Introduction

Amyotrophic lateral sclerosis (ALS) is a fatal neurodegenerative disease characterized by a deficiency of upper and lower motor neurons in the motor cortex and lumbar spinal cord, respectively [1]. Frontotemporal dementia (FTD) is a devastating disease that mainly involves the frontal and temporal lobes [2]. A GGGGCC (G4C2) hexanucleotide repeat expansion (HRE) in the intron of the C9ORF72 gene was identified in 2011 and is the most common genetic cause of both ALS and FTD [3,4,5]. Tens to thousands of G4C2 repeats have been identified in carriers and patients with the C9ORF72-related ALS and FTD (c9ALS/FTD) mutation, while only ~30 repeats occur in normal individuals [3, 4]. The HRE can be further transcribed and translated into sense and antisense RNAs, as well as dipeptide repeat proteins (DPRs) that include poly-GA, poly-GP, poly-GR, poly-PA, and poly-PR [6,7,8,9,10,11,12].

The underlying mechanisms of c9ALS/FTD can be classified into three prototypes (Fig. 1): (1) loss-of-function of the C9ORF72 protein [13,14,15,16,17]; (2) formation of sense and antisense RNA foci in the nucleus [18,19,20]; and (3) gain-of-function caused by repeat-associated non-ATG-initiated translation of DPRs [21,22,23,24]. In recent years, several studies have implicated C9ORF72 in cellular protein transport and that loss of C9ORF72 impairs autophagy [13, 14, 25,26,27] and lysosome biogenesis [28]. Despite the fact that C9ORF72 loss-of-function contributes to microglial activation and a “cytokine storm” in several transgenic mouse models, reducing the expression of C9ORF72 alone does not induce c9ALS/FTD phenotypes and is dispensable for neuronal survival [15,16,17, 26, 29]. Thus, we summarize the literature focusing on gain-of-function in this review.

Fig. 1
figure 1

Mechanisms underlying the GGGGCC hexanucleotide repeat expansion. (1) Loss-of-function: Decreased expression of C9orf72 mRNA in the frontal cortex of individuals with C9ORF72 mutation. (2) Gain-of-function caused by RNA foci: The GGGGCC repeat expansion is transcribed into repeat RNA that can interact with DNA to fold into a G-quadruplex structure. The repeat RNA forms RNA foci that sequester RNA-binding proteins in the nucleus of vulnerable neurons. (3) Gain-of-function caused by dipeptide repeat proteins (DPRs): In a repeat-associated non-ATG-initiated manner, the repeat RNA is translated into DPRs that form toxic aggregates in residual neurons.

Clinical Features

Genetics

The G4C2 HRE in the noncoding region of the C9ORF72 gene contributes to ~25.1% of familial FTD and 37.6% of familial ALS world-wide [30,31,32]. Strikingly, in the Finnish population, C9ORF72 repeat expansion accounts for up to 46.0% of familial ALS and 21.1% of sporadic ALS [4], but the frequency of C9ORF72 repeat expansion is extremely low in Asian populations [33,34,35]. In recent years, several studies have comprehensively described the clinical and pathological features of C9ORF72 patients who exhibit wide variation of age at onset and disease duration [36,37,38,39,40,41,42,43,44]. Most bulbar-onset C9ORF72 patients exhibit symptomatic heterogeneity [38, 39, 43], including behavioral-variant FTD, progressive non-fluent aphasia, motor neuron disease, ALS, and mild psychosis and anxiety symptoms [41,42,43]. The ALS patients with C9ORF72 expansion have shorter survival after disease onset than those without C9ORF72 expansion [45].

Pathological Features

Neuronal Deficiency

Within a cohort of patients with behavioral variant FTD, compared with patients with mutation in microtubule-associated protein tau (MAPT) and progranulin (GRN) that display specific anteromedial temporal atrophy [46, 47] and temporoparietal atrophy [48, 49], respectively, the patients with HRE in C9ORF72 show a widespread pattern of grey matter loss in the cerebellum and spinal cord, and the extramotor frontal lobes, temporal lobes, and hippocampus, as well as the basal ganglia and occipital lobes [37,38,39,40,41,42,43,44, 50]. Notably, an atrophied cerebellum, which is a characteristic pathology of c9ALS/FTD, is not presented in brain of FTD cases without C9ORF72 expansion [51, 52], suggesting a specific role of C9ORF72 in cerebellar pathogenesis. Several other studies using magnetic resonance imaging (MRI) and unbiased voxel-based morphometric analysis have also found broad brain atrophy in C9ORF72 patients [41, 44], thereby accounting for the clinical heterogeneity of these patients.

Pathological Inclusions

One specific pathological feature of c9ALS/FTD is the presence of G4C2 nuclear RNA foci in both sense and antisense forms [3, 8]; the other is cytoplasmic inclusions of RNA-translated DPRs [9,10,11,12, 53]. Notably, intranuclear inclusions of DPRs that co-localize with nucleoli have been identified in the frontal cortex of C9ORF72 patients [54, 55]. The expression of the five DPRs is predominantly in the form of poly-GA, to a lesser extent poly-GP and poly-GR, compared to poly-PA and poly-PR [53]. However, only slight clinico-pathological correlations of poly-GA, but not other DPRs, have been reported in c9ALS/FTD [53]. It is still unclear how much other DPRs contribute to the pathogenesis, especially arginine-rich poly-GR/PR.

TDP-43 is a major component in the pathological inclusions in the ALS and FTD brain [56,57,58]. FTD-TDP is classified into four types according to the heterogeneity of TDP-43 inclusions: cytoplasmic inclusions, short neurites in the upper cortical layers (Type A), round TDP-43 inclusions throughout the cortex (Type B), long dystrophic neurites (Type C), and intranuclear inclusions (Type D) [59]. Broad inclusions of TDP-43 have also been identified in the vulnerable neurons of c9ALS/FTD, mainly containing type A, type B, or a combination of both [59,60,61,62]. A series of studies have reported that p62-positive, TDP-43-negative cytoplasmic inclusions are present in the cerebellar granular cells of c9ALS/FTD patients, while such inclusions are not found in non-C9ORF72 mutant individuals [40, 42, 53, 61, 63, 64], indicating a specific role of HRE in the protein degradation pathway. The clinical and pathological features of c9ALS/FTD described elsewhere are summarized in Table 1.

Table 1 Clinical and pathological features of frontotemporal dementia and amyotrophic lateral sclerosis with C9ORF72 gene mutation

In recent years, impaired nucleocytoplasmic transport has been identified as a common pathological process in C9ORF72 expansion-induced neurodegeneration [65,66,67,68]. Using large-scale unbiased genetic screening, several genes in the nucleocytoplasmic transport process have been identified as major hints in G4C2-expressing Drosophila models [66]. Consistently, RanGAP1, a key regulator of nucleocytoplasmic transport [69, 70], is abnormally distributed in the cortex of the G4C2 mouse model and C9ORF72 ALS patients [67].

Mouse Models with Hexanucleotide Repeat Expansion

It has been reported that overexpression of HRE causes obvious cellular toxicity in cell cultures [18, 19, 23], G4C2 Drosophila models [71,72,73], and a G4C2 zebrafish model [20]. In addition, Petrucelli and colleagues have developed two G4C2-expressing mouse models using an AAV-packaged hexanucleotide expansion with 66 or 149 repeats [74, 75]. In the 66-repeat model, the mice show evident expression of intranuclear sense RNA foci and DPRs in the central nervous system (CNS), accompanied by motor dysfunction and anxiety-like behaviors, as well as cortical neuronal deficiency [75]. Strikingly, nuclear and cytoplasmic phosphorylated TDP-43 inclusions have been observed in the cortex and hippocampus of (G4C2)66 mice [75]. The 149-repeat mice show similar phenotypes, while antisense RNA foci, in amounts from 10% to 20%, have been found in the hippocampus, cortex, and cerebellum. In addition, the antisense DPRs poly-PA and poly-PR have been detected in the cortex of 149-repeat mice at 3 months of age [74]. Mis-localization of RanGAP1 and aggregation of stress granule-associated proteins have also been reported in the cortex and hippocampus of 149-repeat mice [74], suggesting that overexpression of HRE is enough to model the neuropathological changes of c9ALS/FTD.

To identify the pathological effects of HRE at lower expression levels, several groups have developed bacterial artificial chromosome (BAC) transgenic mouse models [29, 76,77,78]. Despite RNA foci and a subset of DPRs in mice containing 500 or 100–1000 repeats of the G4C2 sequence, two BAC models show no behavioral deficiency or neurodegeneration [76, 77]. However, two other transgenic mice show a clear phenotype [29, 78]. One BAC mouse embedded with a patient-derived C9ORF72 gene harbor either 110 or 450 repeats of G4C2, and the 450-repeat mice display cognitive impairment but not motor deficits, accompanied by size- and dose-dependent expression of RNA foci and DPRs in the CNS [29]. Other transgenic mouse models that have the full-length human C9ORF72 gene with ~30 and ~500 repeats show obvious gait abnormalities, anxiety-like behavior, and decreased survival, as well as widespread neurodegeneration and TDP-43 pathology [78]. Notably, the antisense RNA foci preferentially accumulate in the c9ALS/FTD-vulnerable cell populations [11, 79, 80]. As both of the latter models have higher expression of human C9ORF72 mRNA levels than those in the former models [29, 76,77,78], the expression levels of the human C9ORF72 gene in mice and the antisense RNA foci and related DPRs may contribute to the phenotypes.

Disease Models with Dipeptide Repeat Proteins

The G4C2 expansion can be translated into five DPRs in a repeat-associated non -ATG-initiated manner (poly-GA, poly-GP, poly-GR, poly-PA, and poly-PR) [7, 9,10,11]. Among these, poly-GP and poly-GA display higher expression in the c9ALS/FTD brain than the other three DPRs [53]. As poly-GP and poly-PA have been reported to have no cytotoxicity [22, 23, 81], most studies focus on the roles of poly-GA, poly-GR, and poly-PR in neurodegeneration.

Role of Poly-GA in Cytotoxicity

Cell-Culture Models of Poly-GA

Poly-GA forms cellular inclusions and co-localizes with ubiquitin and p62 in GFP-GA50-transfected cultured cells [82]. Poly-GA overexpression induces cytotoxicity, including an increase in caspase-3-positive cells and release of lactate dehydrogenase, which is also found in cultured primary cortical neurons overexpressing poly-GA [82]. In primary hippocampal neurons, a long repeat length of poly-GA (149 repeats) forms p62-positive inclusions and induces dendrite loss [83]. In addition, GA50 and GA100 cause slight neuronal toxicity in Neuro-2a [81] and NSC-34 cells [84]. However, some studies have reported that 30-repeat lengths of GA have no significant toxicity in NSC-34 and HEK293 cells [24], and that both GA and PA are not neurotoxic even at a length of 200 repeats [23]. To date, it is unclear whether the toxicity of poly-GA depends on the expression levels or cell types.

Drosophila Models with Poly-GA

To determine the role of poly-GA in neurodegeneration, several groups have established poly-GA Drosophila models with different repeat lengths [22, 23, 81]. The expression of GA50 and PA50 in Drosophila does not induce neurotoxicity, which is consistent with the previous finding in primary hippocampal neurons [23]. Moreover, GA100 and PA100 have no effects on the egg-to-adult viability of Drosophila at different temperatures, although GA100 causes a late-onset decrease in survival [22], suggesting a mild toxicity of poly-GA. In addition, overexpression of GA80 does not induce degeneration in Drosophila eyes or wing margins [85].

Zebrafish and Chicken Models with Poly-GA

Although poly-GA in Drosophila does not show cellular toxicity, zebrafishes expressing GA80 show a strong pericardial edema phenotype and dramatically decreased circulation, accompanied by an accumulation of red blood cells, and these phenotypes can be rescued by interfering with the expression of GA80, indicating a toxic property of poly-GA in zebrafish [86]. In addition, poly-GA shows the highest toxicity to neurons in the spinal cord of transgenic chickens as compared to other DPRs [87]. Poly-GA sequesters other DPRs to its aggregates and overexpression of poly-PA inhibits poly-GA aggregation [87], suggesting a role of GA in aggregate formation and the influence of other DPRs on GA aggregation.

Mouse Models with Poly-GA

Given the limitations of cultured cells and Drosophila models, several groups have constructed poly-GA transgenic mouse models [88,89,90]. In poly-GA transgenic mice, in which the expression of poly-GA is controlled by the Thy1 promoter, poly-GA is mainly distributed in the spinal cord and brainstem [88]. The mice show co-aggregation of poly-GA with p62, Rad23b, and Mlf2; this also occurs in c9ALS/FTD patients [88]. The animals show some behavioral changes, including motor imbalance and hypoactivity, but no defects in muscle strength or spatial memory [88]. Moreover, the mice have no significant motor neuron deficiency in the spinal cord, although microglial activation is present there [88]. With intracerebroventricular injection of adeno-associated virus (AAV)-packaged GFP-GA50, the mice present motor deficits, brain atrophy, neurodegeneration, and neuroinflammation at post-natal day 0 [89]. Poly-GA-induced motor dysfunction has been confirmed using a similar method [90]. Thus, overexpressed poly-GA has neurotoxic effects and induces motor defects in vertebrates, although its effects in cultured cells or Drosophila remain controversial.

Roles of Poly-GR and Poly-PR in Cytotoxicity

Cellular Models with Poly-GR and Poly-PR

Within DPRs, poly-GR and poly-PR show strong toxicity in both cellular and animal models, in which several methods have been applied to avoid the toxic effects caused by G4C2 RNA foci. In vitro analyses show that synthetic GR20 or PR20 forms nucleolar inclusions and is significantly cytotoxic in U2OS cells and primary astrocytes [21]. In primary cortical neurons, overexpression of PR50 has high neurotoxicity, while poly-GR-induced cell death is repeat-length dependent [23]. Moreover, the cellular toxicity of arginine-rich poly-GR and poly-PR has been reported in multiple cell lines, including Neuro-2a [81], NSC-34 [24, 84], SH-SY5Y [91], and primary cortical neurons [92]. In addition, overexpression of PR100 [93] and GR80 [94] in induced pluripotent stem cell (iPSC)-derived neurons induces significant neuron deficiency and a DNA damage response, further indicating the high toxicity of arginine-rich poly-GR and poly-PR.

Drosophila Models with Poly-GR and Poly-PR

To address whether DPR-induced neurotoxicity depends on repeat RNA, Mizielinska and colleagues have constructed Drosophila models in which they use two strategies: (1) a 6 base-pair interruption that contains a stop codon in both the sense and antisense directions is inserted in every G4C2 repeat, which construct produces “RNA-only” repeats without DPR expression; and (2) using alternative codons that encode the same amino-acids but disrupt the G4C2 repeat, which construct expresses “DPR-only” without G4C2 repeats [22]. In Drosophila, GR100 and PR100 exhibit strong toxicity, leading to severe neurodegeneration and reduced survival, while GA100 causes late-onset disease [22]. However, “RNA-only” repeats do not induce neurodegeneration and disease phenotypes, suggesting that DPRs, especially poly-GR and poly-PR, are major factors in C9ORF72-induced pathogenesis [22]. Expression of PR50 in motor neurons induces developmental failure in Drosophila, despite a normal body morphology [23]. In addition to the neuronal toxicity of arginine-rich DPRs, GR80 exhibits non-neuronal cellular toxicity in Drosophila, such as wing margin defects [85], suggesting that PR and GR are highly toxic to both neuronal and non-neuronal cells in Drosophila.

Mouse Models with Poly-GR and Poly-PR

Several C9ORF72 BAC transgenic mice have been established to identify the role of the C9ORF72 gene in disease [29, 76,77,78]. Nuclear RNA foci and DPRs have been found in the brain of BAC transgenic mice that have no motor neuron degeneration, although one BAC mouse with an FVB/NJ background shows decreased survival and motor deficits [29, 76,77,78]. Thus, transgenic mice with long hexanucleotide repeats in the C9ORF72 gene do not present disease phenotypes, or have only mild phenotypes. Due to the high toxicity of poly-GR and poly-PR in cultured cells, iPSC-derived neurons, and Drosophila models, arginine-rich DPR mouse models have been created [95,96,97]. With intracerebroventricular injections of AAV1 that expresses GFP-GR100, the GFP-GR100 is mainly expressed in the CNS, with a cytoplasmic and diffuse distribution, while there is little expression in the spinal cord [95]. The GR100 mice display progressive motor deficits and memory loss accompanied by age-dependent cortical and hippocampal neurodegeneration. In addition, glial activation occurs at 1.5 months of age [95]. In another mouse model, the spatial and temporal expression of poly-GR is controlled by Tet expression systems, and the expression of GR80 is mainly distributed in the frontal cortex relative to other cortical regions, with diffuse distribution in the cytoplasm and nucleus of 95% of neurons [96]. In addition, the transgenic mice display age-dependent social behavioral deficits, impaired synaptic transmission, cortical neuronal loss, and microglial activation, but no changes in body weight, locomotor activity, and working memory [96]. Thus studies suggest that a long repeat length of poly-GR has a diffuse cytoplasmic distribution and is able to induce severe neurodegeneration and related behavioral deficits in vivo.

The poly-PR that shows the highest neurotoxicity in cellular and Drosophila models has toxic effects in poly-PR AAV-infected mice and transgenic mice [55, 97]. AAV-infected GFP-PR50 mice show behavioral deficiencies and neurodegeneration at an early stage [55]. Moreover, the overexpression of poly-PR in neurons is highly toxic; up to 60% of GFP-PR50-expressing mice die by 4 weeks of age [55]. Strikingly, in this AAV-mediated poly-PR expressing mouse, besides a nucleolar distribution, poly-PR mainly exhibits heterochromatic localization, which elicits aberrant post-translational modifications of histone H3 [55]. In the poly-PR transgenic mice with intermediate repeat lengths of poly-PR, the neuronal expression of GFP-PR28 driven by Cre recombinase under control of the Thy1 promoter is highly toxic [97]. Homozygous transgenic mice develop body weight loss and premature death, similar to GFP-PR50 AAV mice [97]. Heterozygous mice exhibit age-dependent motor dysfunction, decreased survival time, motor-related neuronal deficiency, and neuroinflammation [97]. Unlike the heterochromatic localization of GFP-PR50 in AAV mice [55], GFP-PR28 is mainly distributed in the nucleolus of neurons, as well as a diffuse cytoplasmic distribution in lumbar spinal motor neurons [97]. It is still unknown whether the difference of repeat length in poly-PR between these two models leads to the difference in poly-PR distribution. The animal models of C9ORF72 with either HRE or dipeptide repeat proteins are summarized in Table 2.

Table 2 Transgenic mouse models expressing hexanucleotide repeat expansion and dipeptide repeat proteins

In addition to the difference in cellular localization between poly-GA and arginine-rich DPRs [24], poly-GR and -PR are more neurotoxic than poly-GA in cultured cells [24] and Drosophila [22]. Consistent with this, poly-GA mainly influences the cytoplasmic ubiquitin-proteasome system [98], while poly-GR and -PR have effects on nuclear processes. Although all these DPR mice display serious behavioral deficits and neurodegeneration [55, 95,96,97], they do not develop TDP-43 pathology that is a general neuropathological feature of c9ALS/FTD [56,57,58]. These data also suggest that the pathogenesis in patients may be complicated. Synergic effects, such as cooperativity between gain- and loss-of-function mechanisms, may be essential for the induction of pathological events.

Pathological Mechanisms

Since the discovery of the non-ATG-initiated translation of C9ORF72 repeat expansions, a variety of pathological mechanisms linked to DPRs have been identified, ranging from DNA processes to RNA processing to protein translation (Fig. 2).

Fig. 2
figure 2

Potential mechanisms linked to DPR overexpression. DPR overexpression causes neuronal deficiency through mechanisms that range from DNA processes to RNA processing to protein translation. DPRs cause a DNA damage response and poly-PR induces abnormal histone methylation and dysfunction of heterochromatin. Defects in RNA processing such as nonsense-mediated RNA decay, RNA splicing, and ribosomal RNA processing are linked to arginine-rich DPRs. Moreover, these DPRs contribute to dysfunction of protein homeostasis through nucleocytoplasmic transport, the unfolding protein response (ER-stress), and translation. The diagram in the lower right corner summarizes the biological processes related to poly-PR-binding proteins, which mainly include the nucleolus, RNA splicing, the nuclear pore complex, and the ribosome. rDNA, ribosomal DNA; rRNA, ribosomal RNA.

Shedding Light on DNA Processes

DNA Damage Response (DDR)

Previous studies have demonstrated that an impaired DDR is essential for neuronal deficiency in neurodegenerative diseases [99, 100]. Individuals with mutations in XRCC1 that encodes a scaffold protein that is involved in DNA single-strand break repair, present ocular motor apraxia, axon neuropathy, and progressive cerebellar ataxia [101]. Mutations in another ALS/FTD-related gene FUS (Fused-in-Sarcoma) lead to a DDR and DNA repair dysfunction [102,103,104,105,106]. Impairment of the DDR can also be induced by mutations in TDP-43 [107, 108], ATM [109], APTX [110], and other genes, suggesting a role of the DDR in neurodegeneration.

An essential role of the DDR in the pathogenesis of c9ALS/FTD has been documented in several studies that demonstrated genomic instability as a key event in C9ORF72-linked neurodegeneration. In iPSC-derived C9ORF72 motor neurons, there is increased expression of γH2AX, a marker of DNA double-strand breaks, and increased tail length in the comet assay, suggesting the presence of DNA damage in c9ALS/FTD [94]. Moreover, the expression of γH2AX is also increased in the spinal motor neurons of C9ORF72 patients [90, 93]. In primary cortical neurons and the SH-SY5Y cell line, overexpression of poly-GR and poly-PR induces an accumulation of γH2AX foci and an increase of phosphorylated ATM, the main kinase for DNA repair [93]. Moreover, in poly-GA and repeat RNA overexpression cellular models, the levels of DNA–RNA hybrids (R-loops), a three-stranded nucleic acid structure produced during the transcription of repeat sequences, are increased, which can lead to genome instability [90]. Furthermore, the ATM signaling pathway is impaired in poly-GA transgenic mice, suggesting that defects in the DNA repair machinery are associated with DPR and RNA repeat-induced DNA damage [90]. DNA damage is increased in iPSC-derived C9ORF72 motor neurons [94]. Overexpression of poly-GR but not poly-GA in control iPSC-derived motor neurons induces DNA damage [94], further suggesting a role of DPRs in DNA damage.

Heterochromatin and Histone Methylation

Poly-PR is localized to heterochromatin in the cortex of transgenic mice as well as in c9ALS/FTD patients where it causes abnormal histone methylation [55]. In addition, poly-PR reduces the expression levels of HP1α and disrupts the phase separation of HP1α, leading to lamin invaginations and double-stranded RNA accumulation [55]. Meanwhile, repetitive elements that make up a large portion of heterochromatin are broadly upregulated in the brains of c9ALS/FTD patients and mice overexpressing poly-PR [55, 111].The upregulation of abnormal repetitive elements and accumulation of double-stranded RNA induce neurodegeneration. Thus, the abnormality of histone methylation and the dysfunction of heterochromatin and DNA components induced by poly-PR may contribute to the neurodegeneration in c9ALS/FTD.

RNA Processing

Nonsense-Mediated RNA Decay (NMD)

NMD, a cellular RNA degradation system that is activated in response to stress, rapidly degrades mRNA containing premature termination codons to prevent the translation of defective proteins [112, 113]. UPF1–3 are master factors for NMD activation in which UPFs interact with each other, the ribosome, and multiple mRNA decay factors [112, 113]. NMD can be inhibited, which is indicated by an overlap of upregulated NMD substrate genes in c9ALS/FTD-derived iPSCs and UPF3B-/-lymphocytes [91]. Numerous NMD substrate genes accumulate in PR36 Drosophila [91]. In addition, there are significant overlaps in the accumulated NMD genes in flies overexpressing PR36 and those deficient in UPF1 or UPF2 [91]. Furthermore, the NMD substrate genes Sin3A, Gadd45, Xrp1, and Arc1 do not accumulate in flies expressing α-synuclein, Htt-128Q, or human FUS, although these genes do accumulate in flies overexpressing GR36 or PR36 or with UPF1 knockdown [91]. In contrast, overexpression of UPF1 significantly inhibits the neurotoxicity induced by GR36 or PR36.

Although the primary cause of NMD by DPRs remains unknown, a reactivation of NMD significantly ameliorates the neurotoxicity in flies expressing DPRs, indicating that impairment of NMD contributes to the pathology of C9ORF72.

RNA Splicing

Pre-mRNA splicing and the biogenesis of rRNA are significantly impaired in cultured human astrocytes expressing poly-PR/GR [21]. Notably, FUS and TDP-43, two other ALS-related gene products, have been shown to cause a pre-mRNA splicing dysfunction and reduce downstream gene expression [114,115,116,117]. Another study using unbiased quantitative mass spectrometry demonstrated that poly-GR/PR binds to U2 snRNPs (small nuclear ribonucleoproteins) and inhibits spliceosome assembly [118]. Moreover, bioinformatics analysis showed that U2-dependent exons are mis-spliced in the cortex and cerebellum of C9ORF72 patients [118]. Previous results have revealed that arginine-rich poly-PR and poly-GR, but not poly-GA, bind to proteins that contain low complexity domains [81, 119]. Consistently, a number of binding proteins with low complexity domains, including several U2 snRNPs, have also been detected [118]. These data suggest that U2 snRNPs sequestered by arginine-rich DPRs contribute to the blocked alternative splicing in C9ORF72 patients.

Protein Homeostasis

Proteinopathies such as ALS mainly exhibit protein inclusions with impaired quality control systems in cells [13, 14, 120]. Poly-GR and poly-PR interact with proteins that contain low complexity domains and disturb the phase separation of membrane-less organelles, such as the nucleolus, the nuclear pore complex, and stress granules [81, 119, 121, 122]. The function of poly-GR and poly-PR in the nuclear pore complex and stress granules has been systematically described in other reviews [123, 124]. Here, we mainly focus on the role of poly-GR and poly-PR in the protein quality control system, including the unfolded protein response and translation.

Unfolded Protein Response (UPR)

Using unbiased CRISPR-Cas9 screens, TMX2, an endoplasmic reticulum (ER)-resident transmembrane thioredoxin protein, has been identified as a major modifier of neurotoxicity in primary cortical neurons infected with poly-PR [92]. Moreover, RNA-sequencing analysis has demonstrated that the genes in the ER-stress/UPR pathway are significantly upregulated, including Atf4, Bbc3, and Chac1. Knockdown of TMX2 ameliorates the neurotoxicity caused by poly-PR overexpression in primary cortical neurons and in motor neurons derived from C9ORF72 patients [92]. Notably, the ER-stress-related genes Chac1 and Atf5 are upregulated in the cerebellum of poly-PR transgenic mice at 2 months of age, before motor dysfunction occurs, suggesting that ER-stress is an early event in the pathogenesis of poly-PR [97, 125]. In addition, poly-GA overexpression induces pathological inclusions, neurotoxicity, and ER-stress [82]. Pharmacological inhibitors of ER-stress decrease the expression of ER-stress markers and alleviate the neurotoxicity caused by poly-GA [62]. Besides c9ALS/FTD, SOD1-mutated ALS and other neurodegenerative diseases also show the presence of ER-stress in the CNS of patients and animal models [126,127,128]. Interestingly, using techniques that combine translational ribosome affinity purification and high-throughput RNA sequencing, a cascade of cell type-specific dysregulated processes in SOD1-mutated mice has been identified, showing that the UPR starts within neurons, followed by metabolic and inflammatory gene changes in astrocytes and membrane protein gene alteration in oligodendrocytes [126]. Thus, ER-stress is an early reaction occurring specifically in motor neurons in ALS, and maybe a potential target in ALS therapy.

One interesting but still unanswered question is how poly-PR induces the UPR, as arginine-rich poly-PR is localized in the nucleolus, not the ER. Recently, the nucleolus has been identified as a phase-separated protein control compartment [129]. Under stress, nucleoplasmic proteins are transported to the nucleolus, where Hsp70 mediates refolding, preventing the irreversible aggregation of misfolded proteins [129]. In cells expressing poly-PR, disaggregation of misfolded proteins is inhibited and mobile fractions of liquid-separated proteins are significantly reduced, suggesting an inhibition of nucleolar quality control [129]. Thus, misfolded protein aggregation in the nucleolus, which is caused by poly-PR overexpression, may contribute to the misfolded protein response in c9ALS/FTD. However, it remains unknown whether the nucleolus transmits a stress signal to the ER, thereby leading to ER-stress.

Translation

It is well documented that arginine-rich poly-GR and poly-PR interact with ribosomal proteins, and proteins involved in translation [81, 94, 130, 131]. An inhibition of global translation is consistently found in cultured cells expressing poly-PR [132], primary rat cortical neurons [130], adult Drosophila [131], and human iPSC-derived motor neurons [131]. Overexpression of the translation initiation factor eIF1A rescues the translational defects and neurotoxicity caused by poly-PR [131]. In addition, overexpression of poly-GR100 in mice causes significant neurodegeneration with a dysregulation of genes involved in the ribosomal pathway [95]. The impaired canonical translation occurs in neurons of the GR-expressing mouse model; besides, the non-canonical translation (repeat-associated non-ATG-initiated translation, RAN translation) is also found [95], which argues against previous studies suggesting that DPR-induced ER-stress contributes to the selective activation of RAN translation in vitro [133,134,135]. One reason for the difference may be that the expression of DPR-induced ER-stress thereby activates RAN translation at an early stage, while RAN translation is inhibited under chronic stress.

Therapeutic Advances

Due to the high frequency of C9ORF72 mutation and the urgent requirement for clinical treatment in ALS and FTD, a wide variety of efforts have been put into the identification of promising therapeutic approaches, from DNA to RNA to DPR protein processes.

DNA Processes

DNA methylation plays a pivotal role in regulating the expression of downstream genes. Hypermethylation has been found in the promoter of the C9ORF72 gene, which contains G4C2 repeat expansions, leading to reduced C9orf72 mRNA levels [136]. Neurons from carriers of hypermethylated C9ORF72 repeat expansions display reduced RNA foci and DPR aggregations, suggesting a protective role of hypermethylation in C9ORF72-related pathology [136]. Similarly, genetic manipulation using CRISPR-Cas9 technology that targets the promoter region of the C9ORF72 gene decreases the levels of C9ORF72 and DPR proteins, and ameliorates neurotoxicity in iPSC-derived neurons [137]. Despite the promising effects, further efforts are needed to avoid the adverse effects of downregulation of C9ORF72 protein that is essential for immune function [16] and protein trafficking [28].

RNA Processing

Using unbiased large-scale screens, a number of possible remedies for c9ALS/FTD have focused on transcriptional regulation, targeting the transcriptional regulator PAF1 [138], transcriptional elongation factor SPT4 [139], and AFF2/FMR2 [140]. In addition, antisense oligonucleotides (ASOs) that are single-stranded show clear alleviation of nuclear RNA foci and neurotoxicity [19, 141, 142]. Given the essential role of normal C9ORF72 protein, optimized ASOs exhibiting no C9ORF72 RNA reduction have been designed and are successful in reducing the number of sense RNA foci and the expression of DPRs in transgenic mouse models [29], indicating a hopeful outlook for ASOs in the treatment of ALS and FTD. In addition, RNA-targeting Cas9 is also effective in eliminating the G4C2 repeat RNA foci [143].

Protein Homeostasis

As a result of the high toxicity of DPRs, several strategies have been designed to reduce their levels. RPS25 [144], a small ribosomal protein subunit, and DDX3X [145], an RNA helicase, suppress the translation of DPRs and improve the survival of patient-derived induced motor neurons. Moreover, specific antibodies targeting poly-GA significantly reduce GA protein levels, ameliorate neuronal deficiency, and improve motor dysfunction in transgenic mice [146, 147]. Therefore, selective targeting of DPR expression may also be an alternative therapeutic approach.

Conclusions and Perspectives

Despite the broad achievements in exploring the role of DPRs in disease, several key questions remain unanswered. The DPRs form neuronal cytoplasmic inclusions that are widespread in c9ALS/FTD patients, while overexpression of DPRs in vivo, especially arginine-rich poly-PR, only show intranuclear inclusions in selective neurons. It remains unknown whether the repeat length of poly-PR, or the localization of other DPRs that may interact with poly-PR, contributes to the cellular distribution. Similarly, the distribution of TDP-43 is inconsistent with the pathological features and a priority is to identify which factors are essential for TDP-43 inclusions, RNA foci, or aggregated DPRs. It will also be interesting to determine whether DPR inclusions in muscle contribute to clinical pathology [148, 149].