

Yadav
N.
Lee
J.
Kim
J.
Shen
J.
Hu
M. C.
Aldaz
C. M.
Bedford
M. T
.
(2003)
Specific protein methylation defects and gene expression perturbations in coactivator-associated arginine methyltransferase 1-deficient mice
.
Proc. Natl. Acad. Sci. U. S. A.
,
100
,
6464
6468
.

96.

Yadav
N.
Cheng
D.
Richard
S.
Morel
M.
Iyer
V. R.
Aldaz
C. M.
Bedford
M. T
.
(2008)
CARM1 promotes adipocyte differentiation by coactivating PPARgamma
.
EMBO Rep.
,
9
,
193
198
.

97.

O’Brien
K. B.
Alberich-Jordà
M.
Yadav
N.
et al. 
(2010)
CARM1 is required for proper control of proliferation and differentiation of pulmonary epithelial cells
.
Development
,
137
,
2147
2156
.

98.

Kim
J.
Lee
J.
Yadav
N.
Wu
Q.
Carter
C.
Richard
S.
Richie
E.
Bedford
M. T
.
(2004)
Loss of CARM1 results in hypomethylation of thymocyte cyclic AMP-regulated phosphoprotein and deregulated early T cell development
.
J. Biol. Chem.
,
279
,
25339
25344
.

99.

Ito
T.
Yadav
N.
Lee
J.
et al. 
(2009)
Arginine methyltransferase CARM1/PRMT4 regulates endochondral ossification
.
BMC Dev. Biol.
,
9
,
47
.

100.

Kim
D.
Lee
J.
Cheng
D.
Li
J.
Carter
C.
Richie
E.
Bedford
M. T
.
(2010)
Enzymatic activity is required for the in vivo functions of CARM1
.
J. Biol. Chem.
,
285
,
1147
1152
.

101.

Kawabe
Y.
Wang
Y. X.
McKinnell
I. W.
Bedford
M. T.
Rudnicki
M. A
.
(2012)
Carm1 regulates Pax7 transcriptional activity through MLL1/2 recruitment during asymmetric satellite stem cell divisions
.
Cell Stem Cell
,
11
,
333
345
.

102.

Dacwag
C. S.
Bedford
M. T.
Sif
S.
Imbalzano
A. N
.
(2009)
Distinct protein arginine methyltransferases promote ATP-dependent chromatin remodeling function at different stages of skeletal muscle differentiation
.
Mol. Cell. Biol.
,
29
,
1909
1921
.

103.

Chen
D.
Ma
H.
Hong
H.
Koh
S. S.
Huang
S. M.
Schurter
B. T.
Aswad
D. W.
Stallcup
M. R
.
(1999)
Regulation of transcription by a protein methyltransferase
.
Science
,
284
,
2174
2177
.

104.

Ohkura
N.
Takahashi
M.
Yaguchi
H.
Nagamura
Y.
Tsukada
T
.
(2005)
Coactivator-associated arginine methyltransferase 1, CARM1, affects pre-mRNA splicing in an isoform-specific manner
.
J. Biol. Chem.
,
280
,
28927
28935
.

105.

El Messaoudi
S.
Fabbrizio
E.
Rodriguez
C.
et al. 
(2006)
Coactivator-associated arginine methyltransferase 1 (CARM1) is a positive regulator of the Cyclin E1 gene
.
Proc. Natl. Acad. Sci. U. S. A.
,
103
,
13351
13356
.

106.

Lee
Y. H.
Bedford
M. T.
Stallcup
M. R
.
(2011)
Regulated recruitment of tumor suppressor BRCA1 to the p21 gene by coactivator methylation
.
Genes Dev.
,
25
,
176
188
.

107.

Lee
Y. H.
Stallcup
M. R
.
(2011)
Roles of protein arginine methylation in DNA damage signaling pathways is CARM1 a life-or-death decision point?
Cell Cycle
,
10
,
1343
1344
.

108.

Lupien
M.
Eeckhoute
J.
Meyer
C. A.
Krum
S. A.
Rhodes
D. R.
Liu
X. S.
Brown
M
.
(2009)
Coactivator function defines the active estrogen receptor alpha cistrome
.
Mol. Cell. Biol.
,
29
,
3413
3423
.

109.

Schurter
B. T.
Koh
S. S.
Chen
D.
et al. 
(2001)
Methylation of histone H3 by coactivator-associated arginine methyltransferase 1
.
Biochemistry
,
40
,
5747
5756
.

110.

Zhao
H. Y.
Zhang
Y. J.
Dai
H.
Zhang
Y.
Shen
Y. F
.
(2011)
CARM1 mediates modulation of Sox2
.
PLoS One
,
6
,
e27026
.

111.

Ceschin
D. G.
Walia
M.
Wenk
S. S.
et al. 
(2011)
Methylation specifies distinct estrogen-induced binding site repertoires of CBP to chromatin
.
Genes Dev.
,
25
,
1132
1146
.

112.

Naeem
H.
Cheng
D.
Zhao
Q.
Underhill
C.
Tini
M.
Bedford
M. T.
Torchia
J
.
(2007)
The activity and stability of the transcriptional coactivator p/CIP/SRC-3 are regulated by CARM1-dependent methylation
.
Mol. Cell. Biol.
,
27
,
120
134
.

113.

Feng
Q.
Yi
P.
Wong
J.
O’Malley
B. W
.
(2006)
Signaling within a coactivator complex: methylation of SRC-3/AIB1 is a molecular switch for complex disassembly
.
Mol. Cell. Biol.
,
26
,
7846
7857
.

114.

Chevillard-Briet
M.
Trouche
D.
Vandel
L
.
(2002)
Control of CBP co-activating activity by arginine methylation
.
EMBO J.
,
21
,
5457
5466
.

115.

Sims
R. J.
3rd
Rojas
L. A.
Beck
D.
Bonasio
R.
Schüller
R.
Drury
W. J.
3rd
Eick
D.
Reinberg
D
.
(2011)
The C-terminal domain of RNA polymerase II is modified by site-specific methylation
.
Science
,
332
,
99
103
.

116.

Carascossa
S.
Dudek
P.
Cenni
B.
Briand
P. A.
Picard
D
.
(2010)
CARM1 mediates the ligand-independent and tamoxifen-resistant activation of the estrogen receptor alpha by cAMP
.
Genes Dev.
,
24
,
708
719
.

117.

Kuhn
P.
Chumanov
R.
Wang
Y.
Ge
Y.
Burgess
R. R.
Xu
W
.
(2011)
Automethylation of CARM1 allows coupling of transcription and mRNA splicing
.
Nucleic Acids Res.
,
39
,
2717
2726
.

118.

Frietze
S.
Lupien
M.
Silver
P. A.
Brown
M
.
(2008)
CARM1 regulates estrogen-stimulated breast cancer growth through up-regulation of E2F1
.
Cancer Res.
,
68
,
301
306
.

119.

Mann
M.
Cortez
V.
Vadlamudi
R
.
(2013)
PELP1 oncogenic functions involve CARM1 regulation
.
Carcinogenesis
,
34
,
1468
1475
.

120.

Vadlamudi
R. K.
Manavathi
B.
Balasenthil
S.
Nair
S. S.
Yang
Z.
Sahin
A. A.
Kumar
R
.
(2005)
Functional implications of altered subcellular localization of PELP1 in breast cancer cells
.
Cancer Res.
,
65
,
7724
7732
.

121.

Gururaj
A. E.
Peng
S.
Vadlamudi
R. K.
Kumar
R
.
(2007)
Estrogen induces expression of BCAS3, a novel estrogen receptor-alpha coactivator, through proline-, glutamic acid-, and leucine-rich protein-1 (PELP1)
.
Mol. Endocrinol.
,
21
,
1847
1860
.

122.

Rajhans
R.
Nair
S.
Holden
A. H.
Kumar
R.
Tekmal
R. R.
Vadlamudi
R. K
.
(2007)
Oncogenic potential of the nuclear receptor coregulator proline-, glutamic acid-, leucine-rich protein 1/modulator of the nongenomic actions of the estrogen receptor
.
Cancer Res.
,
67
,
5505
5512
.

123.

Habashy
H. O.
Powe
D. G.
Rakha
E. A.
Ball
G.
Macmillan
R. D.
Green
A. R.
Ellis
I. O
.
(2010)
The prognostic significance of PELP1 expression in invasive breast cancer with emphasis on the ER-positive luminal-like subtype
.
Breast Cancer Res. Treat.
,
120
,
603
612
.

124.

Roy
S.
Chakravarty
D.
Cortez
V.
et al. 
(2012)
Significance of PELP1 in ER-negative breast cancer metastasis
.
Mol. Cancer Res.
,
10
,
25
33
.

125.

Cheng
H.
Qin
Y.
Fan
H.
Su
P.
Zhang
X.
Zhang
H.
Zhou
G
.
(2013)
Overexpression of CARM1 in breast cancer is correlated with poorly characterized clinicopathologic parameters and molecular subtypes
.
Diagn. Pathol.
,
8
,
129
.

126.

Habashy
H. O.
Rakha
E. A.
Ellis
I. O.
Powe
D. G
.
(2013)
The oestrogen receptor coactivator CARM1 has an oncogenic effect and is associated with poor prognosis in breast cancer
.
Breast Cancer Res. Treat.
,
140
,
307
316
.

127.

Davis
M. B.
Liu
X.
Wang
S.
Reeves
J.
Khramtsov
A.
Huo
D.
Olopade
O. I
.
(2013)
Expression and sub-cellular localization of an epigenetic regulator, co-activator arginine methyltransferase 1 (CARM1), is associated with specific breast cancer subtypes and ethnicity
.
Mol. Cancer
,
12
,
40
.

128.

Lee
D. Y.
Northrop
J. P.
Kuo
M. H.
Stallcup
M. R
.
(2006)
Histone H3 lysine 9 methyltransferase G9a is a transcriptional coactivator for nuclear receptors
.
J. Biol. Chem.
,
281
,
8476
8485
.

129.

Al-Dhaheri
M.
Wu
J.
Skliris
G. P.
et al. 
(2011)
CARM1 is an important determinant of ERα-dependent breast cancer cell differentiation and proliferation in breast cancer cells
.
Cancer Res.
,
71
,
2118
2128
.

130.

Wang
L.
Zhao
Z.
Meyer
M. B.
et al. 
(2014)
CARM1 methylates chromatin remodeling factor BAF155 to enhance tumor progression and metastasis
.
Cancer Cell
,
25
,
21
36
.

131.

Herrmann
F.
Pably
P.
Eckerich
C.
Bedford
M. T.
Fackelmayer
F. O
.
(2009)
Human protein arginine methyltransferases in vivo–distinct properties of eight canonical members of the PRMT family
.
J. Cell Sci.
,
122
,
667
677
.

132.

Tee
W. W.
Pardo
M.
Theunissen
T. W.
Yu
L.
Choudhary
J. S.
Hajkova
P.
Surani
M. A
.
(2010)
Prmt5 is essential for early mouse development and acts in the cytoplasm to maintain ES cell pluripotency
.
Genes Dev.
,
24
,
2772
2777
.

133.

Fabbrizio
E.
El Messaoudi
S.
Polanowska
J.
et al. 
(2002)
Negative regulation of transcription by the type II arginine methyltransferase PRMT5
.
EMBO Rep.
,
3
,
641
645
.

134.

Deng
X.
Gu
L.
Liu
C.
et al. 
(2010)
Arginine methylation mediated by the Arabidopsis homolog of PRMT5 is essential for proper pre-mRNA splicing
.
Proc. Natl. Acad. Sci. U. S. A.
,
107
,
19114
19119
.

135.

Sanchez
S. E.
Petrillo
E.
Beckwith
E. J.
et al. 
(2010)
A methyl transferase links the circadian clock to the regulation of alternative splicing
.
Nature
,
468
,
112
116
.

136.

Bezzi
M.
Teo
S. X.
Muller
J.
Mok
W. C.
Sahu
S. K.
Vardy
L. A.
Bonday
Z. Q.
Guccione
E
.
(2013)
Regulation of constitutive and alternative splicing by PRMT5 reveals a role for Mdm4 pre-mRNA in sensing defects in the spliceosomal machinery
.
Genes Dev.
,
27
,
1903
1916
.

137.

Jansson
M.
Durant
S. T.
Cho
E. C.
Sheahan
S.
Edelmann
M.
Kessler
B.
La Thangue
N. B
.
(2008)
Arginine methylation regulates the p53 response
.
Nat. Cell Biol.
,
10
,
1431
1439
.

138.

Andreu-Pérez
P.
Esteve-Puig
R.
de Torre-Minguela
C.
et al. 
(2011)
Protein arginine methyltransferase 5 regulates ERK1/2 signal transduction amplitude and cell fate through CRAF
.
Sci. Signal.
,
4
,
ra58
.

139.

Hsu
J. M.
Chen
C. T.
Chou
C. K.
et al. 
(2011)
Crosstalk between Arg 1175 methylation and Tyr 1173 phosphorylation negatively modulates EGFR-mediated ERK activation
.
Nat. Cell Biol.
,
13
,
174
181
.

140.

Cho
E. C.
Zheng
S.
Munro
S.
et al. 
(2012)
Arginine methylation controls growth regulation by E2F-1
.
EMBO J.
,
31
,
1785
1797
.

141.

Ren
J.
Wang
Y.
Liang
Y.
Zhang
Y.
Bao
S.
Xu
Z
.
(2010)
Methylation of ribosomal protein S10 by protein-arginine methyltransferase 5 regulates ribosome biogenesis
.
J. Biol. Chem.
,
285
,
12695
12705
.

142.

Zhou
Z.
Sun
X.
Zou
Z.
et al. 
(2010)
PRMT5 regulates Golgi apparatus structure through methylation of the golgin GM130
.
Cell Res.
,
20
,
1023
1033
.

143.

Dacwag
C. S.
Ohkawa
Y.
Pal
S.
Sif
S.
Imbalzano
A. N
.
(2007)
The protein arginine methyltransferase Prmt5 is required for myogenesis because it facilitates ATP-dependent chromatin remodeling
.
Mol. Cell. Biol.
,
27
,
384
394
.

144.

Mallappa
C.
Hu
Y. J.
Shamulailatpam
P.
Tae
S.
Sif
S.
Imbalzano
A. N
.
(2011)
The expression of myogenic microRNAs indirectly requires protein arginine methyltransferase (Prmt)5 but directly requires Prmt4
.
Nucleic Acids Res.
,
39
,
1243
1255
.

145.

Ancelin
K.
Lange
U. C.
Hajkova
P.
Schneider
R.
Bannister
A. J.
Kouzarides
T.
Surani
M. A
.
(2006)
Blimp1 associates with Prmt5 and directs histone arginine methylation in mouse germ cells
.
Nat. Cell Biol.
,
8
,
623
630
.

146.

Eckert
D.
Biermann
K.
Nettersheim
D.
Gillis
A. J.
Steger
K.
Jäck
H. M.
Müller
A. M.
Looijenga
L. H.
Schorle
H
.
(2008)
Expression of BLIMP1/PRMT5 and concurrent histone H2A/H4 arginine 3 dimethylation in fetal germ cells, CIS/IGCNU and germ cell tumors
.
BMC Dev. Biol.
,
8
,
106
.

147.

Powers
M. A.
Fay
M. M.
Factor
R. E.
Welm
A. L.
Ullman
K. S
.
(2011)
Protein arginine methyltransferase 5 accelerates tumor growth by arginine methylation of the tumor suppressor programmed cell death 4
.
Cancer Res.
,
71
,
5579
5587
.

148.

Wen
Y. H.
Shi
X.
Chiriboga
L.
Matsahashi
S.
Yee
H.
Afonja
O
.
(2007)
Alterations in the expression of PDCD4 in ductal carcinoma of the breast
.
Oncol. Rep.
,
18
,
1387
1393
.

149.

Yang
H. S.
Knies
J. L.
Stark
C.
Colburn
N. H
.
(2003)
Pdcd4 suppresses tumor phenotype in JB6 cells by inhibiting AP-1 transactivation
.
Oncogene
,
22
,
3712
3720
.

150.

Jiang
W.
Roemer
M. E.
Newsham
I. F
.
(2005)
The tumor suppressor DAL-1/4.1B modulates protein arginine N-methyltransferase 5 activity in a substrate-specific manner
.
Biochem. Biophys. Res. Commun.
,
329
,
522
530
.

151.

Jiang
W.
Newsham
I. F
.
(2006)
The tumor suppressor DAL-1/4.1B and protein methylation cooperate in inducing apoptosis in MCF-7 breast cancer cells
.
Mol. Cancer
,
5
,
4
.

152.

Singhroy
D. N.
Mesplède
T.
Sabbah
A.
Quashie
P. K.
Falgueyret
J. P.
Wainberg
M. A
.
(2013)
Automethylation of protein arginine methyltransferase 6 (PRMT6) regulates its stability and its anti-HIV-1 activity
.
Retrovirology
,
10
,
73
.

153.

Neault
M.
Mallette
F. A.
Vogel
G.
Michaud-Levesque
J.
Richard
S
.
(2012)
Ablation of PRMT6 reveals a role as a negative transcriptional regulator of the p53 tumor suppressor
.
Nucleic Acids Res.
,
40
,
9513
9521
.

154.

Hyllus
D.
Stein
C.
Schnabel
K.
Schiltz
E.
Imhof
A.
Dou
Y.
Hsieh
J.
Bauer
U. M
.
(2007)
PRMT6-mediated methylation of R2 in histone H3 antagonizes H3 K4 trimethylation
.
Genes Dev.
,
21
,
3369
3380
.

155.

Michaud-Levesque
J.
Richard
S
.
(2009)
Thrombospondin-1 is a transcriptional repression target of PRMT6
.
J. Biol. Chem.
,
284
,
21338
21346
.

156.

Herglotz
J.
Kuvardina
O. N.
Kolodziej
S.
Kumar
A.
Hussong
H.
Grez
M.
Lausen
J
.
(2013)
Histone arginine methylation keeps RUNX1 target genes in an intermediate state
.
Oncogene
,
32
,
2565
2575
.

157.

Phalke
S.
Mzoughi
S.
Bezzi
M.
et al. 
(2012)
p53-Independent regulation of p21Waf1/Cip1 expression and senescence by PRMT6
.
Nucleic Acids Res.
,
40
,
9534
9542
.

158.

Casadio
F.
Lu
X.
Pollock
S. B.
LeRoy
G.
Garcia
B. A.
Muir
T. W.
Roeder
R. G.
Allis
C. D
.
(2013)
H3R42me2a is a histone modification with positive transcriptional effects
.
Proc. Natl. Acad. Sci. U. S. A.
,
110
,
14894
14899
.

159.

Waldmann
T.
Izzo
A.
Kamieniarz
K.
et al. 
(2011)
Methylation of H2AR29 is a novel repressive PRMT6 target
.
Epigenetics Chromatin
,
4
,
11
.

160.

Dowhan
D. H.
Harrison
M. J.
Eriksson
N. A.
et al. 
(2012)
Protein arginine methyltransferase 6-dependent gene expression and splicing: association with breast cancer outcomes
.
Endocr. Relat. Cancer
,
19
,
509
526
.

161.

Kim
N. H.
Kim
S. N.
Seo
D. W.
Han
J. W.
Kim
Y. K
.
(2013)
PRMT6 overexpression upregulates TSP-1 and downregulates MMPs: its implication in motility and invasion
.
Biochem. Biophys. Res. Commun.
,
432
,
60
65
.

162.

Chu
M. C.
Selam
F. B.
Taylor
H. S
.
(2004)
HOXA10 regulates p53 expression and matrigel invasion in human breast cancer cells
.
Cancer Biol. Ther.
,
3
,
568
572
.

163.

Somiari
S. B.
Somiari
R. I.
Heckman
C. M.
Olsen
C. H.
Jordan
R. M.
Russell
S. J.
Shriver
C. D
.
(2006)
Circulating MMP2 and MMP9 in breast cancer – potential role in classification of patients into low risk, high risk, benign disease and breast cancer categories
.
Int. J. Cancer
,
119
,
1403
1411
.

164.

Somiari
S. B.
Shriver
C. D.
Heckman
C.
et al. 
(2006)
Plasma concentration and activity of matrix metalloproteinase 2 and 9 in patients with breast disease, breast cancer and at risk of developing breast cancer
.
Cancer Lett.
,
233
,
98
107
.

165.

Qin
L.
Liao
L.
Redmond
A.
Young
L.
Yuan
Y.
Chen
H.
O’Malley
B. W.
Xu
J
.
(2008)
The AIB1 oncogene promotes breast cancer metastasis by activation of PEA3-mediated matrix metalloproteinase 2 (MMP2) and MMP9 expression
.
Mol. Cell. Biol.
,
28
,
5937
5950
.

166.

Wang
X.
Lu
H.
Urvalek
A. M.
Li
T.
Yu
L.
Lamar
J.
DiPersio
C. M.
Feustel
P. J.
Zhao
J
.
(2011)
KLF8 promotes human breast cancer cell invasion and metastasis by transcriptional activation of MMP9
.
Oncogene
,
30
,
1901
1911
.

167.

Liu
Y.
Xin
T.
Jiang
Q. Y.
et al. 
(2013)
CD147, MMP9 expression and clinical significance of basal-like breast cancer
.
Med. Oncol.
,
30
,
366
.

168.

Prasad
C. P.
Chaurasiya
S. K.
Axelsson
L.
Andersson
T
.
(2013)
WNT-5A triggers Cdc42 activation leading to an ERK1/2 dependent decrease in MMP9 activity and invasive migration of breast cancer cells
.
Mol. Oncol.
,
7
,
870
883
.

169.

Harrison
M. J.
Tang
Y. H.
Dowhan
D. H
.
(2010)
Protein arginine methyltransferase 6 regulates multiple aspects of gene expression
.
Nucleic Acids Res.
,
38
,
2201
2216
.

170.

Miranda
T. B.
Webb
K. J.
Edberg
D. D.
Reeves
R.
Clarke
S
.
(2005)
Protein arginine methyltransferase 6 specifically methylates the nonhistone chromatin protein HMGA1a
.
Biochem. Biophys. Res. Commun.
,
336
,
831
835
.

171.

Zou
Y.
Webb
K.
Perna
A. D.
Zhang
Q.
Clarke
S.
Wang
Y
.
(2007)
A mass spectrometric study on the in vitro methylation of HMGA1a and HMGA1b proteins by PRMTs: methylation specificity, the effect of binding to AT-rich duplex DNA, and the effect of C-terminal phosphorylation
.
Biochemistry
,
46
,
7896
7906
.

172.

Liu
F.
Chau
K. Y.
Arlotta
P.
Ono
S. J
.
(2001)
The HMG I proteins: dynamic roles in gene activation, development, and tumorigenesis
.
Immunol. Res.
,
24
,
13
29
.

173.

Reeves
R.
Beckerbauer
L
.
(2001)
HMGI/Y proteins: flexible regulators of transcription and chromatin structure
.
Biochim. Biophys. Acta
,
1519
,
13
29
.

174.

Dolde
C. E.
Mukherjee
M.
Cho
C.
Resar
L. M
.
(2002)
HMG-I/Y in human breast cancer cell lines
.
Breast Cancer Res. Treat.
,
71
,
181
191
.

175.

Flohr
A. M.
Rogalla
P.
Bonk
U.
et al. 
(2003)
High mobility group protein HMGA1 expression in breast cancer reveals a positive correlation with tumour grade
.
Histol. Histopathol.
,
18
,
999
1004
.

176.

Chiappetta
G.
Botti
G.
Monaco
M.
et al. 
(2004)
HMGA1 protein overexpression in human breast carcinomas: correlation with ErbB2 expression
.
Clin. Cancer Res.
,
10
,
7637
7644
.

177.

Chiappetta
G.
Ottaiano
A.
Vuttariello
E.
et al. 
(2010)
HMGA1 protein expression in familial breast carcinoma patients
.
Eur. J. Cancer
,
46
,
332
339
.

178.

Treff
N. R.
Pouchnik
D.
Dement
G. A.
Britt
R. L.
Reeves
R
.
(2004)
High-mobility group A1a protein regulates Ras/ERK signaling in MCF-7 human breast cancer cells
.
Oncogene
,
23
,
777
785
.

179.

Stimpfl
M.
Tong
D.
Fasching
B.
Schuster
E.
Obermair
A.
Leodolter
S.
Zeillinger
R
.
(2002)
Vascular endothelial growth factor splice variants and their prognostic value in breast and ovarian cancer
.
Clin. Cancer Res.
,
8
,
2253
2259
.

180.

Wang
L.
Duke
L.
Zhang
P. S.
Arlinghaus
R. B.
Symmans
W. F.
Sahin
A.
Mendez
R.
Dai
J. L
.
(2003)
Alternative splicing disrupts a nuclear localization signal in spleen tyrosine kinase that is required for invasion suppression in breast cancer
.
Cancer Res.
,
63
,
4724
4730
.

181.

Albertella
M. R.
Lau
A.
O’Connor
M. J
.
(2005)
The overexpression of specialized DNA polymerases in cancer
.
DNA Repair (Amst).
,
4
,
583
593
.

182.

Abdel-Fatah
T. M.
Russell
R.
Agarwal
D.
et al. 
(2014)
DNA polymerase β deficiency is linked to aggressive breast cancer: a comprehensive analysis of gene copy number, mRNA and protein expression in multiple cohorts
.
Mol. Oncol.
,
8
,
520
532
.

183.

Newman
L.
Xia
W.
Yang
H. Y.
Sahin
A.
Bondy
M.
Lukmanji
F.
Hung
M. C.
Lee
M. H
.
(2001)
Correlation of p27 protein expression with HER-2/neu expression in breast cancer
.
Mol. Carcinog.
,
30
,
169
175
.

184.

Bearss
D. J.
Lee
R. J.
Troyer
D. A.
Pestell
R. G.
Windle
J. J
.
(2002)
Differential effects of p21(WAF1/CIP1) deficiency on MMTV-ras and MMTV-myc mammary tumor properties
.
Cancer Res.
,
62
,
2077
2084
.

185.

Spataro
V. J.
Litman
H.
Viale
G.
et al.  ;
International Breast Cancer Study Group
.
(2003)
Decreased immunoreactivity for p27 protein in patients with early-stage breast carcinoma is correlated with HER-2/neu overexpression and with benefit from one course of perioperative chemotherapy in patients with negative lymph node status: results from International Breast Cancer Study Group Trial V
.
Cancer
,
97
,
1591
1600
.

186.

Alkarain
A.
Jordan
R.
Slingerland
J
.
(2004)
p27 deregulation in breast cancer: prognostic significance and implications for therapy
.
J. Mammary Gland Biol. Neoplasia
,
9
,
67
80
.

187.

Crawford
Y. G.
Gauthier
M. L.
Joubel
A.
Mantei
K.
Kozakiewicz
K.
Afshari
C. A.
Tlsty
T. D
.
(2004)
Histologically normal human mammary epithelia with silenced p16(INK4a) overexpress COX-2, promoting a premalignant program
.
Cancer Cell
,
5
,
263
273
.

188.

Stein
C.
Riedl
S.
Rüthnick
D.
Nötzold
R. R.
Bauer
U. M
.
(2012)
The arginine methyltransferase PRMT6 regulates cell proliferation and senescence through transcriptional repression of tumor suppressor genes
.
Nucleic Acids Res.
,
40
,
9522
9533
.

189.

Kleinschmidt
M. A.
de Graaf
P.
van Teeffelen
H. A.
Timmers
H. T
.
(2012)
Cell cycle regulation by the PRMT6 arginine methyltransferase through repression of cyclin-dependent kinase inhibitors
.
PLoS One
,
7
,
e41446
.

190.

Wang
X.
Huang
Y.
Zhao
J.
Zhang
Y.
Lu
J.
Huang
B
.
(2012)
Suppression of PRMT6-mediated arginine methylation of p16 protein potentiates its ability to arrest A549 cell proliferation
.
Int. J. Biochem. Cell Biol.
,
44
,
2333
2341
.

191.

Miranda
T. B.
Miranda
M.
Frankel
A.
Clarke
S
.
(2004)
PRMT7 is a member of the protein arginine methyltransferase family with a distinct substrate specificity
.
J. Biol. Chem.
,
279
,
22902
22907
.

192.

Hasegawa
M.
Toma-Fukai
S.
Kim
J. D.
Fukamizu
A.
Shimizu
T
.
(2014)
Protein arginine methyltransferase 7 has a novel homodimer-like structure formed by tandem repeats
.
FEBS Lett.
,
588
,
1942
1948
.

193.

Lee
J. H.
Cook
J. R.
Yang
Z. H.
Mirochnitchenko
O.
Gunderson
S. I.
Felix
A. M.
Herth
N.
Hoffmann
R.
Pestka
S
.
(2005)
PRMT7, a new protein arginine methyltransferase that synthesizes symmetric dimethylarginine
.
J. Biol. Chem.
,
280
,
3656
3664
.

194.

Karkhanis
V.
Wang
L.
Tae
S.
Hu
Y. J.
Imbalzano
A. N.
Sif
S
.
(2012)
Protein arginine methyltransferase 7 regulates cellular response to DNA damage by methylating promoter histones H2A and H4 of the polymerase δ catalytic subunit gene, POLD1
.
J. Biol. Chem.
,
287
,
29801
29814
.

195.

Thomassen
M.
Tan
Q.
Kruse
T. A
.
(2009)
Gene expression meta-analysis identifies chromosomal regions and candidate genes involved in breast cancer metastasis
.
Breast Cancer Res. Treat.
,
113
,
239
249
.

196.

Jelinic
P.
Stehle
J. C.
Shaw
P
.
(2006)
The testis-specific factor CTCFL cooperates with the protein methyltransferase PRMT7 in H19 imprinting control region methylation
.
PLoS Biol.
,
4
,
e355
.

197.

Martin-Kleiner
I
.
(2012)
BORIS in human cancers – a review
.
Eur. J. Cancer
,
48
,
929
935
.

198.

Yost
J. M.
Korboukh
I.
Liu
F.
Gao
C.
Jin
J
.
(2011)
Targets in epigenetics: inhibiting the methyl writers of the histone code
.
Curr. Chem. Genomics
,
5
,
72
84
.

199.

Vhuiyan
M.
Thomas
D.
Hossen
F.
Frankel
A
.
(2013)
Targeting protein arginine N-methyltransferases with peptide-based inhibitors: opportunities and challenges
.
Future Med. Chem.
,
5
,
2199
2206
.

200.

Cheng
D.
Yadav
N.
King
R. W.
Swanson
M. S.
Weinstein
E. J.
Bedford
M. T
.
(2004)
Small molecule regulators of protein arginine methyltransferases
.
J. Biol. Chem.
,
279
,
23892
23899
.

201.

Ragno
R.
Simeoni
S.
Castellano
S.
et al. 
(2007)
Small molecule inhibitors of histone arginine methyltransferases: homology modeling, molecular docking, binding mode analysis, and biological evaluations
.
J. Med. Chem.
,
50
,
1241
1253
.

202.

Bonham
K.
Hemmers
S.
Lim
Y. H.
Hill
D. M.
Finn
M. G.
Mowen
K. A
.
(2010)
Effects of a novel arginine methyltransferase inhibitor on T-helper cell cytokine production
.
FEBS J.
,
277
,
2096
2108
.

203.

Purandare
A. V.
Chen
Z.
Huynh
T.
et al. 
(2008)
Pyrazole inhibitors of coactivator associated arginine methyltransferase 1 (CARM1)
.
Bioorg. Med. Chem. Lett.
,
18
,
4438
4441
.

204.

Therrien
E.
Larouche
G.
Manku
S.
et al. 
(2009)
1,2-Diamines as inhibitors of co-activator associated arginine methyltransferase 1 (CARM1)
.
Bioorg. Med. Chem. Lett.
,
19
,
6725
6732
.

205.

Spannhoff
A.
Heinke
R.
Bauer
I.
et al. 
(2007)
Target-based approach to inhibitors of histone arginine methyltransferases
.
J. Med. Chem.
,
50
,
2319
2325
.

206.

Bissinger
E. M.
Heinke
R.
Spannhoff
A.
et al. 
(2011)
Acyl derivatives of p-aminosulfonamides and dapsone as new inhibitors of the arginine methyltransferase hPRMT1
.
Bioorg. Med. Chem.
,
19
,
3717
3731
.

207.

Wang
J.
Chen
L.
Sinha
S. H.
et al. 
(2012)
Pharmacophore-based virtual screening and biological evaluation of small molecule inhibitors for protein arginine methylation
.
J. Med. Chem.
,
55
,
7978
7987
.

208.

Yan
L.
Yan
C.
Qian
K.
et al. 
(2014)
Diamidine compounds for selective inhibition of protein arginine methyltransferase 1
.
J. Med. Chem.
,
57
,
2611
2622
.

209.

Cheng
D.
Bedford
M. T
.
(2011)
Xenoestrogens regulate the activity of arginine methyltransferases
.
Chembiochem
,
12
,
323
329
.