Next Article in Journal
Comparative Analysis of Computer-Aided Diagnosis and Computer-Assisted Subjective Assessment in Thyroid Ultrasound
Next Article in Special Issue
Structure, Activity and Function of the PRMT2 Protein Arginine Methyltransferase
Previous Article in Journal
A Multiallelic Molecular Beacon-Based Real-Time RT-PCR Assay for the Detection of SARS-CoV-2
Previous Article in Special Issue
Activity and Function of the PRMT8 Protein Arginine Methyltransferase in Neurons
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Structure, Activity, and Function of PRMT1

by
Charlène Thiebaut
1,2,3,
Louisane Eve
1,2,3,
Coralie Poulard
1,2,3 and
Muriel Le Romancer
1,2,3,*
1
Université de Lyon, F-69000 Lyon, France
2
Inserm U1052, Centre de Recherche en Cancérologie de Lyon, F-69000 Lyon, France
3
CNRS UMR5286, Centre de Recherche en Cancérologie de Lyon, F-69000 Lyon, France
*
Author to whom correspondence should be addressed.
Life 2021, 11(11), 1147; https://doi.org/10.3390/life11111147
Submission received: 3 October 2021 / Revised: 25 October 2021 / Accepted: 25 October 2021 / Published: 27 October 2021
(This article belongs to the Special Issue Structure, Activity, and Function of Protein Methyltransferases)

Abstract

:
PRMT1, the major protein arginine methyltransferase in mammals, catalyzes monomethylation and asymmetric dimethylation of arginine side chains in proteins. Initially described as a regulator of chromatin dynamics through the methylation of histone H4 at arginine 3 (H4R3), numerous non-histone substrates have since been identified. The variety of these substrates underlines the essential role played by PRMT1 in a large number of biological processes such as transcriptional regulation, signal transduction or DNA repair. This review will provide an overview of the structural, biochemical and cellular features of PRMT1. After a description of the genomic organization and protein structure of PRMT1, special consideration was given to the regulation of PRMT1 enzymatic activity. Finally, we discuss the involvement of PRMT1 in embryonic development, DNA damage repair, as well as its participation in the initiation and progression of several types of cancers.

1. Introduction

Arginine methylation is a common and widespread post-translational modification (PTM) in eukaryotes that regulates numerous biological processes. Currently, nine protein arginine methyltransferases (PRMTs) have been described which are divided into three families according to the type of methylarginine produced. Type I PRMTs (PRMT-1, 2, 3, 4, 6 and 8) generate ω-NG-monomethylarginine (MMA) and ω-NG, NG-asymmetric dimethylarginine (ADMA), Type II PRMTs (PRMT-5 and 9) generate MMA and ω-NG, N’G-symmetric dimethylarginine (SDMA) and finally the unique Type III PRMT, PRMT7, generates MMA. Mechanistically, all PRMTs catalyze the transfer of a methyl group from S-adenosyl methionine (AdoMet) to the guanidino nitrogen atom of arginine [1]. Though considered for a long time as a stable mark, it is now well-known that arginine methylation is a dynamic PTM that can be removed by arginine demethylases [2].
PRMT1, which is the major type I PRMT, is responsible for 85% of the activity attributed to type I PRMTs in mammals [3]. Moreover, it plays key roles in various cellular processes such as transcriptional regulation, signal transduction or DNA damage repair, owing to the diversity of its histone and non-histone substrates [1].
The aim of this review is to provide an overview of the literature concerning PRMT1 structure, activities and functions. After a detailed description of the genomic organization and the protein structures of the different PRMT1 isoforms, the substrate specificity and the regulatory mechanisms of PRMT1 itself will be discussed. Finally, the cellular roles and functions of PRMT1, as well as its involvement in cancer, will be addressed.

2. Structural Features

2.1. Genomic Organization

Human PRMT1 is encoded by the PRMT1 gene located on chromosome 19 (19q13.3) and composed of 12 exons and 11 introns. At the 5′ end of this genomic locus of 11.3 kilobases (kb) are four alternative exons (e1a-e1d) involved in the synthesis of at least seven splice variants of PRMT1 (v1–v7) [4,5] (Figure 1A,B). More recently, next-generation sequencing led to the identification of a novel exon located between exons 11 and 12, and 58 additional alternative splice variants of the PRMT1 gene. Among them, 34 are speculated to encode additional protein isoforms of PRMT1 but remain to be characterized [6].

2.2. Protein Structure

At the protein level, human PRMT1 shares a high degree of homology with the different members of the PRMT family that is conserved in eukaryotes. Phylogenetic studies based on the methyltransferase domain highlighted that PRMT1 is closely related to PRMT8 [7]. The canonical structure of PRMT1 includes three functional domains: (i) the N-terminal methyltransferase domain characterized by the Rossmann fold constituting the AdoMet binding pocket, (ii) the C-terminal β-barrel domain which forms a cylindrical structure corresponding to the arginine-substrate binding site and (iii) the α-helical dimerization arm which originates from the N-terminal part of β-barrel domain and connects to the Rossmann fold of a second monomer [8].
The catalytic core of PRMT1 is composed of 6 highly conserved peptide motifs essential for the methyltransferase activity. Motif I (VLDVGSGTG) delimits the AdoMet-binding site and is stabilized by motifs II (VDI) and III (LAPDG). The binding of the AdoMet in this pocket is favored by the formation of hydrogen bonds with the glutamic acid residue of the post-motif I (VIGIE). In addition, the double-E motif (SEWMGYCLFYESM) and the THW loop (YTHWK) define the peptidyl arginine-substrate pocket (Figure 1C). The double-E motif is composed of two negatively charged glutamic acid residues (E144 and E153) that neutralize the positively charged guanidium group of the target arginine, whereas the THW loop stabilizes three dynamic α-helices (αX, αY, αZ) located at the N-terminus of the Rossmann fold that participates in peptidyl arginine recognition [9,10,11]. To illustrate the organization of the catalytic core of PRMT1, an extensive study of the crystal structure of rat PRMT1 which shares 96% identity with the amino acid sequence of human PRMT1 was performed by Zhang and Cheng [10].
Dimerization of PRMTs is a conserved process, crucial for their methyltransferase activity. This mechanism is mediated by the dimerization arm that interacts with the outer surface of the AdoMet binding site through hydrophobic interactions and hydrogen bonds [12]. PRMT1 mutants displaying a mutation or a deletion of the dimerization arm were key to demonstrating the importance of dimerization for AdoMet binding, substrate specificity and the processivity of the methyltransferase activity [10]. As previously described for the yeast PRMT1 counterpart, Hmt1, rat PRMT1 dimers can be assembled into oligomers through hydrophilic interactions [13,14]. This oligomerization is notably associated with a stimulation of the PRMT1 methyltransferase activity [14].

2.3. PRMT1 Isoforms

To date, seven PRMT1 isoforms, PRMT1-v1 to PRMT1-v7, that differ in length and sequence of their N-terminal region have been identified (Figure 1B). These variations of the N-terminal sequence can impact enzymatic activity and substrate specificity. Unlike PRMT1-v7 which is catalytically inactive, variants PRMT1-v1 to PRMT1-v6 exhibit a methyltransferase activity in vitro on different previously described PRMT1 substrates. However, PRMT1-v3 and PRMT1-v4 display a lower methylation efficiency compared to the others. Studies of Goulet et al. also showed that each substrate can be preferentially methylated by a particular isoform. For example, Sam68 and SmB are mainly methylated by PRMT1-v1 and PRMT1-v2 [5]. Currently, studies describing the functionality of the PRMT1-v7 variant are lacking. Although it has retained the ability to heterodimerize with other isoforms, it does not seem to be involved in the regulation of their activity [5].
Differences in enzymatic activities observed among the different PRMT1 isoforms can be partly explained by their subcellular localization. Using a GFP-PRMT1 isoform reporter system, Goulet et al. showed that PRMT1-v1 and -v7 are mainly nuclear, whereas PRMT1-v2 is primarily cytoplasmic [5]. The nucleocytoplasmic shuttling of PRMT1-v2 depends on a leucin-rich nuclear export sequence (NES) encoded by the retained exon 2, but also on its enzymatic activity [15]. Interestingly, there is also a tissue-specific expression pattern of the different PRMT1 isoforms. PRMT1-v1, -v2 and -v3 are ubiquitously expressed in human tissues [4], whereas PRMT1-v4 to -v7 are tissue-specific. More precisely, expression of PRMT1-v4 and -v5 is restricted to the heart and pancreas, respectively; yet, PRMT1-v7 is detectable in the heart and skeletal muscle. PRMT1-v6 expression has so far not been detected in any normal human tissues but was detected in certain breast cancer cell lines [5].

3. Biochemical Features

3.1. Sequence Specificity

PRMT1, like the other type I PRMTs, except PRMT4, catalyzes the asymmetric dimethylation of arginine residues localized in glycine/arginine rich regions and more particularly within RGG or RXR motifs [10]. “RGG” sequences that are often located in regions rich in “RG” dinucleotides are also described as “RGG/RG” motifs that can be subdivided into 4 categories according to the number of repeats: “Tri-RGG”, “Di-RGG”, “Tri-RG” or “Di-RG” motifs [16]. Many substrates of PRMT1 contain a combination of these different motifs such as TAF15 (3 Tri-RGG, 1 Di-RGG) or Sam68 (1 Di-RGG, 1 Tri-RG, 1 Di-RG). Structurally, the presence of glycine residues near the target arginine induces a conformational flexibility that facilitates substrate recognition [17].
The modification of a single residue in conserved motifs like “RGG” can abolish the activity of PRMT1 towards the mutated substrate. For instance, the helicase eIF4A1 that contains an “RGG” motif is methylated by PRMT1, whereas the eIF4A3 isoform in which “RGG” is replaced by an “RSG” sequence is not a substrate for PRMT1. However, it was shown in the same study that PRMT1 is able to methylate synthetic peptides that contain a “RSG” sequence [18]. This suggests that other residues located at a long distance from the target arginine can also be involved in its recognition. This hypothesis was substantiated by a study of Osborne et al., which showed that the affinity of PRMT1 for its arginine substrate relies on long-range interactions involving an acidic residue located away from the PRMT1 active site and probably a positively-charged residue on the substrate [19].

3.2. Product Specificity

Understanding mechanisms that regulate the degree (mono- or dimethylation) and the type (symmetric or asymmetric) of methylation catalyzed by each member of the PRMT family is a major challenge. Indeed, MMA, ADMA and SDMA induce distinct and sometimes antagonistic biological effects as notably described for mono- and dimethylated H3R2 [20,21].
Studies conducted by Gui et al. on rat PRMT1 that shares 96% sequence identity with its human counterpart, identified two conserved methionine residues, M48 and M155, located in the active site that position the target arginine in a favorable configuration for asymmetric dimethylation. Interestingly, M48 also participates in the specific recognition of the target arginine in multi-arginine protein substrates [20]. Mutations in M48L and M155A induce an imbalance in the proportion of MMAs and ADMAs, but do not allow SDMA generation [20]. However, when M48 is mutated to phenylalanine (M48F), a switch in PRMT1 activity occurs, enabling it to induce symmetrical dimethylation. This is consistent with the fact that product specificity of PRMT5 which catalyzes SDMA formation is controlled by the conserved F379 residue in its active site [22]. More recently, mutagenesis studies showed that H293S mutation of the PRMT1 active site does not affect the production of MMA and ADMA by itself, but leads to a predominant formation of SDMA when it is associated with the M48F mutation [23].
The product specificity of PRMT1 which is non-stochastic and regioselective can also be guided by the substrate itself. It seems that the N-terminal arginyl-groups of substrates constitute the main targets for PRMT1 methylation, whereas positively-charged C-terminal residues (including arginines) participate in long-range interactions with acidic residues of PRMT1. This strengthens the affinity of PRMT1 for its arginine substrates [19,24].
Interestingly, the amino acid sequence of the substrate can also direct the degree of methylation (mono- or dimethylation) by regulating PRMT1 processivity [24,25]. Whether PRMT1 dimethylates its substrates in a distributive or processive manner is a matter of debate in the literature. While numerous studies support that PRMT1 acts distributively by transiently releasing MMA and replacing the methyl donor between the two methyl-group transfers [26,27,28], Obianyo and co-workers described a semi-processive activity of PRMT1. In this model the mono-methylated intermediate remained associated with the enzyme but the product S-adenosylhomocysteine (AdoHcy) was replaced by a novel AdoMet to allow the second reaction [19,29,30]. Studies on the catalytic activity and processivity of PRMT1 are ongoing, and the latest data indicate that the degree of processivity of PRMT1 depends on its dimerization but is also dependent on cofactor or enzyme concentrations [10,25].

3.3. Regulation of PRMT1 Expression and Enzymatic Activity

Many studies have sought to decipher the different levels of regulation of PRMT1 expression and enzymatic activity. Indeed, substrate methylation by PRMT1 is a highly regulated and dynamic phenomenon, occurring directly through PRMT1 PTMs or through its association with co-regulators. In addition, crosstalk between different PTMs on the same substrate can influence arginine methylation by PRMT1. Finally, methyl marks on arginine can be removed by PAD4 which demethylates histones by converting MMA to citrulline [31] or by JMJD6 which directly removes the methyl group to convert methylarginine into arginine [32]. More recently, JMJD1B, a well-known lysine demethylase for H3K9me2, has also been described as effective in demethylating H4R3me1 and H4R3me2a [33].

3.3.1. Regulation of PRMT1 Expression

PRMT1 can be regulated at the level of its expression. Indeed, a very recent study discovered that the serine/threonine kinase mTOR is involved in the regulation of PRMT1 expression in a fasting context. Forty-eight hours of experimental fasting was shown to induce a decrease in STAT1 phosphorylation mediated by mTOR, leading to the inhibition of STAT1 binding to the PRMT1 promoter. In this fasting condition, the decrease in PRMT1 expression induced a decrease in mitochondrial mass and thus a decrease in cellular energy availability [34]. Moreover, the expression level of PRMT1 can also be regulated by micro-RNAs (miR). This is the case for example for miR-503 that has a tumor suppressor role and whose expression is low in several types of cancers. In hepatocellular carcinoma cells, miR-503 directly targets PRMT1 and reduces its expression level. Consequently a decrease in cell invasion, migration and epithelial-mesenchymal transition are observed [35].

3.3.2. Post-Translational Modification of PRMT1

Unlike other PRMTs, few PTMs of PRMT1 have been described to date. A first study in 2004 conducted using mass spectrometry found that PRMT1 is phosphorylated on Y291. Using non-natural amino acid mutagenesis, the authors showed that phosphorylation of PRMT1 on Y291 alters protein-protein interactions and substrate specificity. Indeed, Y291 phosphorylation of PRMT1 decreases its interaction with hnRNP, and enzymatic activity on hnRNP in vitro. This is due to the negative charge of the phosphate group that modifies the tertiary structure of the enzyme and in particular of the THW loop [36]. Following this first finding, another study in keratinocytes revealed that PRMT1 is a substrate of the kinase CSNK1a1. Although phosphorylation of PRMT1 by CSNK1a1 does not affect the methylation efficiency of PRMT1 on several known substrates, it seems that it modulates its transcriptional activity on some target genes. Indeed, phosphorylated PRMT1 seems to induce the transcription of genes involved in proliferation and repress the expression of genes involved in keratinocyte differentiation [37]. More recently, in ovarian cancer cells, it was shown that PRMT1 can be phosphorylated by DNA-PK in response to cisplatin, thus inducing its recruitment on chromatin and its enzymatic activity towards H4R3 [38].
PRMT1 activity is also modulated by its degradation mediated by the proteasome pathway. In this context, a study in human embryonic kidney cells showed that PRMT1 is polyubiquitinated by the E3 ubiquitin ligase, TRIM48. Thus, the polyubiquitination of PRMT1 decreases the level of methylation of the substrate ASK1, a kinase involved in the cellular stress response. Downregulation of PRMT1 thus promotes cell death induced by ASK1-mediated oxidative stress. Polyubiquination of PRMT1 also negatively impacts FOXO1 methylation and its transcriptional activity [39]. Another in vivo study used an engineered ubiquitin transfer method called “orthogonal UB transfer” to profile E3 substrate specificity. This method showed that PRMT1 is polyubiquitinated by two other E3 ubiquitin ligases, CHIP and E4B, leading to its proteasome-mediated degradation. Nevertheless, the physiological consequences of this polyubiquitination were not investigated in this study [40]. Given the importance of PRMT1, it probably undergoes many other PTMs including methylation, such as PRMT5 which is methylated by PRMT4 [41], or OGT-glycosylation [42]. Although other modifications (i.e., acetylation and sumoylation) have not been described in the PRMT family, it is likely that these events exist.

3.3.3. PRMT1 Association with Co-Regulators

PRMT1 activity can also be regulated through its interaction with non-substrate proteins that modulate its methyltransferase activity. The first regulators were described in 1996, with the BTG1 (B-cell translocation gene 1) and BTG2. This study showed in vitro that the interaction of BTG1 and BTG2 with PRMT1 positively modulates its enzymatic activity towards a substrate, hnRNPA1 [43]. Several years later, our team discovered a new regulator of PRMT1, hCAF1. We showed by in vitro methylation assay that hCAF1 inhibits PRMT1-mediated methylation of histone H4 on arginine 3 (H4R3) by PRMT1. This observation was confirmed in breast cancer cells where depletion of hCAF1 induces a strong reduction in the overall level of asymmetric arginine methylation, indicating that hCAF1 modulates PRMT1 activity towards several substrates [44]. Interestingly, a study in HeLa cells revealed a crosstalk between PRMT1 and PRMT2. Indeed, PRMT2 binds to PRMT1 without methylating it and potentiates its enzymatic activity towards H4R3. Surprisingly, PRMT2-mediated activation of PRMT1 also induces an increase in SDMA levels in vivo, implying possible further crosstalk between the different enzymes of the PRMT family [45].
PRMT1 activity can also be modulated by exogenous regulators. For instance, the serine/threonine phosphatase PP2A has been described to regulate PRMT1 activity. PRMT1 methylate hepatitis C virus NS3 protein and inhibits its helicase activity. PP2A binds to PRMT1 and inhibits its enzymatic activity towards a NS3 protein, which affects inhibitory role of PRMT1 on the helicase activity of NS3. Interestingly, the hepatitis C virus upregulates PP2A expression, thus counteracting the downregulation of NS3 by PRMT1. This study highlights the complexity of the pathways regulating PRMT1 enzymatic activity [46].
In addition, other regulators have been identified, such as RALY [47], TR3 [48], PDGF-BB [49], or GFI1 [50]. Moreover, other mechanisms of regulation of PRMT1 have been uncovered, such as oxidative stress [12] or iron deficiency [51].

3.3.4. PTMs Influencing PRMT1 Activity

In parallel to the direct regulation of PRMT1 by PTMs or by the binding of co-regulators, a crosstalk between arginine methylation and different PTMs deposited by other enzymes on the same substrate has been described. For example, a 2006 study showed that methylation of H4R3 by PRMT1 at the pS2 promoter is required to activate its expression. Interestingly, this study showed that histone hypoacetylation is necessary for the recruitment of PRMT1 to the promoter and for the deposition of the H4R3 methylation mark. The patient SE translocation (SET) protein, which is part of the INHAT complex, prevents the acetylation of the histone at the pS2 promoter [52]. Another study investigated the effect of histone H4 phosphorylation on serine 1 (H4S1). The authors showed by in vitro methylation assays that H4S1 phosphorylation leads to a 3-fold decrease in PRMT1-mediated H4R3 methylation. Interestingly, mass spectrometry analysis revealed MMA as a PRMT1 major product. Indeed, further in vitro methylation assays revealed a 3-fold decrease in ADMA, due to an approximate 11-fold reduction in PRMT1 catalytic efficiency. Moreover, H4S1 phosphorylation also leads to a 8-, 5-, and 3-fold decrease in PRMT3, PRMT8 and PRMT5 activity, respectively [53].
These in vitro studies highlighted the complex crosstalk between the different PTMs in the histone code and the tight regulation of the activity of each enzyme. Although this phenomenon has only been described on H4R3 for PRMT1, this is probably because PRMT1 was first described as a histone methyltransferase catalyzing H4R3 methylation [54]. Many non-histone substrates have since been described, and likely display similar crosstalk that remains to be depicted.

3.4. Substrates

Arginine 3 of histone H4 was the first substrate described for PRMT1 [54,55]. The asymmetric dimethylation of H4R3 constitutes an activating mark of transcription [56]. It was also demonstrated that PRMT1 methylates histone H2A at R3, R11 and R29, although the latter two are not localized within a consensus motif recognized by PRMT1 [57]. Further studies are expected to clarify the impact of these two histone marks on transcriptional activity. In addition to the activity of PRMT1 as a chromatin modifying enzyme, a plethora of non-histone substrates of PRMT1 have been identified and can be classified according to their cellular functions: transcriptional and translational regulation, RNA-processing, DNA damage repair and signal transduction. A list of the currently identified substrates of PRMT1 is available in Table 1.
It is important to note that some substrates are common to different types of PRMTs and that competitive mechanisms may exist. This hypothesis is supported by the observations of Dhar et al. who showed that inhibition of PRMT1 induces a decrease in the level of ADMA concomitant with an increase in MMA and SDMA levels [58].
Table 1. List of non-histone substrates of PRMT1 classified according to their cellular functions.
Table 1. List of non-histone substrates of PRMT1 classified according to their cellular functions.
Biological FunctionSubstrateMethylation SiteBiological OutcomeReference
Transcriptional
Regulation
Transcriptional
regulation
BRCA1Within the 504–802 regionPromotes BRCA1 recruitment to specific promoters[59]
C/EBPαR35, R156, R165Prevents C/EBPα interaction with the corepressor HDAC3[60]
c-MycR299, R346Promotes c-Myc interaction with p300[61]
EZH2R342Prevents EZH2 target gene expression[62]
FOXO1R248, R250Prevents FOXO1 phosphorylation by Akt[63]
FOXP3R48, R51Enhances FOXP3 transcriptional activity[64]
GLI1R597Enhances GLI1 binding to target gene promoters[65]
MyoDR121Promotes MyoD DNA-binding and transcriptional activity[66]
Nrf2R437Promotes Nrf2 DNA-binding and transcriptional activity[67]
PRR637Accelerates PR recycling and transcriptional activity[68]
RACO-1R98, R109Promotes c-Jun/AP1 activation[69]
RelAR30Prevents RelA DNA-binding and represses NF-κB target genes[70]
RIP40R240, R650, R948Favors RIP140 nuclear export and prevents the recruitment of HDAC3[71]
RunX1R206, R210Prevents Sin3a binding and promotes RUNX1 transcriptional activity[72]
STAT1R31Prevents STAT1 association with PIAS1 and enhances IFNα/β induced transcription[73]
TAF15R203Affects the subcellular localization of TAF-15 and enhances its transcriptional activity[74]
FUS/TLSR216, R218, R242, R394Participates in the nuclear cytoplasmic shuttling of FUS/TLS and enhances its transcriptional activity[75,76]
TOP3BR833, R835Promotes TOP3B interaction with TDRD3, stress granule localization and topoisomerase activity[77]
Twist1R34Regulates the nuclear import of Twist1 and represses E-cadherin expression[78]
RNA- processingCNBPR25, R27Prevents its RNA binding activity[79]
G3BP1R435, R447Prevents stress granule formation during oxidative stress[80]
hnRNPA1R214, R226, R223, R240Prevents hnRNPA1 ITAF activity and RNA-binding ability[81]
HSP70R416, R447Enhances HSP70 RNA-binding and -stabilization abilities[82]
NS3R1493Affects NS3 RNA-binding and helicase activity[46,83]
RBM15R578Promotes RBM15 degradation by CNOT4 (RNA splicing)[84]
Sam68Within the 276–343 regionPrevents Sam68 poly(U) RNA-binding activity[85,86]
SF2/ASFR93, R97, R109Affects SF2/ASF nucleocytoplasmic distribution and modulates the alternative splicing of target genes[87,88]
Translational
Regulation
eIF4A1R362Prevents eIF4A1 interaction with eIF4G1 and inhibits ATPase activity[18,89]
eIF4G1R689, R698Regulates eIF4G1 stability and the assembly of the translation initiation complex[90]
rpS3R64, R65, R67Promotes rpS3 import into the nucleolus and ribosome assembly[91]

DNA damage
repair
53BP1Within the 1319–1480 regionPromotes 53BP1 recruitment to DNA-damage sites[92]
APE1R301Promotes APE1 mitochondrial translocation (translocase Tom20) and protects mitochondrial DNA from oxidative damage[93]
DNA pol βR137Prevents DNA pol β interaction with PCNA in BER pathway[94]
E2F-1R109Promotes E2F-1-dependent apoptosis in DNA-damaged cells[95]
FEN1Not determinedStabilizes FEN1 and upregulates its DNA damage repair activities[96]
hnRNPKR296, R299Prevents PKCδ-dependent apoptosis during DNA damage[97]
hnRNPUL1R584, R618, R620, R645, R656Promotes hnRNPUL1 association with NBS1 and recruitment to DNA-damage sites[98]
MRE11GAR domainPromotes MRE11 recruitment to DNA-damage sites and favors its exonuclease activity[99,100]
RunX1R233, R237Confers resistance to apoptosis under stress condition and DNA damage accumulation[101]
ASK1R78, R80Prevents the stress-induced ASK1-JNK1 signaling[102]
Signal transductionAxinR378Favors Axin stability and consequently prevents Wnt/β-catenin signaling[103]
BADR94, R96Prevents BAD phosphorylation by Akt and subsequent survival signaling[104]
CaMKIIR9, R275Prevents CaMKII-dependent signaling in cardiomyocytes[105]
CDK4R55, R73, R82, R163Prevents the formation of a CDK4/Cyc D3 complex and subsequent cell cycle progression[106]
cTnIR146, R148Induces cardiac cell hypertrophy[107]
EGFRR198, R200Upregulates EGFR signaling[108]
ERαR260Promotes the formation of the ERα/PI3K/Src/FAK complex and subsequent activation of downstream kinase cascades[109]
INCENPR887Enhances INCENP binding-affinity to AURKB and promotes cell division[110]
KCNQR333, R345, R353, R435Promotes PIP2 binding and subsequent KCNQ channel activity[111]
MYCNR65Enhances MYCN stability through CDK-dependent phosphorylation[112]
NONOR251Favors NONO oncogenic function[113]
p38 MAPKR49, R149Promotes p38 MAPK phosphorylation by MKK3 and the subsequent activation of MAPKAK2 involved in erythroid differentiation[114]
Smad4R272Promotes Smad4 phosphorylation by GSK3 and support the activation of the canonical Wnt signaling[115]
Smad6R74, R81Participates in BMP signaling and prevents NF-κB activation[116,117]
Smad7R57, R67Facilitates TGF-β signaling[118]
TRAF6R88, R125Prevents TRAF6 ubiquitin ligase activity and regulates Toll-like receptor signaling[119]
TSC2R1457, R1459Blocks the Akt-dependent phosphorylation of TSC2 and regulates mTORC1 activity[120]

4. Cellular Features

4.1. Connection with Chromatin Dynamics and Transcriptional Regulation

Arginine methylation was first described as a PTM of histones that regulates reader protein recruitment and therefore chromatin dynamics. The main target of PRMT1 at the chromatin level is the arginine 3 of histone H4 (H4R3) [54,55]. Asymmetrically dimethylated H4R3, H4R3me2a, is associated with an active form of the chromatin and recognized by different Tudor domain-containing proteins, such as TDRD3 [121]. This protein, with no intrinsic activity, serves as a scaffold coregulator for the assembly of protein complexes at the transcription start sites of target genes. More precisely, TDRD3 can recruit, through its OB-fold domain, the DNA Topoisomerase IIIβ [122] and can directly interact with the RNA Polymerase II, previously methylated at R1810 by PRMT4 also known as CARM1 [123]. Therefore, this complex assembled through TDRD3 and likely involving other actors promotes transcription at H4R3me2a loci (Figure 2).
Interestingly, H4R3me2a can also recruit chromatin modifying enzymes involved in transcriptional regulation by depositing other histone marks on chromatin. Indeed, methylation of H4R3 by PRMT1 promotes the subsequent acetylation of H4K8 and H4K12 by the histone acetyltransferase p300 [56]. An H4R3me2a-dependent induction of H4K5 and H4K12 acetylation, allowing the recruitment of the transcription initiation factor TAFII250 and therefore contributing to chromatin opening, was also suggested using the chicken β-globin locus as a model [124]. Finally, the ability of H3R4me2a to act in trans to promote the acetylation of histone H3K9 and H3K14 by the histone acetyltransferases p300 and PCAF was demonstrated within the β-major globin promoter in murine erythroleukemia cells [124,125]. It is worth noting that PCAF directly interacts with H4R3me2a and this could explain how PRMT1-dependent methylation potentiates H3K9 and H3K14 acetylation [125] (Figure 2).
Conversely, the activity of PRMT1 on H4R3 is inhibited by the presence of acetylation, propionylation, crotonylation, butyrylation or 2-hydroxyisobutyrylation of H4K5 [126]. Moreover, H4K5ac combined with H4K8ac or H4K12ac increases its repressive effect on PRMT1 activity. There is currently one known exception, as acetylated H4K16 is associated with an increase in PRMT1 activity. Interestingly, the inducing effect of H4K16ac dominates the repressive effect of H4K5ac when the 2 histone marks co-exist [53,127] (Figure 2).
Aside from chromatin regulation, a large number of transcription factors whose activity can be regulated by PTMs are known PRMT1 substrates (Table 1). PRMT1-dependent methylation can notably increase their stability and thus promote their transactivation function. This type of mechanism has been described for FOXO1 whose methylation by PRMT1 prevents its proteosomal degradation and favors its nuclear localization [63]. The methyltransferase activity of PRMT1 can also impact interactions between transcription factors and their corepressors. For example, PRMT1 was shown to act as a coactivator of RUNX1 by inducing its methylation at R206 and R210, and thereby preventing its interaction with the transcriptional corepressor SIN3A [72]. Similarly, C/EBPα methylation at R35, R156 and R165 blocks its interaction with the corepressor HDAC3 [60].

4.2. Connection to Cell Signaling Pathways

4.2.1. Steroid Receptors

To date, PRMT1 has been shown to methylate two steroid receptors; estrogen receptor (ERα) and progesterone receptor (PR). These arginine methylation events control different signaling pathways involved in breast tumorigenesis.

Estrogen Receptor (ERα)

ERα regulates many physiological processes, notably the growth and survival of breast tumor cells, acting as a ligand-dependent transcription factor. Aside from the well described transcriptional effects, estrogen also mediates extranuclear events called non-genomic signaling via its receptor [128]. Our group showed that ERα is methylated on the residue R260 (met260ERα) by PRMT1 in response to estrogen or IGF-1 [109,129]. This event is a prerequisite for the formation of a signaling complex containing met260ERα, Src and PI3K, which orchestrates cell proliferation and survival. The involvement of this complex in breast carcinogenesis will be addressed in Section 5.1 of this review. Met260ERα is a transient event downregulated by the arginine demethylase JMJD6 [130].

Progesterone Receptor (PR)

Our group also demonstrated that PRMT1 methylates PR on the residue R637, within a RGG consensus site. This methylation event decreases PR stability in order to accelerate its recycling and its transcriptional activity. In addition, PRMT1 depletion decreases the expression of a specific subset of progesterone-target genes, involved in breast cancer cell proliferation and migration [68].

4.2.2. Akt Signaling Pathway

Several reports demonstrated that specific arginine methylation, catalyzed by PRMT1 within the Akt consensus phosphorylation motif, works as an inhibitor of Akt-dependent survival signaling.

FOXO

Forkhead box O (FOXO) is a family of transcription factors controlling a large variety of biological processes including cell survival [131]. Several studies revealed that FOXO proteins are phosphorylated by Akt, resulting (i) in the export of FOXO proteins from the nucleus to the cytoplasm [132,133] and (ii) in FOXO proteasomal degradation through polyubiquitination [134,135]. Interestingly, a member of the FOXO family, FOXO1 was shown to be methylated by PRMT1 on R248 and R250, in the consensus Akt phosphorylation site, impeding Akt phosphorylation on S253 [63]. This methylation event results in a decrease in its cytoplasmic localization and its subsequent degradation. PRMT1 depletion decreases oxidative-stress-induced apoptosis regulated by the Akt-FOXO1 pathway. These results indicated that PRMT1 arginine methylation can act as a modulator of Akt-phosphorylation by regulating responses to oxidative stress in mammalian cells.

BAD

Similarly, PRMT1 binds and methylates the proapoptotic protein BCL-2 antagonist of cell death (BAD) on R94 and R96, in the Akt consensus site. PRMT1 methylation on these two residues inhibits Akt phosphorylation on S99, a modification that is necessary for its interaction and sequestration with 14-3-3 proteins, resulting in cell survival [104].

4.2.3. NF-κB Signaling

NF-κB plays an important role in the transcriptional regulation of genes involved in inflammation and cell survival. Toll-like receptor (TLR), when activated by lipopolysaccharides, triggers the recruitment of the adaptor protein Myd88 and the subsequent activation of the transcription factor NF-κB. TGFβ inhibits TLR signaling through the methylation of SMAD6 by PRMT1. Indeed, the binding of methylated SMAD6 to Myd88 results in its degradation, impeding TLR signaling to NF-κB [117]. Moreover, PRMT1 serves as a coactivator of NF-κB, synergistically with CARM1, although the underlying mechanisms are not fully elucidated [136]. More recently, the methylation of the RelA subunit of NF-κB by PRMT1 was identified as a repressive mark modulating TNFα/NF-κB response [70].

4.2.4. Wnt Signaling

Wnt signaling plays important roles in embryonic development and cell proliferation. Aberrant Wnt signaling leads to several human diseases including cancer. Axin is a negative regulator of the Wnt pathway, as it is a key scaffold protein for the β-catenin destruction complex. PRMT1-induced methylation of axin enhances its interaction with GSK3β, leading to a decrease in axin ubiquitination and degradation [103]. Therefore, PRMT1 seems to be a new modulator of Wnt/β-catenin signaling. Moreover, PRMT1 also regulates this pathway by methylating substrates prior to their phosphorylation by GSK3β and its sequestration in endolysosomes, a key event in Wnt signaling [115]. Altogether PRMT1 appears as an important modulator of the Wnt pathway at the interface of protein phosphorylation and trafficking.

4.3. Cellular Role and Functions

4.3.1. Embryogenesis and Development

The critical role of PRMT1 in embryogenesis and development was first suggested by the study of Pawlak et al. which showed that PRMT1 knockout mouse embryos, generated by insertion of a gene trap retrovirus in the second intron of the PRMT1 gene, failed to develop beyond embryonic day 6.5, which would coincide with the exhaustion of the maternal stock of PRMT1 enzymes and methylated substrates [137]. It is worth noting that homozygous PRMT1 mutant embryonic stem (ES) cells isolated from mutant preimplantation blastocysts at day 3.5 are viable and retained the morphology and the same doubling time as wild-type ES cells. Moreover, in these cells, loss of PRMT1 activity is not balanced by the activation of other methyltransferases. Therefore, PRMT1 activity does not seem to be required for cell viability [137].
Early lethality of homozygous PRMT1 KO mouse embryos, as well as their uterus-enclosed localization, makes it difficult to study the epigenetic regulation of vertebrate development and emphasizes the importance to develop other models. Among them, Zebra fish embryos constitute a promising model as they are suitable for genetic manipulation approaches and express a highly conserved PRMT1 protein (90% identity with human PRMT1) at different stages of embryogenesis. A study conducted by Tsai et al. showed that PRMT1 knockdown, by antisense morpholino oligo injection into one-cell stage zebra fish embryos, induces developmental defects at gastrulation notably including a shortened body-length. This highlighted the importance of the methyltransferase activity of PRMT1 in early embryogenesis [138]. More recently, Shibata et al. used the TALEN genome editing technology to knockout PRMT1 in the diploid anuran Xenopus tropicalis that undergoes an external and biphasic development (embryogenesis and metamorphosis). They observed that H4R3me2a methylation by PRMT1 is not required for early embryogenesis but is essential for the growth and development of various organs including the brain, liver and intestine during late embryonic developmental stages, occurring prior to metamorphosis. This effect is directly related to the drastic inhibition of cell proliferation associated with PRMT1 KO in this model [139]. Interestingly, Xenopus embryos were already used to demonstrate the involvement of the xPRMT1b gene in early neural determination [140].
Another interesting aspect is the potential involvement of PRMT1 in placental development. A study of Sato et al. showed that murine placental expression of two PRMT1 isoforms is differentially regulated during the gestational period. More precisely, while PRMT1-v1 expression reaches a maximum at embryonic day E11 before decreasing, PRMT1-v2 expression increases from E13. This balance between the two isoforms explains the change in subcellular localization of PRMT1 observed between early and late stages of gestation; though further studies are required to determine the exact role played by PRMT1 in the placenta [141].

4.3.2. DNA Damage Repair

The conditional knockout of PRMT1 in mouse embryonic fibroblasts is associated with a severe genetic instability characterized by the occurrence of spontaneous DNA damage, chromosome copy number variations and defective mitotic checkpoint [142]. The relevance of PRMT1 in the maintenance of genome integrity is based on the methylation and subsequent regulation of key factors involved in the major DNA repair pathways.
The first substrate of PRMT1, involved in DNA damage repair, to be identified was MRE11 (Meiotic recombination 11). This component of the MRN complex (MRE11/RAD50/NBS1), recruited early upon DNA double-strand break (DSB), participates in the initiation of DNA repair pathways by homologous recombination (HR) or by non-homologous end joining (NHEJ). Methylation of the C-terminal GAR motif of MRE11 at R587 by PRMT1 does not seem to participate in the formation of the MRN complex but it promotes the relocalization of MRE11 from PML nuclear bodies to DNA-damage sites and it favors its exonuclease activity [92,99,100]. These events are essential to allow the recruitment of RAD51 and the subsequent activation of HR [100]. By using a model of knock-in mice that express the mutated MRE11RK protein devoid of methylarginines, Yu et al. also demonstrated that MRE11 methylation participates in the activation of the ATR/CHK1 checkpoint signaling [143]. Finally, methylated MRE11 is involved in telomere maintenance and regulates DNA replication by controlling the intra-S phase checkpoint in response to DNA damage [99,144].
The choice of pathways between NHEJ or HR is directly influenced by the DNA-end structure of DNA DSBs. Among the actors that play a pivotal role to orient this choice are the tumor suppressor protein BRCA1, which promotes HR repair by activating DNA-end resection, and p53-Binding Protein 1 (53BP1) that inversely activates NHEJ repair by inhibiting the recruitment of BRCA1 to DNA DSBs [145]. Interestingly, these two proteins are methylated by PRMT1, suggesting that arginine methylation may play an important role in directing the switch from HR to NHEJ repair pathways. More precisely, 53BP1 is methylated by PRMT1 at a canonical GAR motif localized in its kinetochore-binding domain and this methylation is essential for its DNA-binding activities [92,146]. Concerning BRCA1, the methylation status of the 504–802 protein region, that encompasses the DNA-binding domain, directly influences its interaction with transcription factors such as Sp1 or STAT1 and its subsequent recruitment to specific promoters [59].
The base excision repair mechanism (BER) that can correct single-stranded DNA breaks and oxidative or alkylation damage is also regulated by PRMT1, which methylates two major players in this pathway, namely the Flap endonuclease 1 (FEN1) and the DNA polymerase β (DNA Pol β). Methylation of FEN1 by PRMT1, at an arginine residue that remains to be determined, stabilizes the protein without disturbing its localization [96]. Moreover, unlike PRMT5-dependent methylation at residue R192 which strengthens the interaction between FEN1 and the DNA polymerase processivity factor PCNA necessary for a faithful and efficient BER, PRMT1-dependent methylation of FEN1 does not seem to impact this interaction [96,147]. Interestingly, methylation of the DNA Pol β by PRMT1 on R137 abolishes its binding with PCNA without affecting its enzymatic activities (polymerase and dRA-lyase) [94]. This suggests that methylation could regulate the sequential interaction of FEN1 and DNA Pol β with PCNA during BER.

5. PRMT1 in Cancer

Since the substrates methylated by most PRMTs regulate various biological functions essential for cellular homeostasis, it is not surprising that a dysregulation of arginine methylation may contribute to cancer initiation and progression. The involvement of PRMT1 in carcinogenesis is no longer questioned due to its overexpression or aberrant splicing observed in numerous types of cancers.

5.1. Breast Cancer

Various studies have shown that PRMT1 gene expression is higher in breast tumor samples than in healthy tissue suggesting the involvement of PRMT1 in breast carcinogenesis [5,148]. Despite the detection of PRMT1-v1, v2 and v3 isoforms in breast tumor tissue, it seems that only the predominant PRMT1-v1 variant is correlated with clinical parameters such as histological grade [148].
ERα is an important PRMT1 substrate whose methylation can be associated with the development of breast cancer. Our group highlighted that a PRMT1-dependent hypermethylation of ERα at R260, induced in response to estrogen or IGF-1, is observed in different subtypes of breast cancers and regulates cell proliferation and survival [109,129]. We notably showed that the signaling complex containing met260ERα, Src and PI3K (described in Section 4.2.1 of this review) is expressed at low levels in the cytoplasm of normal mammary epithelial cells but highly expressed in 55% of breast tumors [149]. Moreover, its overexpression is correlated with the activation of Akt (pAkt), the main effector of the pathway, showing that this signaling pathway exists in vivo. In addition, a high expression of the complex is an independent marker of poor prognostic [149] and has been linked with resistance to tamoxifen [150,151].
Another interesting aspect is the key role of PRMT1 in the maintenance of stem-cell-like properties of breast cancer cells. PRMT1-dependent EGFR methylation on R198 and R200 upregulates different signaling cascades, notably those involving Akt, ERK or STAT3 in triple-negative breast cancer (TNBC) cells, MDA-MB-468. EGFR/ERK-dependent activation of ZEB1, a transcription factor that regulates epithelial-mesenchymal transition, may be implicated in cancer stem cell maintenance [152]. Interestingly, asymmetric dimethylation of H4R3 by PRMT1 at the ZEB1 promoter is another mechanism described to activate this factor and therefore promotes migration, invasion and acquisition of stem cell characteristics. It is worth noting that ZEB1 may simultaneously contribute to the PRMT1-dependent inhibition of senescence in breast cancer cells [153].
PRMT1-dependent methylation also inhibits the tumor suppressive function of some substrates. For example, methylation of C/EBPα at R35, R156 and R165 by PRMT1 prevents its interaction with the corepressor HDAC3, thus promoting the expression of cell-cycle genes such as cyclin D1 and the subsequent growth of breast cancer cells [60]. In the same line, BRCA1 methylation by PRMT1 affects its recruitment to responsive promoters but also its ability to interact with certain partners such as Sp1 or STAT1. As a result, this can significantly affect the tumor suppressive activity of BRCA1 [59].

5.2. Colorectal Cancer

Two clinical reports demonstrated the unfavorable prognosis associated with PRMT1 expression in colorectal cancer (CRC) patients by discussing the respective involvement of PRMT1-v1 and PRMT-v2 isoforms [154,155]. Mechanistically, it was described that H4R3me2a can recruit SMARCA4, an ATPase subunit of the SWI/SNF complex, to the promoter of certain target genes including EGFR to promote their expression. PRMT1-dependent enhancing of EGFR signaling is associated with a significant increase in the proliferative and migratory abilities of human CRC cells [156]. Moreover, methylation of EGFR at R198 and R200 by PRMT1 leads to an EGF-dependent hyperactivation of EGFR signaling and confers cells with resistance to the anti-EGFR monoclonal antibody, cetuximab. Indeed, in CRC patients, the rate of EGFR methylation is directly correlated with a higher recurrence rate after cetuximab treatment and a poorer overall patient survival [108].
Recently, the non-POU domain-containing octamer-binding protein (NONO), which is overexpressed in CRC tissue, was described as a substrate of PRMT1. Methylation of NONO at R251 is required to promote its oncogenic function including the induction of CRC cell proliferation, migration and invasion [113].

5.3. Lung Cancer

As described for other cancers, PRMT1 expression is significantly increased in lung cancer tissue compared to non-neoplastic ones though very little data are available in the literature to explain its role in lung carcinogenesis [157]. A study by Avasarala et al. highlighted that PRMT1 participates in non-small cell lung cancer progression and metastasis through the methylation of the EMT-associated transcription factor Twist1 at R34. PRMT1-dependent Twist1 methylation is associated with inhibition of E-cadherin expression [78]. Moreover, PRMT1 can methylate the inner centromere protein (INCENP) at R887 to favor its interaction and the subsequent activation of aurora kinase B in A549 non-small cell lung cancer cells. This mechanism regulates the alignment and segregation of chromosomes during cell division to promote the growth of cancer cells [110].

5.4. Other Cancers

Dysregulation of PRMT1 expression has been reported in several other types of cancers, albeit the molecular mechanisms that drive the initiation and progression of these cancers remain incompletely understood. The limited data available in the literature indicate that PRMT1 is particularly dysregulated in bladder cancer, esophageal squamous cell carcinoma, as well as in acute myeloid leukemia [157,158,159]. Interestingly, in ovarian carcinomas, upregulation of PRMT1 expression is associated with an increased methylation of the apoptosis signal-regulated kinase 1 (ASK1), which confers tumor cells with resistance to platinum-based chemotherapeutic agents [160]. Moreover, in prostate cancer, the methylation status of H4R3 is significantly correlated with clinical features, such as tumor grade or the risk of prostate cancer recurrence. This study highlighted the fact that histone modifications can also serve as a prognostic marker [161].

5.5. PRMT1 Inhibitors

In 2004, the symmetrical sulfonated urea salt named arginine methylation inhibitor-1 (AMI-1) was the first PRMT inhibitor characterized [162]. Since then, two substrate competitive inhibitors, MS023 and GSK3368715, that broadly target type I PRMTs (Table 2), were developed and displayed antitumor activities notably on xenograft mouse models of acute myeloid leukemia or breast cancer, respectively [163,164,165]. Promisingly, the GSK3368715 inhibitor is currently undergoing a first-time clinical trial (NCT03666988) for patients with solid tumors and diffuse large B-cell lymphoma. However, high affinity of these inhibitors for other type I PRMTs, renders the identification and characterization of specific PRMT1-dependent effects difficult.
Currently, two PRMT1-specific inhibitors, TC-E-5003 and C7280948, are mentioned in the literature (Table 2). TC-E-5003 displays significant antitumor activity in vitro on breast or lung cancer cell lines and inhibits the growth of xenografted A549 lung cancer cells in mice [166]. Concerning C7280948, a study of Yin et al. showed that it suppresses colorectal cancer cell proliferation, migration and invasion [113]. Additionally, a structure-based virtual screening of different libraries of compounds allowed the identification of several potential PRMT1-specific inhibitors, the properties of which were detailed by Hu et al. [167]. Although these inhibitors are promising, more studies are needed to characterize and consider their clinical potential.

6. Outlook

Over the last twenty years since the discovery of PRMT1, the number of studies conducted on this enzyme has constantly increased. This interest, which persists today, has improved our knowledge on the diversity of its substrates and the numerous biological functions regulated by PRMT1. Its key role in cancer initiation and progression makes PRMT1 an interesting target for the development of new anti-cancer therapeutic strategies. Therefore, the development of inhibitors that target PRMT1 activity is an ongoing challenge that may offer new therapeutic opportunities for various pathologies in the coming years.

Author Contributions

T.C., E.L., P.C. and L.R.M. wrote the manuscript and revised it. All authors have read and agreed to the published version of the manuscript.

Funding

T.C., E.L., P.C. and L.R.M.’s laboratory is funded with grants from “La Ligue contre le Cancer”, “Fondation ARC Cancer” and “Fondation de France”. C.T. was supported from a fellowship from “Fondation de France” and L.E. was supported from a fellowship from “La Ligue contre le Cancer”.

Acknowledgments

We thank B. Manship for proofreading the manuscript. The illustrations were created by using Servier Medical Art (SERVIER SAS, Suresnes, France).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, Y.; Bedford, M.T. Protein arginine methyltransferases and cancer. Nat. Rev. Cancer 2012, 13, 37–50. [Google Scholar] [CrossRef]
  2. Zhang, J.; Jing, L.; Li, M.; He, L.; Guo, Z. Regulation of histone arginine methylation/demethylation by methylase and demethylase. Mol. Med. Rep. 2019, 19, 3963. [Google Scholar] [CrossRef] [Green Version]
  3. Tang, J.; Frankel, A.; Cook, R.; Kim, S.; Paik, W.; Williams, K.; Clarke, S.; Herschman, H. PRMT1 is the predominant type I protein arginine methyltransferase in mammalian cells. J. Biol. Chem. 2000, 275, 7723–7730. [Google Scholar] [CrossRef] [Green Version]
  4. Scorilas, A.; Black, M.H.; Talieri, M.; Diamandis, E.P. Genomic Organization, Physical Mapping, and Expression Analysis of the Human Protein Arginine Methyltransferase 1 Gene. Biochem. Biophys. Res. Commun. 2000, 278, 349–359. [Google Scholar] [CrossRef] [Green Version]
  5. Goulet, I.; Gauvin, G.; Boisvenue, S.; Côté, J. Alternative splicing yields protein arginine methyltransferase 1 isoforms with distinct activity, substrate specificity, and subcellular localization. J. Biol. Chem. 2007, 282, 33009–33021. [Google Scholar] [CrossRef] [Green Version]
  6. Bedford, M.T.; Richard, S. Arginine methylation: An emerging regulator of protein function. Mol. Cell 2005, 18, 263–272. [Google Scholar] [CrossRef]
  7. Wu, H.; Min, J.; Lunin, V.V.; Antoshenko, T.; Dombrovski, L. Structural Biology of Human H3K9 Methyltransferases. PLoS ONE 2010, 5, e8570. [Google Scholar] [CrossRef] [Green Version]
  8. Tewary, S.K.; Zheng, Y.G.; Ho, M.C. Protein arginine methyltransferases: Insights into the enzyme structure and mechanism at the atomic level. Cell. Mol. Life Sci. 2019, 76, 2917–2932. [Google Scholar] [CrossRef]
  9. Jain, K.; Warmack, R.A.; Debler, E.W.; Hadjikyriacou, A.; Stavropoulos, P.; Clarke, S.G. Protein arginine methyltransferase product specificity is mediated by distinct active-site architectures. J. Biol. Chem. 2016, 291, 18299–18308. [Google Scholar] [CrossRef] [Green Version]
  10. Zhang, X.; Cheng, X. Structure of the predominant protein arginine methyltransferase PRMT1 and analysis of its binding to substrate peptides. Structure 2003, 11, 509–520. [Google Scholar] [CrossRef] [Green Version]
  11. Fuhrmann, J.; Clancy, K.; Thompson, P. Chemical biology of protein arginine modifications in epigenetic regulation. Chem. Rev. 2015, 115, 5413–5461. [Google Scholar] [CrossRef] [Green Version]
  12. Morales, Y.; Nitzel, D.V.; Price, O.M.; Gui, S.; Li, J.; Qu, J.; Hevel, J.M. Redox Control of Protein Arginine Methyltransferase 1 (PRMT1) Activity. J. Biol. Chem. 2015, 290, 14915–14926. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Weiss, V.; McBride, A.; Soriano, M.; Filman, D.; Silver, P.; Hogle, J. The structure and oligomerization of the yeast arginine methyltransferase, Hmt1. Nat. Struct. Biol. 2000, 7, 1165–1171. [Google Scholar] [CrossRef]
  14. Feng, Y.; Xie, N.; Jin, M.; Stahley, M.; Stivers, J.; Zheng, Y. A transient kinetic analysis of PRMT1 catalysis. Biochemistry 2011, 50, 7033–7044. [Google Scholar] [CrossRef] [Green Version]
  15. Herrmann, F.; Fackelmayer, F.O. Nucleo-cytoplasmic shuttling of protein arginine methyltransferase 1 (PRMT1) requires enzymatic activity. Genes Cells 2009, 14, 309–317. [Google Scholar] [CrossRef]
  16. Thandapani, P.; O’Connor, T.; Bailey, T.; Richard, S. Defining the RGG/RG motif. Mol. Cell 2013, 50, 613–623. [Google Scholar] [CrossRef] [Green Version]
  17. Morales, Y.; Cáceres, T.; May, K.; Hevel, J. Biochemistry and regulation of the protein arginine methyltransferases (PRMTs). Arch. Biochem. Biophys. 2016, 590, 138–152. [Google Scholar] [CrossRef]
  18. Wooderchak, W.; Zang, T.Z.; Zhou, Z.S.; Acuna, M.; Taharam, S.M.; Hevel, J. Substrate profiling of PRMT1 reveals amino acid sequences that extend beyond the “RGG” paradigm. Biochemistry 2008, 47, 9456–9466. [Google Scholar] [CrossRef] [PubMed]
  19. Osborne, T.C.; Obianyo, O.; Zhang, X.; Cheng, X.; Thompson, P.R. Protein Arginine Methyltransferase 1: Positively Charged Residues in Substrate Peptides Distal to the Site of Methylation Are Important for Substrate Binding and Catalysis. Biochemistry 2007, 46, 13370. [Google Scholar] [CrossRef] [Green Version]
  20. Gui, S.; Wooderchak, W.; Daly, M.; Porter, P.; Johnson, S.; Hevel, J. Investigation of the molecular origins of protein-arginine methyltransferase I (PRMT1) product specificity reveals a role for two conserved methionine residues. J. Biol. Chem. 2011, 286, 29118–29126. [Google Scholar] [CrossRef] [Green Version]
  21. Kirmizis, A.; Santos-Rosa, H.; Penkett, C.J.; Singer, M.A.; Vermeulen, M.; Mann, M.; Bähler, J.; Green, R.D.; Kouzarides, T. Arginine methylation at histone H3R2 controls deposition of H3K4 trimethylation. Nature 2007, 449, 928. [Google Scholar] [CrossRef] [Green Version]
  22. Gui, S.; Gathiaka, S.; Li, J.; Qu, J.; Acevedo, O.; Hevel, J.M. A remodeled protein arginine methyltransferase 1 (PRMT1) generates symmetric dimethylarginine. J. Biol. Chem. 2014, 289, 9320–9327. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Gathiaka, S.; Boykin, B.; Cáceres, T.; Hevel, J.M.; Acevedo, O. Understanding protein arginine methyltransferase 1 (PRMT1) product specificity from molecular dynamics. Bioorg. Med. Chem. 2016, 24, 4949–4960. [Google Scholar] [CrossRef] [Green Version]
  24. Gui, S.; WL, W.-D.; Zang, T.; Chen, D.; Daly, M.; Zhou, Z.; Hevel, J. Substrate-induced control of product formation by protein arginine methyltransferase 1. Biochemistry 2013, 52, 199–209. [Google Scholar] [CrossRef]
  25. Brown, J.I.; Koopmans, T.; van Strien, J.; Martin, N.I.; Frankel, A. Kinetic Analysis of PRMT1 Reveals Multifactorial Processivity and a Sequential Ordered Mechanism. ChemBioChem 2018, 19, 85–99. [Google Scholar] [CrossRef] [Green Version]
  26. Kölbel, K.; Ihling, C.; Bellmann-Sickert, K.; Neundorf, I.; Beck-Sickinger, A.G.; Sinz, A.; Kühn, U.; Wahle, E. Type I arginine methyltransferases PRMT1 and PRMT-3 act distributively. J. Biol. Chem. 2009, 284, 8274–8282. [Google Scholar] [CrossRef] [Green Version]
  27. Lakowski, T.M.; Frankel, A. Kinetic analysis of human protein arginine N-methyltransferase 2: Formation of monomethyl- and asymmetric dimethyl-arginine residues on histone H4. Biochem. J. 2009, 421, 253–261. [Google Scholar] [CrossRef]
  28. Hu, H.; Luo, C.; Zheng, Y.G. Transient kinetics define a complete kinetic model for protein arginine methyltransferase 1. J. Biol. Chem. 2016, 291, 26722–26738. [Google Scholar] [CrossRef] [Green Version]
  29. Obianyo, O.; Osborne, T.C.; Thompson, P.R. Kinetic mechanism of protein arginine methyltransferase. Biochemistry 2008, 47, 10420–10427. [Google Scholar] [CrossRef] [Green Version]
  30. Obianyo, O.; Causey, C.P.; Jones, J.E.; Thompson, P.R. Activity-Based Protein Profiling of Protein Arginine Methyltransferase 1. ACS Chem. Biol. 2011, 6, 1127–1135. [Google Scholar] [CrossRef] [Green Version]
  31. Wang, Y.; Wysocka, J.; Sayegh, J.; Lee, Y.; Perlin, J.; Leonelli, L.; Sonbuchner, L.; McDonald, C.; Cook, R.; Dou, Y.; et al. Human PAD4 regulates histone arginine methylation levels via demethylimination. Science 2004, 306, 279–283. [Google Scholar] [CrossRef]
  32. Chang, B.; Chen, Y.; Zhao, Y.; Bruick, R. JMJD6 is a histone arginine demethylase. Science 2007, 318, 444–447. [Google Scholar] [CrossRef] [Green Version]
  33. Li, S.; Ali, S.; Duan, X.; Liu, S.; Du, J.; Liu, C.; Dai, H.; Zhou, M.; Zhou, L.; Yang, L.; et al. JMJD1B Demethylates H4R3me2s and H3K9me2 to Facilitate Gene Expression for Development of Hematopoietic Stem and Progenitor Cells. Cell Rep. 2018, 23, 389–403. [Google Scholar] [CrossRef] [Green Version]
  34. Zhang, X.; Li, L.; Li, Y.; Li, Z.; Zhai, W.; Sun, Q.; Yang, X.; Roth, M.; Lu, S. mTOR regulates PRMT1 expression and mitochondrial mass through STAT1 phosphorylation in hepatic cell. Biochim. Biophys. Acta Mol. Cell Res. 2021, 1868, 119017. [Google Scholar] [CrossRef]
  35. Li, B.; Liu, L.; Li, X.; Wu, L. miR-503 suppresses metastasis of hepatocellular carcinoma cell by targeting PRMT1. Biochem. Biophys. Res. Commun. 2015, 464, 982–987. [Google Scholar] [CrossRef]
  36. Rust, H.L.; Subramanian, V.; West, G.M.; Young, D.D.; Schultz, P.G.; Thompson, P.R. Using unnatural amino acid mutagenesis to probe the regulation of PRMT1. ACS Chem. Biol. 2014, 9, 649–655. [Google Scholar] [CrossRef]
  37. Bao, X.; Siprashvili, Z.; Zarnegar, B.J.; Shenoy, R.M.; Rios, E.J.; Nady, N.; Qu, K.; Mah, A.; Webster, D.E.; Rubin, A.J.; et al. CSNK1a1 Regulates PRMT1 to Maintain the Progenitor State in Self-Renewing Somatic Tissue. Dev. Cell 2017, 43, 227–239.e5. [Google Scholar] [CrossRef] [Green Version]
  38. Musiani, D.; Giambruno, R.; Massignani, E.; Ippolito, M.R.; Maniaci, M.; Jammula, S.; Manganaro, D.; Cuomo, A.; Nicosia, L.; Pasini, D.; et al. PRMT1 Is Recruited via DNA-PK to Chromatin Where It Sustains the Senescence-Associated Secretory Phenotype in Response to Cisplatin. Cell Rep. 2020, 30, 1208–1222.e9. [Google Scholar] [CrossRef] [Green Version]
  39. Hirata, Y.; Katagiri, K.; Nagaoka, K.; Morishita, T.; Kudoh, Y.; Hatta, T.; Naguro, I.; Kano, K.; Udagawa, T.; Natsume, T.; et al. TRIM48 Promotes ASK1 Activation and Cell Death through Ubiquitination-Dependent Degradation of the ASK1-Negative Regulator PRMT1. Cell Rep. 2017, 21, 2447–2457. [Google Scholar] [CrossRef] [Green Version]
  40. Bhuripanyo, K.; Wang, Y.; Liu, X.; Zhou, L.; Liu, R.; Duong, D.; Zhao, B.; Bi, Y.; Zhou, H.; Chen, G.; et al. Identifying the substrate proteins of U-box E3s E4B and CHIP by orthogonal ubiquitin transfer. Sci. Adv. 2018, 4, e1701393. [Google Scholar] [CrossRef] [Green Version]
  41. Nie, M.; Wang, Y.; Guo, C.; Li, X.; Wang, Y.; Deng, Y.; Yao, B.; Gui, T.; Ma, C.; Liu, M.; et al. CARM1-mediated methylation of protein arginine methyltransferase 5 represses human γ-globin gene expression in erythroleukemia cells. J. Biol. Chem. 2018, 293, 17454–17463. [Google Scholar] [CrossRef] [Green Version]
  42. Sakabe, K.; Hart, G.W. O-GlcNAc Transferase Regulates Mitotic Chromatin Dynamics. J. Biol. Chem. 2010, 285, 34460–34468. [Google Scholar] [CrossRef] [Green Version]
  43. Lin, W.J.; Gary, J.D.; Yang, M.C.; Clarke, S.; Herschman, H.R. The mammalian immediate-early TIS21 protein and the leukemia-associated BTG1 protein interact with a protein-arginine N-methyltransferase. J. Biol. Chem. 1996, 271, 15034–15044. [Google Scholar] [CrossRef] [Green Version]
  44. Robin-Lespinasse, Y.; Sentis, S.; Kolytcheff, C.; Rostan, M.C.; Corbo, L.; Le Romancer, M. hCAF1, a new regulator of PRMT1-dependent arginine methylation. J. Cell Sci. 2007, 120, 638–647. [Google Scholar] [CrossRef] [Green Version]
  45. Pak, M.L.; Lakowski, T.M.; Thomas, D.; Vhuiyan, M.I.; Hüsecken, K.; Frankel, A. A protein arginine N-methyltransferase 1 (PRMT1) and 2 heteromeric interaction increases PRMT1 enzymatic activity. Biochemistry 2011, 50, 8226–8240. [Google Scholar] [CrossRef]
  46. Duong, F.H.T.; Christen, V.; Berke, J.M.; Penna, S.H.; Moradpour, D.; Heim, M.H. Upregulation of protein phosphatase 2Ac by hepatitis C virus modulates NS3 helicase activity through inhibition of protein arginine methyltransferase 1. J. Virol. 2005, 79, 15342–15350. [Google Scholar] [CrossRef] [Green Version]
  47. Gasperini, L.; Rossi, A.; Cornella, N.; Peroni, D.; Zuccotti, P.; Potrich, V.; Quattrone, A.; Macchi, P. The hnRNP RALY regulates PRMT1 expression and interacts with the ALS-linked protein FUS: Implication for reciprocal cellular localization. Mol. Biol. Cell 2018, 29, 3067–3081. [Google Scholar] [CrossRef]
  48. Lei, N.; Zhang, X.; Chen, H.; Wang, Y.; Zhan, Y.; Zheng, Z.; Shen, Y.; Wu, Q. A feedback regulatory loop between methyltransferase PRMT1 and orphan receptor TR3. Nucleic Acids Res. 2009, 37, 832–848. [Google Scholar] [CrossRef]
  49. Sun, Q.; Liu, L.; Mandal, J.; Molino, A.; Stolz, D.; Tamm, M.; Lu, S.; Roth, M. PDGF-BB induces PRMT1 expression through ERK1/2 dependent STAT1 activation and regulates remodeling in primary human lung fibroblasts. Cell. Signal. 2016, 28, 307–315. [Google Scholar] [CrossRef]
  50. Vadnais, C.; Chen, R.; Fraszczak, J.; Yu, Z.; Boulais, J.; Pinder, J.; Frank, D.; Khandanpour, C.; Hébert, J.; Dellaire, G.; et al. GFI1 facilitates efficient DNA repair by regulating PRMT1 dependent methylation of MRE11 and 53BP1. Nat. Commun. 2018, 9, 1418. [Google Scholar] [CrossRef]
  51. Inoue, H.; Hanawa, N.; Katsumata, S.-I.; Aizawa, Y.; Katsumata-Tsuboi, R.; Tanaka, M.; Takahashi, N.; Uehara, M. Iron deficiency negatively regulates protein methylation via the downregulation of protein arginine methyltransferase. Heliyon 2020, 6, e05059. [Google Scholar] [CrossRef]
  52. Wagner, S.; Weber, S.; Kleinschmidt, M.A.; Nagata, K.; Bauer, U.-M. SET-mediated promoter hypoacetylation is a prerequisite for coactivation of the estrogen-responsive pS2 gene by PRMT1. J. Biol. Chem. 2006, 281, 27242–27250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Fulton, M.D.; Dang, T.; Brown, T.; Zheng, Y.G. Effects of substrate modifications on the arginine dimethylation activities of PRMT1 and PRMT5. Epigenetics 2020, 31, 1–18. [Google Scholar] [CrossRef] [PubMed]
  54. Strahl, B.D.; Briggs, S.D.; Brame, C.J.; Caldwell, J.A.; Koh, S.S.; Ma, H.; Cook, R.G.; Shabanowitz, J.; Hunt, D.F.; Stallcup, M.R.; et al. Methylation of histone H4 at arginine 3 occurs in vivo and is mediated by the nuclear receptor coactivator PRMT1. Curr. Biol. 2001, 11, 996–1000. [Google Scholar] [CrossRef] [Green Version]
  55. Stallcup, M.R. Role of protein methylation in chromatin remodeling and transcriptional regulation. Oncogene 2001, 20, 3014–3020. [Google Scholar] [CrossRef] [Green Version]
  56. Wang, H.; Huang, Z.; Xia, L.; Feng, Q.; Erdjument-Bromage, H.; Strahl, B.; Briggs, S.; Allis, C.; Wong, J.; Tempst, P.; et al. Methylation of histone H4 at arginine 3 facilitating transcriptional activation by nuclear hormone receptor. Science 2001, 293, 853–857. [Google Scholar] [CrossRef] [PubMed]
  57. Waldmann, T.; Izzo, A.; Kamieniarz, K.; Richter, F.; Vogler, C.; Sarg, B.; Lindner, H.; Young, N.L.; Mittler, G.; Garcia, B.A.; et al. Methylation of H2AR29 is a novel repressive PRMT6 target. Epigenetics Chromatin 2011, 4, 11. [Google Scholar] [CrossRef] [Green Version]
  58. Dhar, S.; Vemulapalli, V.; Patananan, A.N.; Huang, G.L.; Di Lorenzo, A.; Richard, S.; Comb, M.J.; Guo, A.; Clarke, S.G.; Bedford, M.T. Loss of the major type i arginine methyltransferase PRMT1 causes substrate scavenging by other PRMTs. Sci. Rep. 2013, 3, 138–152. [Google Scholar] [CrossRef]
  59. Guendel, I.; Carpio, L.; Pedati, C.; Schwartz, A.; Teal, C.; Kashanchi, F.; Kehn-Hall, K. Methylation of the Tumor Suppressor Protein, BRCA1, Influences Its Transcriptional Cofactor Function. PLoS ONE 2010, 5, e11379. [Google Scholar] [CrossRef] [Green Version]
  60. Liu, L.-M.; Sun, W.-Z.; Fan, X.-Z.; Xu, Y.-L.; Cheng, M.-B.; Zhang, Y. Molecular Cell Biology Methylation of C/EBPa by PRMT1 Inhibits Its Tumor-Suppressive Function in Breast Cancer. Cancer Res. 2019, 79, 2865–2877. [Google Scholar] [CrossRef] [Green Version]
  61. Tikhanovich, I.; Zhao, J.; Bridges, B.; Kumer, S.; Roberts, B.; Weinman, S.A. Arginine methylation regulates c-Myc–dependent transcription by altering promoter recruitment of the acetyltransferase p300. J. Biol. Chem. 2017, 292, 13333–13344. [Google Scholar] [CrossRef] [Green Version]
  62. Li, Z.; Wang, D.; Lu, J.; Huang, B.; Wang, Y.; Dong, M.; Fan, D.; Li, H.; Gao, Y.; Hou, P.; et al. Methylation of EZH2 by PRMT1 regulates its stability and promotes breast cancer metastasis. Cell Death Differ. 2020, 27, 3226–3242. [Google Scholar] [CrossRef]
  63. Yamagata, K.; Daitoku, H.; Takahashi, Y.; Namiki, K.; Hisatake, K.; Kako, K.; Mukai, H.; Kasuya, Y.; Fukamizu, A. Arginine methylation of FOXO transcription factors inhibits their phosphorylation by Akt. Mol. Cell 2008, 32, 221–231. [Google Scholar] [CrossRef] [PubMed]
  64. Kagoya, Y.; Saijo, H.; Matsunaga, Y.; Guo, T.; Saso, K.; Anczurowski, M.; Wang, C.H.; Sugata, K.; Murata, K.; Butler, M.O.; et al. Arginine methylation of FOXP3 is crucial for the suppressive function of regulatory T cells. J. Autoimmun. 2019, 97, 10–21. [Google Scholar] [CrossRef] [PubMed]
  65. Wang, Y.; Hsu, J.-M.; Kang, Y.; Wei, Y.; Lee, P.-C.; Chang, S.-J.; Hsu, Y.-H.; Hsu, J.L.; Wang, H.-L.; Chang, W.-C.; et al. Oncogenic functions of Gli in pancreatic adenocarcinoma are supported by its PRMT1-mediated methylation. Cancer Res. 2016, 76, 7049. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Liu, Q.; Zhang, X.; Cheng, M.; Zhang, Y. PRMT1 activates myogenin transcription via MyoD arginine methylation at R121. Biochim. Biophys. Acta Gene Regul. Mech. 2019, 1862, 194442. [Google Scholar] [CrossRef] [PubMed]
  67. Liu, X.; Li, H.; Liu, L.; Lu, Y.; Gao, Y.; Geng, P.; Li, X.; Huang, B.; Zhang, Y.; Lu, J. Methylation of arginine by PRMT1 regulates Nrf2 transcriptional activity during the antioxidative response. Biochim. Biophys. Acta 2016, 1863, 2093–2103. [Google Scholar] [CrossRef]
  68. Malbeteau, L.; Poulard, C.; Languilaire, C.; Mikaelian, I.; Flamant, F.; Le Romancer, M.; Corbo, L. PRMT1 Is Critical for the Transcriptional Activity and the Stability of the Progesterone Receptor. IScience 2020, 23, 101236. [Google Scholar] [CrossRef]
  69. Davies, C.C.; Chakraborty, A.; Diefenbacher, M.E.; Skehel, M.; Behrens, A. Arginine methylation of the c-Jun coactivator RACO-1 is required for c-Jun/AP-1 activation. EMBO J. 2013, 32, 1556. [Google Scholar] [CrossRef]
  70. Reintjes, A.; Fuchs, J.E.; Kremser, L.; Lindner, H.H.; Liedl, K.R.; Huber, L.A.; Valovka, T. Asymmetric arginine dimethylation of RelA provides a repressive mark to modulate TNFα/NF-κB response. Proc. Natl. Acad. Sci. USA 2016, 113, 4326–4331. [Google Scholar] [CrossRef] [Green Version]
  71. Huq, M.D.M.; Gupta, P.; Tsai, N.-P.; White, R.; Parker, M.G.; Wei, L.-N. Suppression of receptor interacting protein 140 repressive activity by protein arginine methylation. EMBO J. 2006, 25, 5094–5104. [Google Scholar] [CrossRef] [Green Version]
  72. Zhao, X.; Jankovic, V.; Gural, A.; Huang, G.; Pardanani, A.; Menendez, S.; Zhang, J.; Dunne, R.; Xiao, A.; Erdjument-Bromage, H.; et al. Methylation of RUNX1 by PRMT1 abrogates SIN3A binding and potentiates its transcriptional activity. Genes Dev. 2008, 22, 640. [Google Scholar] [CrossRef] [Green Version]
  73. Mowen, K.A.; Tang, J.; Zhu, W.; Schurter, B.T.; Shuai, K.; Herschman, H.R.; David, M. Arginine Methylation of STAT1 Modulates IFNα/β-Induced Transcription. Cell 2001, 104, 731–741. [Google Scholar] [CrossRef]
  74. Jobert, L.; Argentini, M.; Tora, L. PRMT1 mediated methylation of TAF15 is required for its positive gene regulatory function. Exp. Cell Res. 2009, 315, 1273–1286. [Google Scholar] [CrossRef] [PubMed]
  75. Tradewell, M.; Yu, Z.; Tibshirani, M.; Boulanger, M.; Durham, H.; Richard, S. Arginine methylation by PRMT1 regulates nuclear-cytoplasmic localization and toxicity of FUS/TLS harbouring ALS-linked mutations. Hum. Mol. Genet. 2012, 21, 136–149. [Google Scholar] [CrossRef] [Green Version]
  76. Du, K.; Arai, S.; Kawamura, T.; Matsushita, A.; Kurokawa, R. TLS and PRMT1 synergistically coactivate transcription at the survivin promoter through TLS arginine methylation. Biochem. Biophys. Res. Commun. 2011, 404, 991–996. [Google Scholar] [CrossRef] [PubMed]
  77. Huang, L.; Wang, Z.; Narayanan, N.; Yang, Y. Arginine methylation of the C-terminus RGG motif promotes TOP3B topoisomerase activity and stress granule localization. Nucleic Acids Res. 2018, 46, 3061–3074. [Google Scholar] [CrossRef]
  78. Avasarala, S.; Van Scoyk, M.; Kumar, M.; Rathinam, K.; Zerayesus, S.; Zhao, X.; Zhang, W.; Pergande, M.R.; Borgia, J.A.; Degregori, J.; et al. PRMT1 Is a Novel Regulator of Epithelial-Mesenchymal-Transition in Non-small Cell Lung Cancer. J. Biol. Chem. 2015, 290, 13479–13489. [Google Scholar] [CrossRef] [Green Version]
  79. Wei, H.-M.; Hu, H.-H.; Chang, G.-Y.; Lee, Y.-J.; Li, Y.-C.; Chang, H.-H.; Li, C. Arginine methylation of the cellular nucleic acid binding protein does not affect its subcellular localization but impedes RNA binding. FEBS Lett. 2014, 588, 1542–1548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Tsai, W.-C.; Gayatri, S.; Reineke, L.C.; Sbardella, G.; Bedford, M.T.; Lloyd, R.E. Arginine Demethylation of G3BP1 Promotes Stress Granule Assembly. J. Biol. Chem. 2016, 291, 22671. [Google Scholar] [CrossRef] [Green Version]
  81. Wall, M.L.; Lewis, S.M. Methylarginines within the RGG-Motif Region of hnRNP A1 Affect Its IRES Trans-Acting Factor Activity and Are Required for hnRNP A1 Stress Granule Localization and Formation. J. Mol. Biol. 2017, 429, 295–307. [Google Scholar] [CrossRef]
  82. Wang, L.; Jia, Z.; Xie, D.; Zhao, T.; Tan, Z.; Zhang, S.; Kong, F.; Wei, D.; Xie, K. Methylation of HSP70 orchestrates its binding to and stabilization of BCL2 mRNA and renders pancreatic cancer cells resistant to therapeutics. Cancer Res. 2021, 80, 4500–4513. [Google Scholar] [CrossRef] [PubMed]
  83. Rho, J.; Choi, S.; Seong, Y.R.; Choi, J.; Im, D.-S. The Arginine-1493 Residue in QRRGRTGR1493G Motif IV of the Hepatitis C Virus NS3 Helicase Domain Is Essential for NS3 Protein Methylation by the Protein Arginine Methyltransferase 1. J. Virol. 2001, 75, 8031. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Zhang, L.; Tran, N.-T.; Su, H.; Wang, R.; Lu, Y.; Tang, H.; Aoyagi, S.; Guo, A.; Khodadadi-Jamayran, A.; Zhou, D.; et al. Cross-talk between PRMT1-mediated methylation and ubiquitylation on RBM15 controls RNA splicing. eLife 2015, 4, e60742. [Google Scholar] [CrossRef] [PubMed]
  85. Côté, J.; Boisvert, F.-M.; Boulanger, M.-C.; Bedford, M.T.; Richard, S. Sam68 RNA Binding Protein Is an In Vivo Substrate for Protein Arginine N-Methyltransferase 1. Mol. Biol. Cell 2003, 14, 274. [Google Scholar] [CrossRef] [Green Version]
  86. 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. [Google Scholar] [CrossRef]
  87. Sinha, R.; Allemand, E.; Zhang, Z.; Karni, R.; Myers, M.P.; Krainer, A.R. Arginine Methylation Controls the Subcellular Localization and Functions of the Oncoprotein Splicing Factor SF2/ASF. Mol. Cell. Biol. 2010, 30, 2762–2774. [Google Scholar] [CrossRef] [Green Version]
  88. Jia, H.; Du, C.H.; Bao, S.L.; Zheng, H.Y. Protein arginine methyltransferase 1 methylates SF2/ASF at arginine. Chin. J. Cancer Biother. 2009, 16, 216–220. [Google Scholar] [CrossRef]
  89. Tahara, S.M.; Acuna, M. Discrimination of eIF4A isoforms by protein arginine methyltransferase 1 (PRMT1). FASEB J. 2006, 20, A109–A110. [Google Scholar] [CrossRef]
  90. Hsu, J.H.-R.; Hubbell-Engler, B.; Adelmant, G.; Huang, J.; Joyce, C.E.; Vazquez, F.; Weir, B.A.; Montgomery, P.; Tsherniak, A.; Giacomelli, A.O.; et al. Prmt1-mediated translation regulation is a crucial vulnerability of cancer. Cancer Res. 2017, 77, 4613. [Google Scholar] [CrossRef] [Green Version]
  91. Shin, H.-S.; Jang, C.-Y.; Kim, H.D.; Kim, T.-S.; Kim, S.; Kim, J. Arginine methylation of ribosomal protein S3 affects ribosome assembly. Biochem. Biophys. Res. Commun. 2009, 385, 273–278. [Google Scholar] [CrossRef]
  92. Boisvert, F.-M.; Rhie, A.; Richard, S.; Doherty, A.J. The GAR Motif of 53BP1 is Arginine Methylated by PRMT1 and is Necessary for 53BP1 DNA Binding Activity. Cell Cycle 2005, 4, 1834–1841. [Google Scholar] [CrossRef]
  93. Zhang, Y.; Zhang, Q.; Li, L.L.; Mu, D.; Hua, K.; Ci, S.; Shen, L.; Zheng, L.; Shen, B.; Guo, Z. Arginine methylation of APE1 promotes its mitochondrial translocation to protect cells from oxidative damage. Free Radic. Biol. Med. 2020, 158, 60–73. [Google Scholar] [CrossRef]
  94. El-Andaloussi, N.; Valovka, T.; Toueille, M.; Hassa, P.O.; Gehrig, P.; Covic, M.; Hübscher, U.; Hottiger, M.O. Methylation of DNA polymerase ß by protein arginine methyltransferase 1 regulates its binding to proliferating cell nuclear antigen. FASEB J. 2007, 21, 26–34. [Google Scholar] [CrossRef] [PubMed]
  95. Zheng, S.; Moehlenbrink, J.; Lu, Y.; Zalmas, L.; Sagum, C.; Carr, S.; McGouran, J.; 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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. He, L.; Hu, Z.; Sun, Y.; Zhang, M.; Zhu, H.; Jiang, L.; Zhang, Q.; Mu, D.; Zhang, J.; Gu, L.; et al. PRMT1 is critical to FEN1 expression and drug resistance in lung cancer cells. DNA Repair 2020, 95, 102953. [Google Scholar] [CrossRef] [PubMed]
  97. Yang, J.; Chiou, Y.; Fu, S.; Shih, I.; Weng, T.; Lin, W.; Lin, C. Arginine methylation of hnRNPK negatively modulates apoptosis upon DNA damage through local regulation of phosphorylation. Nucleic Acids Res. 2014, 42, 9908–9924. [Google Scholar] [CrossRef] [Green Version]
  98. Gurunathan, G.; Yu, Z.; Coulombe, Y.; Masson, J.-Y.; Richard, S. Arginine methylation of hnRNPUL1 regulates interaction with NBS1 and recruitment to sites of DNA damage. Sci. Rep. 2015, 5, 1–9. [Google Scholar] [CrossRef] [Green Version]
  99. Boisvert, F.M.; Déry, U.; Masson, J.Y.; Richard, S. Arginine methylation of MRE11 by PRMT1 is required for DNA damage checkpoint control. Genes Dev. 2005, 19, 671–676. [Google Scholar] [CrossRef] [Green Version]
  100. Déry, U.; Coulombe, Y.; Rodrigue, A.; Stasiak, A.; Richard, S.; Masson, J.-Y. A Glycine-Arginine Domain in Control of the Human MRE11 DNA Repair Protein. Mol. Cell. Biol. 2008, 28, 3058. [Google Scholar] [CrossRef] [Green Version]
  101. Matsumura, T.; Nakamura-Ishizu, A.; Anurag Muddineni, S.S.N.; Tan, D.Q.; Wang, C.Q.; Tokunaga, K.; Tirado-Magallanes, R.; Sian, S.; Benoukraf, T.; Okuda, T.; et al. Hematopoietic stem cells acquire survival advantage by loss of RUNX1 methylation identified in familial leukemia. Blood 2020, 136, 1919–1932. [Google Scholar] [CrossRef]
  102. Cho, J.-H.; Lee, M.-K.; Yoon, K.W.; Lee, J.; Cho, S.-G.; Choi, E.-J. Arginine methylation-dependent regulation of ASK1 signaling by PRMT1. Cell Death Differ. 2012, 19, 859. [Google Scholar] [CrossRef] [Green Version]
  103. Cha, B.; Kim, W.; Kim, Y.K.; Hwang, B.N.; Park, S.Y.; Yoon, J.W.; Park, W.S.; Cho, J.W.; Bedford, M.T.; Jho, E.H. Methylation by protein arginine methyltransferase 1 increases stability of Axin, a negative regulator of Wnt signaling. Oncogene 2011, 30, 2379–2389. [Google Scholar] [CrossRef] [Green Version]
  104. Sakamaki, J.; Daitoku, H.; Ueno, K.; Hagiwara, A.; Yamagata, K.; Fukamizu, A. Arginine methylation of BCL-2 antagonist of cell death (BAD) counteracts its phosphorylation and inactivation by Akt. Proc. Natl. Acad. Sci. USA 2011, 108, 6085–6090. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Pyun, J.; Kim, H.; Jeong, M.; Ahn, B.; Vuong, T.; Lee, D.; Choi, S.; Koo, S.; Cho, H.; Kang, J. Cardiac specific PRMT1 ablation causes heart failure through CaMKII dysregulation. Nat. Commun. 2018, 9, 1–15. [Google Scholar] [CrossRef] [Green Version]
  106. Dolezal, E.; Infantino, S.; Drepper, F.; Börsig, T.; Singh, A.; Wossning, T.; Fiala, G.J.; Minguet, S.; Warscheid, B.; Tarlinton, D.M.; et al. The BTG2-PRMT1 module limits pre-B cell expansion by regulating the CDK4-Cyclin-D3 complex. Nat. Immunol. 2017, 18, 911–920. [Google Scholar] [CrossRef] [PubMed]
  107. Onwuli, D.O.; Samuel, S.F.; Sfyri, P.; Welham, K.; Goddard, M.; Abu-Omar, Y.; Loubani, M.; Rivero, F.; Matsakas, A.; Benoit, D.M.; et al. The inhibitory subunit of cardiac troponin (cTnI) is modified by arginine methylation in the human heart. Int. J. Cardiol. 2019, 282, 76–80. [Google Scholar] [CrossRef]
  108. Liao, H.-W.; Hsu, J.-M.; Xia, W.; Wang, H.-L.; Wang, Y.-N.; Chang, W.-C.; Arold, S.T.; Chou, C.-K.; Tsou, P.-H.; Yamaguchi, H.; et al. PRMT1-mediated methylation of the EGF receptor regulates signaling and cetuximab response. J. Clin. Investig. 2015, 125, 4529–4543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Le Romancer, M.; Treilleux, I.; Leconte, N.; Robin-Lespinasse, Y.; Sentis, S.; Bouchekioua-Bouzaghou, K.; Goddard, S.; Gobert-Gosse, S.; Corbo, L. Regulation of Estrogen Rapid Signaling through Arginine Methylation by PRMT1. Mol. Cell 2008, 31, 212–221. [Google Scholar] [CrossRef]
  110. Deng, X.; Keudell, G. Von 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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Kim, H.J.; Jeong, M.H.; Kim, K.R.; Jung, C.Y.; Lee, S.Y.; Kim, H.; Koh, J.; Vuong, T.A.; Jung, S.; Yang, H.; et al. Protein arginine methylation facilitates KCNQ channel-PIP2 interaction leading to seizure suppression. eLife 2016, 5, e17159. [Google Scholar] [CrossRef] [PubMed]
  112. Eberhardt, A.; Hansen, J.N.; Koster, J.; Lotta, L.T., Jr.; Wang, S.; Livingstone, E.; Qian, K.; Valentijn, L.J.; Zheng, Y.G.; Schor, N.F.; et al. Protein arginine methyltransferase 1 is a novel regulator of MYCN in neuroblastoma. Oncotarget 2016, 7, 63629. [Google Scholar] [CrossRef] [PubMed]
  113. Yin, X.-K.; Wang, Y.-L.; Wang, F.; Feng, W.-X.; Bai, S.-M.; Zhao, W.-W.; Feng, L.-L.; Wei, M.-B.; Qin, C.-L.; Wang, F.; et al. PRMT1 enhances oncogenic arginine methylation of NONO in colorectal cancer. Oncogene 2021, 40, 1375. [Google Scholar] [CrossRef] [PubMed]
  114. Liu, M.; Hua, W.; Chen, C.; Lin, W. The MKK-Dependent Phosphorylation of p38α Is Augmented by Arginine Methylation on Arg49/Arg149 during Erythroid Differentiation. Int. J. Mol. Sci. 2020, 21, 3546. [Google Scholar] [CrossRef]
  115. Albrecht, L.V.; Ploper, D.; Tejeda-Muñoz, N.; De Robertis, E.M. Arginine methylation is required for canonical Wnt signaling and endolysosomal trafficking. Proc. Natl. Acad. Sci. USA 2018, 115, E5317–E5325. [Google Scholar] [CrossRef] [Green Version]
  116. Xu, J.; Wang, A.H.; Oses-Prieto, J.; Makhijani, K.; Katsuno, Y.; Pei, M.; Yan, L.; Zheng, Y.G.; Burlingame, A.; Brückner, K.; et al. Arginine Methylation Initiates BMP-Induced Smad Signaling. Mol. Cell 2013, 51, 5–19. [Google Scholar] [CrossRef] [Green Version]
  117. Zhang, T.; Wu, J.; Ungvijanpunya, N.; Jackson-Weaver, O.; Gou, Y.; Feng, J.; Ho, T.; Shen, Y.; Liu, J.; Richard, S.; et al. Smad6 Methylation Represses NFκB Activation and Periodontal Inflammation. J. Dent. Res. 2018, 97, 810–819. [Google Scholar] [CrossRef]
  118. Katsuno, Y.; Qin, J.; Oses-Prieto, J.; Wang, H.; Jackson-Weaver, O.; Zhang, T.; Lamouille, S.; Wu, J.; Burlingame, A.; Xu, J.; et al. Arginine methylation of SMAD7 by PRMT1 in TGF-β-induced epithelial-mesenchymal transition and epithelial stem-cell generation. J. Biol. Chem. 2018, 293, 13059–137072. [Google Scholar] [CrossRef] [Green Version]
  119. Tikhanovich, I.; Kuravi, S.; Artigues, A.; Villar, M.T.; Dorko, K.; Nawabi, A.; Roberts, B.; Weinman, S.A. Dynamic Arginine Methylation of Tumor Necrosis Factor (TNF) Receptor-associated Factor 6 Regulates Toll-like Receptor Signaling. J. Biol. Chem. 2015, 290, 22236–22249. [Google Scholar] [CrossRef] [Green Version]
  120. Gen, S.; Matsumoto, Y.; Kobayashi, K.-I.; Suzuki, T.; Inoue, J.; Yamamoto, Y. Stability of tuberous sclerosis complex 2 is controlled by methylation at R1457 and R1459. Sci. Rep. 2020, 10, 1–9. [Google Scholar] [CrossRef]
  121. Yang, Y.; Lu, Y.; Espejo, A.; Wu, J.; Xu, W.; Liang, S.; Bedford, M.T. TDRD3 is an Effector Molecule for Arginine Methylated Histone Marks. Mol. Cell 2010, 40, 1016. [Google Scholar] [CrossRef] [PubMed]
  122. Yang, Y.; McBride, K.M.; Hensley, S.; Lu, Y.; Chedin, F.; Bedford, M.T. Arginine methylation facilitates the recruitment of TOP3B to chromatin to prevent R-loop accumulation. Mol. Cell 2014, 53, 484. [Google Scholar] [CrossRef] [Green Version]
  123. Sims, R.J., III; Rojas, L.A.; Beck, D.B.; Bonasio, R.; Schüller, R.; Drury, W.J., III; Eick, D.; Reinberg, D. The C-Terminal Domain of RNA Polymerase II Is Modified by Site-Specific Methylation. Science 2011, 332, 99. [Google Scholar] [CrossRef] [Green Version]
  124. Huang, S.; Litt, M.; Felsenfeld, G. Methylation of histone H4 by arginine methyltransferase PRMT1 is essential in vivo for many subsequent histone modifications. Genes Dev. 2005, 19, 1885–1893. [Google Scholar] [CrossRef] [Green Version]
  125. Li, X.; Hu, X.; Patel, B.; Zhou, Z.; Liang, S.; Ybarra, R.; Qiu, Y.; Felsenfeld, G.; Bungert, J.; Huang, S. H4R3 methylation facilitates β-globin transcription by regulating histone acetyltransferase binding and H3 acetylation. Blood 2010, 115, 2028–2037. [Google Scholar] [CrossRef] [Green Version]
  126. Fulton, M.D.; Zhang, J.; He, M.; Ho, M.-C.; Zheng, Y.G. Intricate Effects of α-Amino and Lysine Modifications on Arginine Methylation of the N-Terminal Tail of Histone H4. Biochemistry 2017, 56, 3539–3548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Feng, Y.; Wang, J.; Asher, S.; Hoang, L.; Guardiani, C.; Ivanov, I.; Zheng, Y.G. Histone H4 acetylation differentially modulates arginine methylation by an in cis mechanism. J. Biol. Chem. 2011, 286, 20323–20334. [Google Scholar] [CrossRef] [Green Version]
  128. Thiebaut, C.; Vlaeminck-Guillem, V.; Trédan, O.; Poulard, C.; Le Romancer, M. Non-genomic signaling of steroid receptors in cancer. Mol. Cell. Endocrinol. 2021, 538, 111453. [Google Scholar] [CrossRef]
  129. Choucair, A.; Pham, T.H.; Omarjee, S.; Jacquemetton, J.; Kassem, L.; Trédan, O.; Rambaud, J.; Marangoni, E.; Corbo, L.; Treilleux, I.; et al. The arginine methyltransferase PRMT1 regulates IGF-1 signaling in breast cancer. Oncogene 2019, 38, 4015–4027. [Google Scholar] [CrossRef]
  130. Poulard, C.; Rambaud, J.; Hussein, N.; Corbo, L.; Le Romancer, M. JMJD6 regulates ERα methylation on arginine. PLoS ONE 2014, 9, e87982. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Greer, E.L.; Brunet, A. FOXO transcription factors at the interface between longevity and tumor suppression. Oncogene 2005, 24, 7410–7425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Brunet, A.; Bonni, A.; Zigmond, M.J.; Lin, M.Z.; Juo, P.; Hu, L.S.; Anderson, M.J.; Arden, K.C.; Blenis, J.; Greenberg, M.E. Akt Promotes Cell Survival by Phosphorylating and Inhibiting a Forkhead Transcription Factor. Cell 1999, 96, 857–868. [Google Scholar] [CrossRef] [Green Version]
  133. Kops, G.; Burgering, B. Forkhead transcription factors: New insights into protein kinase B (c-akt) signaling. J. Mol. Med. 1999, 77, 656–665. [Google Scholar] [CrossRef] [PubMed]
  134. Huang, H.; Tindall, D.J. Regulation of FoxO protein stability via ubiquitination and proteasome degradation. Biochim. Biophys. Acta 2011, 1813, 1961. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Matsuzaki, H.; Daitoku, H.; Hatta, M.; Tanaka, K.; Fukamizu, A. Insulin-induced phosphorylation of FKHR (Foxo1) targets to proteasomal degradation. Proc. Natl. Acad. Sci. USA 2003, 100, 11285–11290. [Google Scholar] [CrossRef] [Green Version]
  136. Hassa, P.O.; Covic, M.; Bedford, M.T.; Hottiger, M.O. Protein Arginine Methyltransferase 1 Coactivates NF-κB-Dependent Gene Expression Synergistically with CARM1 and PARP1. J. Mol. Biol. 2008, 377, 668–678. [Google Scholar] [CrossRef]
  137. Pawlak, M.R.; Scherer, C.A.; Chen, J.; Roshon, M.J.; Ruley, H.E. Arginine N-Methyltransferase 1 Is Required for Early Postimplantation Mouse Development, but Cells Deficient in the Enzyme Are Viable. Mol. Cell. Biol. 2000, 20, 4859. [Google Scholar] [CrossRef] [Green Version]
  138. Tsai, Y.; Pan, H.; Hung, C.; Hou, P.; Li, Y.; Lee, Y.; Shen, Y.; Wu, T.; Li, C. The predominant protein arginine methyltransferase PRMT1 is critical for zebrafish convergence and extension during gastrulation. FEBS J. 2011, 278, 905–917. [Google Scholar] [CrossRef]
  139. Shibata, Y.; Okada, M.; Miller, T.C.; Shi, Y.-B. Knocking out histone methyltransferase PRMT1 leads to stalled tadpole development and lethality in Xenopus tropicalis. Biochim. Biophys. Acta Gen. Subj. 2020, 1864, 129482. [Google Scholar] [CrossRef]
  140. Batut, J.; Vandel, L.; Leclerc, C.; Daguzan, C.; Moreau, M.; Néant, I. The Ca2+-induced methyltransferase xPRMT1b controls neural fate in amphibian embryo. Proc. Natl. Acad. Sci. USA 2005, 102, 15128–15133. [Google Scholar] [CrossRef] [Green Version]
  141. Sato, A.; Kim, J.D.; Mizukami, H.; Nakashima, M.; Kako, K.; Ishida, J.; Itakura, A.; Takeda, S.; Fukamizu, A. Gestational changes in PRMT1 expression of murine placentas. Placenta 2018, 65, 47–54. [Google Scholar] [CrossRef]
  142. 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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Yu, Z.; Vogel, G.; Coulombe, Y.; Dubeau, D.; Spehalski, E.; Hébert, J.; Ferguson, D.O.; Masson, J.Y.; Richard, S. The MRE11 GAR motif regulates DNA double-strand break processing and ATR activation. Cell Res. 2012, 22, 305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Mitchell, T.R.H.; Glenfield, K.; Jeyanthan, K.; Zhu, X.-D. Arginine Methylation Regulates Telomere Length and Stability. Mol. Cell. Biol. 2009, 29, 4918–4934. [Google Scholar] [CrossRef] [Green Version]
  145. Zhao, F.; Kim, W.; Kloeber, J.A.; Lou, Z. DNA end resection and its role in DNA replication and DSB repair choice in mammalian cells. Exp. Mol. Med. 2020, 52, 1705–1714. [Google Scholar] [CrossRef]
  146. Adams, M.M.; Wang, B.; Xia, Z.; Morales, J.C.; Lu, X.; Donehower, L.A.; Bochar, D.A.; Elledge, S.J.; Carpenter, P.B. 53BP1 Oligomerization is Independent of its Methylation by PRMT1. Cell Cycle 2005, 4, 12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Guo, Z.; Zheng, L.; Xu, H.; Dai, H.; Zhou, M.; Pascua, M.; Chen, Q.; Shen, B. Methylation of FEN1 suppresses nearby phosphorylation and facilitates PCNA binding. Nat. Chem. Biol. 2010, 6, 766–773. [Google Scholar] [CrossRef] [Green Version]
  148. Mathioudaki, K.; Scorilas, A.; Ardavanis, A.; Lymberi, P.; Tsiambas, E.; Devetzi, M.; Apostolaki, A.; Talieri, M. Clinical evaluation of PRMT1 gene expression in breast cancer. Tumor Biol. 2011, 32, 575–582. [Google Scholar] [CrossRef]
  149. Poulard, C.; Treilleux, I.; Lavergne, E.; Bouchekioua-Bouzaghou, K.; Goddard-Léon, S.; Chabaud, S.; Trédan, O.; Corbo, L.; Le Romancer, M. Activation of rapid oestrogen signalling in aggressive human breast cancers. EMBO Mol. Med. 2012, 4, 1200–1213. [Google Scholar] [CrossRef]
  150. Poulard, C.; Jacquemetton, J.; Trédan, O.; Cohen, P.A.; Vendrell, J.; Ghayad, S.E.; Treilleux, I.; Marangoni, E.; Le Romancer, M. Oestrogen non-genomic signalling is activated in tamoxifen-resistant breast cancer. Int. J. Mol. Sci. 2019, 20, 2773. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Jacquemetton, J.; Kassem, L.; Poulard, C.; Dahmani, A.; De Plater, L.; Montaudon, E.; Sourd, L.; Morisset, L.; El Botty, R.; Chateau-Joubert, S.; et al. Analysis of genomic and non-genomic signaling of estrogen receptor in PDX models of breast cancer treated with a combination of the PI3K inhibitor Alpelisib (BYL719) and fulvestrant. Breast Cancer Res. 2021; in press. [Google Scholar]
  152. Nakai, K.; Xia, W.; Liao, H.; Saito, M.; Hung, M.; Yamaguchi, H. The role of PRMT1 in EGFR methylation and signaling in MDA-MB-468 triple-negative breast cancer cells. Breast Cancer 2018, 25, 74–80. [Google Scholar] [CrossRef]
  153. Gao, Y.; Zhao, Y.; Zhang, J.; Lu, Y.; Liu, X.; Geng, P.; Huang, B.; Zhang, Y.; Lu, J. The dual function of PRMT1 in modulating epithelial-mesenchymal transition and cellular senescence in breast cancer cells through regulation of ZEB1. Sci. Rep. 2016, 6, 1–13. [Google Scholar] [CrossRef] [PubMed]
  154. Papadokostopoulou, A.; Mathioudaki, K.; Scorilas, A.; Xynopoulos, D.; Ardavanis, A.; Kouroumalis, E.; Talieri, M. Colon cancer and protein arginine methyltransferase 1 gene expression. Anticancer Res. 2009, 29, 1361–1366. [Google Scholar] [PubMed]
  155. 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. [Google Scholar] [CrossRef] [Green Version]
  156. Yao, B.; Gui, T.; Zeng, X.; Deng, Y.; Wang, Z.; Wang, Y.; Yang, D.; Li, Q.; Xu, P.; Hu, R.; et al. PRMT1-mediated H4R3me2a recruits SMARCA4 to promote colorectal cancer progression by enhancing EGFR signaling. Genome Med. 2021, 13, 1–21. [Google Scholar] [CrossRef]
  157. Yoshimatsu, M.; Toyokawa, G.; Hayami, S.; Unoki, M.; Tsunoda, T.; Field, H.I.; Kelly, J.D.; Neal, D.E.; Maehara, Y.; Ponder, B.A.J.; et al. Dysregulation of PRMT1 and PRMT6, Type I arginine methyltransferases, is involved in various types of human cancers. Int. J. Cancer 2011, 128, 562–573. [Google Scholar] [CrossRef]
  158. Zhao, Y.; Lu, Q.; Li, C.; Wang, X.; Jiang, L.; Huang, L.; Wang, C.; Chen, H. PRMT1 regulates the tumour-initiating properties of esophageal squamous cell carcinoma through histone H4 arginine methylation coupled with transcriptional activation. Cell Death Dis. 2019, 10, 1–17. [Google Scholar] [CrossRef]
  159. He, X.; Zhu, Y.; Lin, Y.; Li, M.; Du, J.; Dong, H.; Sun, J.; Zhu, L.; Wang, H.; Ding, Z.; et al. PRMT1-mediated FLT3 arginine methylation promotes maintenance of FLT3-ITD + acute myeloid leukemia. Blood 2019, 134, 548–560. [Google Scholar] [CrossRef]
  160. Matsubara, H.; Fukuda, T.; Awazu, Y.; Nanno, S.; Shimomura, M.; Inoue, Y.; Yamauchi, M.; Yasui, T.; Sumi, T. PRMT1 expression predicts sensitivity to platinum-based chemotherapy in patients with ovarian serous carcinoma. Oncol. Lett. 2021, 21, 1. [Google Scholar] [CrossRef]
  161. Seligson, D.B.; Horvath, S.; Shi, T.; Yu, H.; Tze, S.; Grunstein, M.; Kurdistani, S.K. Global histone modification patterns predict risk of prostate cancer recurrence. Nature 2005, 435, 1262–1266. [Google Scholar] [CrossRef]
  162. Cheng, D.; Yadav, N.; King, R.W.; Swanson, M.S.; Weinstein, E.J.; Bedford, M.T. Small Molecule Regulators of Protein Arginine Methyltransferases. J. Biol. Chem. 2004, 279, 23892–23899. [Google Scholar] [CrossRef] [Green Version]
  163. Eram, M.S.; Shen, Y.; Szewczyk, M.; Wu, H.; Senisterra, G.; Li, F.; Butler, K.V.; Kaniskan, H.Ü.; Speed, B.A.; Seña, C.; et al. A Potent, Selective and Cell-active Inhibitor of Human Type I Protein Arginine Methyltransferases. ACS Chem. Biol. 2016, 11, 772. [Google Scholar] [CrossRef] [Green Version]
  164. Fong, J.Y.; Pignata, L.; Goy, P.-A.; Kawabata, K.C.; Lee, S.C.-W.; Koh, C.M.; Musiani, D.; Massignani, E.; Kotini, A.G.; Penson, A.; et al. Therapeutic Targeting of RNA Splicing Catalysis through Inhibition of Protein Arginine Methylation. Cancer Cell 2019, 36, 194. [Google Scholar] [CrossRef] [PubMed]
  165. Fedoriw, A.; Rajapurkar, S.R.; O’Brien, S.; Gerhart, S.V.; Mitchell, L.H.; Adams, N.D.; Rioux, N.; Lingaraj, T.; Ribich, S.A.; Pappalardi, M.B.; et al. Anti-tumor Activity of the Type I PRMT Inhibitor, GSK3368715, Synergizes with PRMT5 Inhibition through MTAP Loss. Cancer Cell 2019, 36, 100–114.e25. [Google Scholar] [CrossRef]
  166. Zhang, P.; Tao, H.; Yu, L.; Zhou, L.; Zhu, C. Developing protein arginine methyltransferase 1 (PRMT1) inhibitor TC-E-5003 as an antitumor drug using INEI drug delivery systems. Drug Deliv. 2020, 27, 491–501. [Google Scholar] [CrossRef] [PubMed]
  167. Hu, H.; Qian, K.; Ho, M.-C.; Zheng, Y.G. Small Molecule Inhibitors of Protein Arginine Methyltransferases. Expert Opin. Investig. Drugs 2016, 25, 335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Genomic organization and protein structure of PRMT1. (A) Genomic organization of the PRMT1 gene which spans 11.3 kilobases (kb) and possesses 12 constitutive exons including exon 1 subdivided into 4 alternative exons, represented by a scale of blue-colored boxes; (B) Exon composition of the different PRMT-1 isoforms (v1 to v7). The sequences of intron boundaries are represented by the black boxes. Molecular weight of each protein isoform is indicated in kilodaltons (kDa); (C) Protein structure of PRMT1-v2. The PRMT signature motifs (I, Post-I (PI), II, III) as well as double-E motif, the THW loop and the phosphorylation site Y291 are represented (adapted from [5,6]).
Figure 1. Genomic organization and protein structure of PRMT1. (A) Genomic organization of the PRMT1 gene which spans 11.3 kilobases (kb) and possesses 12 constitutive exons including exon 1 subdivided into 4 alternative exons, represented by a scale of blue-colored boxes; (B) Exon composition of the different PRMT-1 isoforms (v1 to v7). The sequences of intron boundaries are represented by the black boxes. Molecular weight of each protein isoform is indicated in kilodaltons (kDa); (C) Protein structure of PRMT1-v2. The PRMT signature motifs (I, Post-I (PI), II, III) as well as double-E motif, the THW loop and the phosphorylation site Y291 are represented (adapted from [5,6]).
Life 11 01147 g001
Figure 2. PRMT1 regulates chromatin dynamics. PRMT1-dependent H4R3 methylation (R3me2a) allows the recruitment of the Tudor domain-containing protein, TDRD3, which in turn associates with topoisomerase IIIβ (Top IIIβ) to reduce R-loop formation and RNA polymerase II (RNA pol II) to promote transcriptional activity. Concomitantly, H4R3me2a-dependent activation of histone acetyltransferases p300 and pCAF induces acetylation of H4K5, H4K8, H4K12 but also of H3K9 and H3K14. H4K5 and H4K12 are involved in the recruitment of TAFII250 that associates with RNA pol II. H4K5ac and H4K16ac are also involved in PRMT1-activity regulation. Ac = Acetylation, m = methylation.
Figure 2. PRMT1 regulates chromatin dynamics. PRMT1-dependent H4R3 methylation (R3me2a) allows the recruitment of the Tudor domain-containing protein, TDRD3, which in turn associates with topoisomerase IIIβ (Top IIIβ) to reduce R-loop formation and RNA polymerase II (RNA pol II) to promote transcriptional activity. Concomitantly, H4R3me2a-dependent activation of histone acetyltransferases p300 and pCAF induces acetylation of H4K5, H4K8, H4K12 but also of H3K9 and H3K14. H4K5 and H4K12 are involved in the recruitment of TAFII250 that associates with RNA pol II. H4K5ac and H4K16ac are also involved in PRMT1-activity regulation. Ac = Acetylation, m = methylation.
Life 11 01147 g002
Table 2. List of PRMT inhibitors targeting PRMT1. ND: Not defined in literature.
Table 2. List of PRMT inhibitors targeting PRMT1. ND: Not defined in literature.
NameMechanism of ActionTarget(s)IC50Reference
AMI-1Substrate competitive
SAM uncompetitive
PRMT18.81 µM[162]
MS023Substrate competitive
SAM uncompetitive
PRMT130 nM[163]
PRMT3119 nM
PRMT4/CARM183 nM
PRMT64 nM
PRMT85 nM
GSK3368715Substrate competitive
SAM uncompetitive
Reversible
PRMT133.1 nM[165]
PRMT3162 nM
PRMT4/CARM138 nM
PRMT64.7 nM
PRMT83.9 nM
TC-E-5003NDPRMT11.5 µM[166]
C7280948Interaction with the
substrate-binding pocket
PRMT112.8 µM[113]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Thiebaut, C.; Eve, L.; Poulard, C.; Le Romancer, M. Structure, Activity, and Function of PRMT1. Life 2021, 11, 1147. https://doi.org/10.3390/life11111147

AMA Style

Thiebaut C, Eve L, Poulard C, Le Romancer M. Structure, Activity, and Function of PRMT1. Life. 2021; 11(11):1147. https://doi.org/10.3390/life11111147

Chicago/Turabian Style

Thiebaut, Charlène, Louisane Eve, Coralie Poulard, and Muriel Le Romancer. 2021. "Structure, Activity, and Function of PRMT1" Life 11, no. 11: 1147. https://doi.org/10.3390/life11111147

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop