Skip to main content

Nucleo–cytoplasmic transport defects and protein aggregates in neurodegeneration

Abstract

In the ongoing process of uncovering molecular abnormalities in neurodegenerative diseases characterized by toxic protein aggregates, nucleo-cytoplasmic transport defects have an emerging role. Several pieces of evidence suggest a link between neuronal protein inclusions and nuclear pore complex (NPC) damage. These processes lead to oxidative stress, inefficient transcription, and aberrant DNA/RNA maintenance. The clinical and neuropathological spectrum of NPC defects is broad, ranging from physiological aging to a suite of neurodegenerative diseases. A better understanding of the shared pathways among these conditions may represent a significant step toward dissecting their underlying molecular mechanisms, opening the way to a real possibility of identifying common therapeutic targets.

Background

Neuronal protein aggregates are a characteristic neuropathological hallmark of neurodegenerative disorders such as amyotrophic lateral sclerosis (ALS), frontotemporal dementia (FTD), Alzheimer’s disease (AD), Huntington’s disease (HD), polyglutamine expansion related ataxias, and Parkinson’s disease (PD) [1]. One possible consequence of these aggregates is interference with the proper functioning of the highly conserved nucleo–cytoplasmic transport (NCT) mechanism in the cell, which ensures the transport of nucleic acids and proteins across the nuclear membrane. This pathway is fundamental for long-lived cells such as neurons [2, 3], and emerging evidence links damage to the nuclear membrane and nuclear pores to neurodegenerative diseases and physiological neuronal aging [4].

A properly functional NCT is fundamental for neurons to allow transcription factors to enter into the nucleus [5]. One essential NCT role is to activate regulatory elements necessary for adaptation and response to external stimuli, especially those implicated in neuronal plasticity [6]. Moreover, in non-dividing post-mitotic cells such as neurons, proteins involved in DNA maintenance and repair reach the nucleus through this mechanism [7, 8]. In addition, all ribonucleic proteins involved in processes of RNA maturation, stability, splicing, and export require this transport for localization to the nucleus [9].

The cell relies on an essential, conserved, and dynamic structure called Nuclear Pore Complex (NPC) to accomplish a correct NCT [5]. Small molecules (molecular weight < 40 kDa, diameter < 5 nm) can relatively freely diffuse through the NPC [5]. In contrast, higher molecular-weight proteins depend on the highly regulated and very specific mechanisms of the NCT machinery [5].

NPCs are large cellular structures that span the nuclear envelope [10] and control exchanges between the cytoplasm and nucleus. The NPC is constituted by proteins collectively referred to as nucleoporins (Nups) [11] that have different roles in regulating NPC functions and transport of matter, energy, and information between the nucleus and cytoplasm [12]. In addition, the NPC is involved in cell cycle regulation, chromatin organization, and gene activation [6, 13].

Despite the relatively large size of NPCs, most proteins and RNAs cannot freely diffuse through them. Indeed, the passage through NPC is a highly coordinated and selective process, mediated by a family of soluble receptors called karyopherins (Kaps), also known as importins/exportins [14]. The small Ras-related nuclear protein (Ran) plays a central role in this transport and regulates interactions between Kaps and their cargos on either side of the NPC [15]. Ran is the only known member of the Ras superfamily of small GTP-binding proteins that is localized principally inside the nucleus [16, 17]. A gradient of RanGDP/GTP across the nuclear membrane is essential in establishing the directionality of the transport. To maintain the gradient, RanGTP is hydrolyzed to RanGDP on the cytoplasmic side and RanGDP is converted to RanGTP in the nucleus [17].

The nuclei of human cells have several thousand NPCs, with a structure that seems to have been highly conserved, underlining their importance for cells [18]. Nevertheless, despite this high degree of structural conservation of the NPC complex, two crucial time points in evolution from prokaryotes to humans have been suggested in NPC development, implying that the nuclear pore is still highly adaptive and flexible at the sequence level [19].

Neurons seem particularly sensitive to the damage of these structures, as demonstrated by the exclusively neurodegenerative consequence of many diseases due to damage of the NPC function and the NCT. So far, several reasons for this vulnerability have been studied. One clue is the inability of neurons to undergo mitosis, the usual process of cellular renewal [20]. Through DNA replication, cells maintain physiological protection of the genome from exogenous damages. To preserve genomic integrity, there are at least four active DNA repair pathways in nervous system each corresponding to a particular type of DNA lesions [21]. Genome instability can appear when the accumulation of DNA damage exceeds a neuron’s repair capacity or the DNA repair machinery is defective [20]. Thus, transport across the nuclear membrane represents a unique possibility for neurons to guarantee DNA integrity, and genome instability is a major factor in neuronal aging [22, 23]. For example, somatic single nucleotide variants have been reported to be in excess in neurons from people affected by early-onset degeneration with DNA repair gene mutations [24].

Cellular oxidative stress is one of the primary sources of DNA damage, mostly in the brain due to its high demand for energy and increased radical oxygen species (ROS) formation [25]. This process, in addition to direct effects on DNA stability, is increasingly reported to affect NCT [26, 27]. At least four mechanisms have been associated with oxidative stress: reduced Ran GDP/GTP ratio, mislocalization of Nups, altered functions (binding, docking) of importins/exportins, and decreased integrity of the nuclear membrane [28].

RNA transport defects are an additional sign of neuronal sensitivity to NCT impairment, and of course, the export of RNAs from the nucleus to the cytoplasm is key to gene expression [29]. Moreover, a tRNA retrograde transport between cytoplasm and nucleus has recently been proposed as part of the cellular response to oxidative stress [30]. Among neurons, proper RNA shuttling function appears even more relevant for cells with a high transcriptional activity such as Purkinje cells and motor neurons [31,32,33]. All of these NCT defects that significantly affect DNA and RNA are related to aging and are even more relevant in neurodegeneration [26]. RNA-binding proteins mutated in specific neurodegenerative disorders (TDP-43, FUS, and hnRNPA1) can alter the dynamics of membrane-less organelles such as stress granules, or the more recently identified liquid droplets, and accelerate fibrillization in neurons, resulting in the formation of pathological amyloid-like fibrils that deposit in the cell bodies and neuropil [34,35,36].

Neurodegenerative disorders display a unique pathological hallmark arising from specific gene mutations. This hallmark is protein aggregates, which are known to damage NPCs [34, 37, 38], and evidence indicates that protein and RNA aggregates may interfere directly with specific Nups [34, 37]. In some neurons, Nups mislocalization and altered NPC function could lead to an even greater increase in aggregate accumulation and genome stress due to chromatin and DNA/RNA impairment [39, 40]. Further complicating this picture, the pathogenic process includes aberrant transcript maturation, transport, and translation [41] (Fig. 1).

Fig. 1
figure 1

Representation of the nucleus and NPCs with Nups, protein/RNA aggregates in several neurodegenerative diseases. ALS/FTD: - Damaged NPC and specific Nups involved in the presence of TDP-43 cytoplasmic aggregates and impaired TDP-43 nuclear import. - Impaired FUS import in the presence of aggregates containing FUS with the importin TNPO1 or alone. - Impaired RNA export in C9ORF72 mutations with DPRs and TDP-43 protein aggregates formation. - Altered Ran and RCC1 nucleo-cytoplasmic distribution; C9ORF72 toxicity is increased by GLE1. - EXOSC3 dysfunction. - Altered distribution of importin-α and -β in SOD1 mutation with Nup62 impairment. HD: HTT physiological transport across the NPC through the interaction of NES with TPR and XPO1. Aberrant shuttling of RAN proteins and MAP2 due to PolyQ HTT affecting the NPC. Intranuclear aggregates of PolyQ HTT sequestering Nup62, Nup88, GLE1, and RanGAP1. AD: Phospho-tau aggregates induce NPC damage and accumulation of NTF2 and Nup98 in the cytoplasm; Nup98 loss, in turn, may facilitate tau aggregation. PD: Cytoplasmic aggregates and intranuclear alpha-synuclein in Parkinson’s disease; pCREB aggregates and nuclear accumulation of NFkB are associated with NPC and Nup358 defects in PD

The neuronal inclusions detected in neurodegenerative forms have been linked to deficient mechanisms of degradation, especially proteasomal and autophagy processes [42].

Whether the pathological cascade, starting from the formation of the aggregates and the damage of the NPC, to the enhanced fibrillation, may represent a common paradigm underlying different neurodegenerative diseases is still unknown.

In summary, several lines of evidences are emerging to suggest a link between the formation of neuronal aggregates and the structural and functional damages of the NPC, as well as the NCT pathways. However, it is not clear yet whether NCT dysfunction act as an upstream common pathogenetic mechanism in the neurodegenerative process or a downstream event triggered by specific pathological aggregates of different neurological disorders. To give an overview that may facilitate the design of future studies, we review here the most relevant pieces of evidence of NCT impairment in neurodegeneration. We explore several neurodegenerative diseases, their pathogenic mechanisms, and genetic causes, highlighting the role of NPC and NCT as key factors in modulating the neurodegenerative process and also physiological aging.

Main text

NPC structure and function

The nucleus is the central and distinguishing organelle of eukaryotic cells, encompassed by a double membrane dynamic structure called the nuclear envelope [43]. This envelope consists of an outer membrane that is directly continuous with the rough endoplasmic reticulum and an inner membrane that contains a specific set of nuclear envelope transmembrane proteins [44]. Internal to the nuclear envelope is the nuclear lamina, a dense fibrillar network of intermediate filaments that surrounds the cellular genome [45]. These structures guarantee the specific eukaryotic compartmentalization that segregates the DNA from the cytoplasm. Accomplishing this function requires an accurate system providing proper communication and molecular transport among the different cellular compartments [46, 47], which is the role of NPCs.

Vertebrate NPCs are 70 nm protein channels spanning the nuclear envelope, with a cylindrical scaffold of 125 nm and internal diameter of 40 nm [48]. The channel connects the nucleus and the cytoplasm. NPC is the largest cellular protein structure at 125 MDa [12], consisting of more than 30 different proteins called Nups [49, 50]. When assembled, they form a cytoplasmic ring, spoke ring, and nuclear ring [12]. Eight filaments are attached to the rings at the nuclear and cytoplasmic sides [51,52,53]. On the nuclear side, 50 nm filaments are connected in a basket-like structure, while on the cytoplasmic side, the filaments are linked to the Nup214 complex [9, 48, 54, 55].

The NPC has a complex and highly regulated function, and most of the specific roles of single Nups are not well known. What is known is that a particular group of Nups in the central channel is fundamental to the selective barrier and substrate-specific transport role of NPCs [56, 57]. These Nups are characterized by phenylalanine-glycine (FG) domains and are anchored to the core scaffold through linker Nups [58]. Moreover, several Nups seem to exhibit a certain level of redundancy and functional overlap, forming an extremely dynamic barrier [56, 57].

With their intrinsic disordered FG domains, FG Nups form a dynamic filter that prevents passive diffusion of molecules through the NPC and allows for regulated transport of larger protein complexes of up to 40 nm [59, 60]. A single pore can contain 6 MDa of FG repeats, providing docking sites for import and export nuclear transport receptors (NTRs) that are crucial for selective passage [54, 61]. Molecules passing from the cytoplasm to the nucleus and vice versa must bear specific signaling sequences to interact with the NPC [54]. NPCs do not change from a defined “closed” to an “open” state during this passage, and the bond with the substrates consists of multiple weak contacts between the NTRs and the FG Nups assembly [58, 62, 63].

Among the many different NTRs, the largest group is the highly conserved family of proteins known as Kaps, consisting of more than 20 members (importins, exportins, and transportins) in humans [64]. Kaps can transiently interact with Nups and their FG domains to facilitate the shuttling of specific macromolecules through the NPC [64]. Motifs adjacent to FG repeats coordinate termination of transport and release of transporting complexes from the NPC [65].

Only a few other Kaps, such as importin-α, importin-β1, importin-β2, and chromosome region maintenance 1 (CRM1), have been well characterized [66, 67]. The same is true for the interactions among these proteins, with only a few well elucidated, such as the link between importin β1 and Nup153 at the nuclear face of NPCs [44, 68]. Many transport receptors, e.g., importin-β and NTF2, have hydrophobic binding sites on their surface for these FG-Nups [69, 70].

The NTRs bind short signaling peptides presented on their substrates for nuclear import and export. These signaling regions have been generally referred to as nuclear localization signals (NLSs) and nuclear export signals (NESs). The link established between the receptor and the shuttled substrate can be direct or mediated by additional adaptors, such as importin-α itself between importin-β and cargoes [71].

The processes of shuttling cargo–receptor complexes into the nucleus or into the cytoplasm do not directly use energy from ATP/GTP hydrolysis but depend strictly on the small GTPase Ran [72,73,74]. The nucleoporin Nup358 (RanBP2) has four domains that can bind Ran [72]. This member of the protein Ras superfamily exists in two nucleotide states, bound to nucleotide guanine triphosphate (RanGTP) or to guanine diphosphate (RanGDP) [73, 74]. The intrinsic GTPase activity of Ran is low, but the GTPase-activating protein RanGAP1 [75] catalyzes the shift between the two different forms, from cytosolic RanGTP to RanGDP [76, 77].

In contrast, the Ran guanidine exchange factor RCC1 (Regulator of chromosome condensation 1) in vertebrates can promote the exchange of guanine nucleotides by RanGDP to RanGTP in the presence of GTP [75]. This process leads to a higher nuclear concentration of RanGTP in the nucleus. A continuous supply RanGDP from the cytoplasm to the nucleus is accomplished by nuclear transport factor 2 (NTF2) [78]. RCC1 only can be found attached to nuclear chromatin to guarantee unequal distribution of GTPase Ran states, as the presence of a RanGDP/GTP gradient across the nuclear envelope gives directionality to the transport [79].

In summary, importins can recognize their cargo in the cytoplasm, where the RanGTP concentration is low, and then bind to the Nups to go through the NPC. In the nucleus, RanGTP stimulates the separation of the complex and cargo release. The importin–RanGTP complex is recycled, and the RanGTP becomes RanGDP in the cytoplasm through the activity of RanGAP. In contrast, exportins act in the nucleus only in the presence of RanGTP, forming a trimeric cargo–receptor–RanGTP complex that, once transported in the cytoplasm, undergoes dissociation (Fig. 2a–e).

Fig. 2
figure 2

Representation of the NPC structure highlighting the import–export function of proteins and transcripts. a Cargo binds importins to be shuttled to the nucleus. b Nuclear RanGTP induces dissociation of the importin–cargo complex. c The importins bound to RanGTP are recycled to the cytoplasm. d RanGAP1 hydrolyzes RanGTP, maintaining the RanGDP/GTP gradient across the nuclear membrane. e RanGDP is imported into the nucleus through NTF2 and converted to RanGTP by RCC1. f RNAs are shuttled to the cytoplasm, binding their specific exportins. g mRNAs only may undergo an alternative NPC transport interacting with the NFX1–TREX complex. GLE1, Nup214, and DDX19 act as NFX1 modifiers to release mRNAs into the cytoplasm

NPCs are essential not only for protein shuttling across the nuclear envelope but also for RNAs that need to be exported into the cytoplasm. However, the detailed mechanism of RNA transport is not yet fully solved [80]. Several NTRs are implicated, and the directionality is not often accomplished through the RanGDP/RanGTP gradient but by specific RNA helicases [81].

For example, exportin-t (XPO-t) and exportin 5 (XPO5) have key roles in shuttling specific RNAs. The former is implicated in the export of tRNAs, and the latter in the shuttling of double-stranded RNAs and pre-miRNAs [82, 83]. They use an RNA hairpin structure as an NES in a Ran-dependent process [84, 85]. A small subset of mRNAs use CRM1 (XPO1) for their export through the NPC [86]. The principal mRNA export receptor is NXF1 (Nuclear export factor 1) [87], which binds the mRNA in combination with several adaptors (THO complex, UAP56, and AlyRef), collectively forming the TRanscription and EXport complex (TREX) [88]. NTF2 mediates the interaction between NXF1 and the FG-Nups [89]. Once on the cytoplasmic side, Nups such as Gle1, Nup214, and the DEAD-box protein RNA helicase DDX19, that use ATP hydrolysis, revert NXF1 to a lower affinity RNA-binding state to release mRNAs [90]. The proteins involved in mRNAs shuttling are intricately linked with the mRNA modifications at several stages, including transcription, processing, splicing, and poly-adenylation (Fig. 2f-g). All these mRNA maturation processes and NCT components require a tight orchestration to ensure a correct cellular functioning.

Neurodegenerative diseases

Amyotrophic lateral sclerosis and frontotemporal dementia

ALS, also known as Lou Gehrig’s disease, is a fatal neurodegenerative disease characterized by progressive loss of motor neurons in the brain and spinal cord [35]. ALS patients frequently present with behavioral and cognitive deficits, which are symptoms associated with atrophy in the frontotemporal cortex [91]. This particular degeneration had traditionally been linked to another disease, FTD, the most common cause of dementia after AD. Similarly, FTD patients can show motor neuron disease symptoms commonly associated with ALS [92, 93]. This duality prompted consideration of ALS and FTD as two ends of a spectrum disorder termed ALS/FTD.

Although only a small proportion of cases (10%) with ALS/FTD are familial, genetic factors have been firmly established as causative in both familial and sporadic cases, as several mutations display a low penetrance [35]. Mutations in several genes are widely accepted as relevant in ALS, FTD, or ALS/FTD, including SOD1, TARDBP, FUS, C9ORF72, MAPT, TBK1, BCP, GLE1, OPTN, UBQLN2, SQSTM1, ANG, TUBA4A, MATR3, VCP, and CHCHD10 [35].

ALS and FTD cases with different monogenic forms present distinctive molecular neuropathologic signatures. The findings range from inclusions of TDP-43 to FUS and TAU to no inclusions of these proteins, as observed in patients with SOD1 mutations [94]. In post-mortem brains of most patients with sporadic ALS/FTD (97% of ALS and 50% of FTD) [95], TDP-43 and FUS show an altered subcellular localization and appear to be at least partially lost from the nucleus in neuronal and glial cells [94].

Moreover, substantial evidence implicates oxidative stress as a central mechanism of motor neuron death in ALS [96].

TDP-43–related pathology

TARDBP encodes the predominantly nuclear RNA-binding TDP-43, a protein with a NLS and a “putative” NES that can shuttle between the nucleus and cytosol [97]. In addition to nuclear functions, TDP-43 is involved in multiple processes in the cytoplasm, including mRNA stability, transport, and translation [98]. TDP-43 is essential for neuronal survival and selective neuronal degeneration appears in the presence of pathological TDP-43 [99, 100]. The precise mechanism by which TDP-43 passes through the nuclear membrane is unclear; however, recent evidence suggests that despite the predicted NES, the protein is exported independently of the export receptor CRM1/XPO1 [101,102,103]. Thus, no single exporter is necessary for TDP-43 export and TDP-43 nuclear egress seems to follow a model of passive diffusion; these findings, provided by silencing specific Nups in Hela cells, suggest that redundant pathways regulate its transport [101, 102]. Other factors may influence TDP-43 cytoplasmic translocation, as recently demonstrated for the tumor-suppressor folliculin protein, which directly interacts with the RNA-recognition motif domain [104]. These experiments, performed in non-neuronal cell lines and in rat primary cortical neurons, leave unexplored the possibility of other mechanisms specific of human neurons, this should be investigated in future experiments [102].

Whether the TDP43 pathology causes import–export defects or whether defects in NPC components precede the TDP43 pathology is still unclear. TDP-43 is imported into the nucleus by the import pathway Kaps α/β1, using the Ran GDP/GTP gradient; “Kap α” is recycled to the cytoplasm via cellular apoptosis susceptibility protein (CAS) [105]. Disruption of this pathway in human and mouse neuroblastoma cell lines and primary murine cortical neurons, by downregulation of Kap β1 or CAS, leads to TDP-43 cytoplasmic accumulation [105]. Of note, some studies have shown decreased levels of CAS, Kap α2, or α6 in motor neurons of FTD patients with TDP-43 mutations and in ALS patients [105]. Direct evidence of TDP-43 translocation to the cytoplasm was nicely demonstrated in vivo in zebrafish, with an innovative fluorescent tool that visualizes the dynamic of protein transport in neurons [106]. In particular, TDP-43 reaches the cytoplasm, dendrites, and extracellular matrix, inducing neuronal cell death, and can form cytoplasmic aggregates in the absence of microglia [106].

However, if TDP-43 translocation after UV-mediated injury affects also other neuronal subtypes is still unexplored. Future studies aimed at filling this gap might help to clarify the mechanisms of motor neuron pathology selectivity in ALS/FTD.

TDP-43 pathology causes the mislocalization and aggregation of Nups and transport factors, as well as disruption of nuclear membrane and NPCs, leading to impaired nuclear protein import and mRNA export [34]. Differently, nuclear alterations of Nup62 or Kap β1 have been detected by nuclear staining of spinal motor neurons from ALS patients [107]. Recently, new emerging observations show that ribonucleoprotein complexes, such as those containing TDP-43, form liquid droplets with different biophysical characteristics at different neuronal sub-cellular locations [36, 108]. Interestingly, mutations of TDP-34 may confer toxic properties to these droplets [108]. Furthermore, TDP-43 was shown to undergo phase-separation into liquid droplets in the cytosol in vivo after TDP-43 overexpression or exposure to amyloid-like fibrillar TDP-43 or FUS [36]. Droplets of cytosolic TDP-43 were shown to recruit importin-α and Nup62, and to induce mislocalization of RanGap1, Ran, and Nup107, with the consequent inhibition of NCT [36]. Further experiments in this promising direction will define the intricate interplay between the role of pathological aggregates and the alteration of Nups and consequently of the NCT, perhaps also taking into consideration other neurodegenerative pathologies characterized by alterations of different ribonucleoproteins.

Experimental models using Neuroblastoma cell lines overexpressing several Nups and wild-type or mutated TDP-43 have shown that FG repeat-containing Nups, scaffold Nups, and nuclear export factors co-aggregate with mutated TDP-43. Nup205 has been found to mislocalize in patients’ fibroblasts, induced pluripotent stem cell (iPSC)-derived motor neurons, and showed co-aggregation with TDP-43 positive cytoplasmic inclusions in brain tissue of ALS patient [34]. Starting from these observations, future studies evaluating additional Nups mislocalization and co-aggregation with TDP-43 in ALS and FTD brain samples and iPSCs-derived motor neurons will define a clearer picture of the neurodegenerative process involving NPC and protein aggregates. In the same study, NPC defects in the Drosophila model expressing human mutant TDP-43 could be rescued by reducing TDP-43 aggregates and vice versa, suggesting a complex interplay of both components of the degenerative process [34]. In ALS/FTD fly models with the hexanucleotide C9ORF72 repeat expansion, accumulated cytosolic TDP-43 is found to cause Kap α2/ α4 pathology, to increase levels of dipeptide repeat proteins (DPRs), and to enhance expansion-related toxicity [109].

Finally, actin polymerization was recently linked to NPC integrity and function. Modulation of actin homeostasis in primary motor neurons seems to rescue NPC instability arising from mutant Profilin1 and C9ORF72 repeat expansion [110].

FUS-related pathology

The term “FUSopathies” designates neurodegenerative diseases characterized by neuronal and glial cytoplasmic fused in sarcoma (FUS) inclusions [111]. Cytoplasmic aggregates of FUS can be found in 5% of familial ALS and up to 10% of ALS/FTD brains. Mutations in the FUS gene occur in 5% of familial ALS and < 1% of sporadic ALS cases, but are rare in FTD [112,113,114].

FUS is a predominantly nuclear protein with DNA and RNA binding properties involved in regulating transcription, splicing, miRNA biogenesis, and RNA transport [115, 116]. Among its functions, FUS is also essential for DNA repair, being recruited by PARP1 at DNA damage sites [117, 118]. It belongs to the family of RNA-binding proteins known as FET (FUS/TLS, EWS, and TAF15). At their C-terminus, FET proteins have a homologous PY-NLS motif interacting with β-Kaps TNPO1 or TNPO2, which are involved in nuclear import [119, 120]. FUS was seen to translocate through a calcium-dependent mechanism, and NCT alterations result from its cytosolic translocation [121]. The majority of ALS-associated FUS mutations affect the C-terminal NLS, interfering with TNPO1 binding and consequently leading to FUS cytoplasmic accumulation [122]. Post-translational modifications of FET proteins can trigger the same outcome, affecting their interaction with TNPO1 [123].

FUS inclusions in ALS and FTD differ in that FUS is methylated in ALS and not methylated in FTD. In the first case, TNPO1 does not co-localize with the inclusions, and FUS is not properly imported into the nucleus. In the second, FUS inclusions are not methylated, and TNPO1 co-localizes with FUS inclusions, implicating other factors in cytoplasmic accumulation of the protein [123, 124].

Several lines of evidence suggest a major role of DNA repair impairment resulting from FUS mutations [125, 126]. Loss of nuclear FUS causes DNA nick ligation defects in motor neurons because of reduced recruitment of XRCC1/LigIII to DNA strand breaks, a mechanism that also has recently been associated with AOA4 [125, 126].

In iPSC-derived motor neurons, the induced DNA damage leads to FUS mislocalization in the cytosol, enhancing a vicious cycle by increasing cytoplasmic FUS shuttling [127]. The trigger of FUS cytoplasmic accumulation by XPO1 is a still controversial issue. So far, the most important link between NPC and neurodegeneration associated with FUS pathology derives from a study using transgenic flies. In this model, cytoplasmic stress granules containing FUS could be rescued by downregulating XPO1 or Nup154 (Nup155 in humans), proteins needed to export FUS from the nucleus [101, 128]. To clarify the role of NPC in FUS related-ALS and FTD, further studies aimed at exploring the NPC defects and FUS transport in brain tissues and iPSC-derived neurons are needed.

C9ORF72-related pathology

GGGGCC repeat expansion in the noncoding region of C9ORF72 accounts for up to 80% of familial ALS/FTD, 20–50% of familial ALS, 5–20% of sporadic ALS, and 10–30% of FTD cases [129,130,131]. Of note, most patients with ALS carrying the expansions present TDP-43 pathology [95]. How C9ORF72 dysfunction leads to neurodegeneration is a matter of intensive debate. Some lines of evidence suggest a reduction in C9ORF72 transcript expression in cell lines carrying the pathological expansion, which lies in the promoter region [132]. However, C9ORF72 knockout mice do not show neurodegeneration, raising concerns about haploinsufficiency as a primary cause of the disease [133,134,135,136]. In this view, the observation that patients who are homozygous for C9ORF72 expansion have similar phenotypes to those who are heterozygous suggests a less important role of haploinsufficiency.

Cytoplasmic and nuclear RNA foci formed by the expanded GGGGCC repeat transcripts have been suggested to sequester RNA binding proteins and alter RNA metabolism as possible mechanisms underlying the disease [137]. The expanded hexanucleotide repeat disrupts nucleolar integrity by binding the aborted transcripts to nucleolin [138]. Of interest, nucleolar disruption also has been observed in BAC transgenic mice, Drosophila, human motor neurons derived from patient iPSCs, and brains [131, 139].

C9ORF72 is also translated by an alternative mechanism to produce sense and antisense transcripts, which undergo repeat-associated non-AUG translation in all reading frames [140]. Repeat-associated non-AUG encodes five species of highly toxic DPRs, glycine-proline (GP), glycine-alanine (GA), glycine-arginine (GR), proline-alanine (PA), and proline-arginine (PR) [141, 142]. RNA foci and DPR proteins are found in the brains of C9ORF72-ALS/FTD patients, in human motor neurons derived from C9ORF72-iPSCs [143,144,145], and in C9ORF72 BAC transgenic mice [139, 146, 147]. In Drosophila, GR and PR DPRs show toxicity, while RNA-only repeat PA or GA DPRs are not sufficient to induce degeneration [148]. In line with these findings, studies on yeast and flies confirm the toxicity of arginine-containing DPR proteins and the modifying effect of components of the NCT [149, 150]. DPRs also interact with many proteins components of the NCT [149,150,151,152]. Most proteins that interact with DPRs are low complexity domain proteins, including FUS and TDP-43 [152]. Moreover, a large-scale genetic screen approach in Drosophila showed an enhancement or suppression of the expansion toxicity related to several Nups, including Nup50, Nup152, TNPO1, protein shuttle mediators like Nup50, Ran, CRM1, and KPNB1, and RNA-exportins (TREX complex NXF1, Nup107, Nup160, EXOSC3, GLE1, CRM1, and others) [153]. Worth special mention are EXOSC3 and GLE1, which are strong enhancers of C9ORF72 expansion toxicity and causative in a rare form of motor neuron disease [141].

In line with the critical role of NCT-associated proteins in C9ORF72 pathology is the finding of reduced toxicity linked to the expansion if RanGAP is overexpressed in Drosophila [154]. Nuclear retention of RNAs has been observed in C9ORF72 iPSC-derived cortical neurons, indicating impairment of nuclear RNA export. Moreover, C9ORF72 iPSC-derived neurons display a reduced nucleo–cytoplasmic ratio of Ran and RCC1, leading to defects in the NCT mechanism [150]. Recently, the mislocalization of the RNA-editing enzyme adenosine deaminase, which acts on RNA 2 (ADAR2), was found in C9ORF72-associated ALS/FTD, and is possibly responsible for the associated alterations in RNA processing events [155].

Taken together, these studies suggest that C9ORF72 expansion may play distinctive roles in inducing neurodegeneration, and DPRs that lead to NCT defects is a primary pathogenic mechanism.

SOD1-related pathology

SOD1 mutations are present in approximately 15% of familial and 1–3% of sporadic ALS cases [156]. Although SOD1 was the first gene causatively linked to ALS, no widely accepted mechanisms have emerged to explain the toxicity of the mutant protein [35]. Several hypotheses have been proposed, from the impaired superoxide dismutase activity to misfolded protein aggregation and microglia activation [35]. The loss of the antioxidant function of Cu,Zn-superoxide dismutase leads to RNA oxidation and selective motor neuron degeneration [156, 157].

Some recent evidence suggests an impact of SOD1 mutations on NCT. Mice expressing mutant SOD1 show misregulation of different NPC components and import factors [158]. Specifically, a transgenic G93A mouse displays reduced immunoreactivity of importin-α and importin-β in the nucleus, with an increase in the cytoplasm [158]. In the same mouse model, the nuclear membrane shows irregular Nup62 staining, and Lewy body-like hyaline cytoplasmic inclusions are found in spinal motor neurons [158, 159]. Finally, irregular NM morphology was found in spinal motor neurons of ALS patients with SOD1 mutations. Taking as examples the murine models, studies aimed at investigating alterations of several Nups in cortical and spinal motor neurons of patients with SOD1 mutations will be useful to understand whether a more complex defect of NPCs and NCT underlines the observed nuclear membrane alterations [159].

GLE1-related pathology

GLE1 was first identified as causative in a severe autosomal recessive lethal congenital contracture syndrome 1 and lethal arthrogryposis with anterior horn cell disease [160, 161]. Later, nonsense and missense mutations in this gene were found in ALS [162].

GLE1 is a conserved multifunctional protein that regulates gene expression, and its role has been studied in several processes, including nuclear mRNA export, translation initiation, and termination [163]. GLE1 interacts with the nucleoporin Nup155 [161], and impairment in GLE1 thus is expected to lead to NCT deficiencies. GLE1 mutations found in ALS are linked to mRNA degradation and reduced interaction with NupL2 [162]. Moreover, GLE1 is a strong enhancer of C9ORF72 expansion–related toxicity in a Drosophila model [141].

Huntington’s disease

A consistent proportion of NCT impairment in neurodegenerative diseases is represented by the polyglutamine expansion that interferes with the normal structure and function of the NPC [164]. In HD, the CAG repeat expansion in the Huntingtin (HTT) gene is responsible for most of the deleterious effects that lead to death of striatum medium spiny neurons [165, 166]. Among these mechanisms, damage of polyglutamine on the neurons can be mediated by interference with NCT components, or a direct impairment of NPCs, or indirect interaction with specific Nups. An example of the former is the interaction between the N-terminal of HTT and XPO1, shown in HEK cells, which promotes shuttling between the cytoplasm and nucleus [167]. These fragments also interact with Nup TPR to export HTT from the nucleus, a mechanism that is impaired in the presence of pathological polyglutamine expansion [38]. Studies about HTT-NCT performed on iPSC-derived neurons would help to better comprehend these mechanisms and the selectivity of the disease to specific neuronal populations.

The finding of GLE1 among proteins sequestered within polyglutamine inclusions in HD suggests an indirect mechanism of HTT interference with NPCs and further establishes a link between HTT expansion and NPC defects [168]. Moreover, oxidative stress has long been held to be key to disease progression in HD [169]. Several other recent lines of evidence suggest a role for HTT and its interactions with NCP in neurodegeneration [37]. In mouse models of HD, for example, several proteins important for NCT, including RanGAP1, Nup62, and Nup88, form intranuclear inclusions that co-localize with HTT aggregates in striatal and cortical neurons [37].

A Drosophila model highlights the importance of antisense repeat–associated non-ATG (RAN) proteins in nuclear import [170]. These proteins accumulate in several brain regions with neuronal loss and microglia, including the striatum and white matter [170]. Of note, the pathological phenotype in Drosophila is enhanced with the expression of a dominant-negative form of Ran and rescued by the overexpression of Ran and RanGAP1, which seem to be neuroprotective in HD [37]. In HD patients’ iPSC-derived neurons, the MAP2 nucleo–cytoplasm ratio is significantly increased, indicating that NPCs may be damaged and dysfunctional [37]. Additional results along these lines come from neuropathological analyses of brains from patients with HD. Staining for RanGAP1 and Nup62 showed profound mislocalization in striatal neurons, consistent with the observation of selective striatal pathology in infantile bilateral striatal necrosis caused by Nup62 mutations [37].

Other neurodegenerative disorders with expansion repeats

Spinocerebellar ataxias include several autosomal dominant forms and one recessive form, which are caused by the expansion of CAG or CGG repeats. In spinocerebellar ataxia (SCA) 1–2–3-6-7-17 and Friedreich’s ataxia, the CAG expansions encode for polyglutamine [171]. Other forms (SCA 8–10–31-36) are not polyglutamine repeat diseases, however, and the expansions may follow an aberrant pathogenic mechanism similar to that of C9ORF72. In particular, various combinations of the expansions (GGGGCC-sense and GGCCCC-antisense) are improperly translated into RAN proteins, which are known to be toxic for the cell [164].

For example, Ataxin-3 enters into the nucleus through NPCs, imported by importin α/β. XPO1 and Kap α3 modulate trafficking in HEK cells, Drosophila, and mice. Increased cytoplasmic localization of the expanded Ataxin-3 and reduction of protein aggregates have been observed by Kap α3 knockdown and XPO1 overexpression [172, 173]. The altered interaction between Kap α3 and pathologic expanded Ataxin-3 may suggest a link between the pathogenic mechanism and the NCT, but further studies are needed.

Myotonic dystrophy type 1 (DM1) is caused by CGG repeat expansion in the 3′ untranslated region, similar to the site of expansion in SCA8. Reduced levels of Ref1, the Drosophila orthologue of AlyRef in mammals, exacerbates neurodegeneration in a DM1 model of Drosophila, suggesting a possible underlying role of an impaired RNA transport mechanism, which deserves more studies [164].

RAN proteins translation also appears to have a prominent role in the nuclear retention of transcripts in Fragile-X tremor ataxia syndrome [174]. Fragile-X tremor ataxia is caused by the premutation of FMR1 CGG promoter expansion [175, 176], possibly indicating a pathogenic mechanism similar to that observed in HD, which is an interesting avenue of study to pursue. The protein encoded seems to be important during neural differentiation to read N6-methyladenosine modification of mRNA and promote nuclear export of methylated mRNAs [177].

Recently, impairment in the nuclear export of polyQ-expanded androgen receptor and formation of intranuclear inclusions were demonstrated in spinal and bulbar muscular atrophy [178]. In dentatorubral-pallidoluysian atrophy, which is dominantly caused by a CAG repeat in the Atrophin1 gene, neuropathological findings show intranuclear filamentous inclusions and nuclear membrane defects in cerebellar granule cells, suggesting defects in NCT function or in NPC machinery [179, 180].

Recessive ataxia syndromes

SETX encodes the RNA/DNA helicase Senataxin. Mutations in this gene are responsible for ataxia with oculomotor apraxia (AOA) type 2 and have been linked to NCT impairment [173, 181].

In transgenic mice harboring a single SETX human mutation, the nuclear membrane appears damaged. The protein is less able to reach the nucleus in the presence of mutations, leading to DNA/RNA defects. In SETX knock-in mice, motor neurons and primary neuronal cultures show mislocalization of TDP-43 and FUS in the cytoplasm [173]. SETX mutations also have been reported in rare juvenile forms of ALS [182], prompting speculation about a possible link between motor neuron degeneration and TDP-43/FUS alterations.

Aprataxin, a protein involved in single-strand DNA repair breaks, is mutated in AOA type 1 [183]. Aprataxin reaches the nucleus to exert its function, binding NPCs [184]. In particular, it interacts with Kap α/β and binds the nucleoporin Aladin on the cytoplasmic side of the NPC, initial steps for import into the nucleus. This pattern suggests the importance of a correct NCT for Aprataxin to reach the nucleus and guarantees optimal DNA maintenance, which is particularly relevant for non-dividing cells.

Recently, mutations in the Nup93 were identified in an autosomal form of congenital ataxia [185]. In all of these rare forms of neurodegenerative syndromes, NPC and NCT appear to be relevant, but these preliminary observations need to be further studied.

Parkinson’s disease

Cytoplasmic accumulation of the protein alpha-synuclein in dopaminergic neurons of the substantia nigra is the hallmark of PD [186]. Alpha-synuclein belongs to the family of synucleins, which are proteins identified inside the nucleus, although their normal localization is in the cytoplasm. Different alpha-synuclein species bind the chromatin to induce a transcriptional modification [187]. In this role, alpha-synuclein needs phosphorylation at S129, which is specifically found in neurons of patients with an aggressive form of parkinsonism, called Lewy body dementia. Whether alpha-synuclein is transported into the nucleus is still unknown, and uncovering a possible pathological effect of alpha-synuclein aggregates on NPCs or NCT may shed light on the mechanism of PD-related neurodegeneration [187]. Oxidative damage contributes to the cascade of events leading to degeneration of highly functioning dopaminergic neurons in PD [188].

Further suggesting a possible link between altered NCT and PD, several studies have demonstrated a mislocalization of specific transcription factors in PD brains. In particular, NF-κB shows increased nuclear translocation in dopaminergic neurons of PD [189]. In addition, cytoplasmic aggregates of pCREB and the lack of the expected nuclear staining are observed in PD substantia nigra pars compacta [190]. In line with this observation, primary midbrain cultures treated with 6-hydroxydopamine show progressive accumulation of pCREB in the cytoplasm and decreased pCREB in the nucleus of dopaminergic neurons but not in nondopaminergic neurons [190].

Of note, proteins involved in recessive forms of PD have also been found to interact with Nups. Parkin, encoded by PARK2, is mutated in juvenile PD. It is an E3 ubiquitin ligase and specifically ubiquitinates Nup358/RanBP2, a protein fundamental to binding Ran to NPCs [191]. The role of alpha-synuclein in PARK2-related PD is controversial because most studies do not show alpha-synuclein accumulation in dopaminergic neurons in patients with Parkin mutations [192,193,194]. In addition, mice lacking one copy of Nup358/RanBP2 and treated with 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine, known to induce a PD-phenotype in animal models, have a more severe disease course and slower recovery [195, 196]. These findings suggest that NPC alterations in PD might occur regardless of alpha-synuclein accumulation. Further studies are needed to corroborate these initial observations of NCT impairment in PD, with a focus on the role of alpha-synuclein or other possible mechanisms.

Alzheimer’s disease

The classical pathological hallmarks of AD consist of deposition of beta-amyloid in cortical plaques, hyperphosphorylation of tau, and formation of neurofibrillary tangles, causing neuronal degeneration [197]. Initial observations suggest the presence of nuclear membrane alteration and aggregation of NPC in AD brains [198, 199]. Of interest, NTF2 accumulates in the cytoplasm of some hippocampal AD neurons regardless of the presence of tangles, and accumulation of importin-α1 has been found in CA1-hippocampal neuronal inclusions in AD brains [200]. More recently, the number of NPCs has been found to be significantly reduced in AD brains [40], and Ran-reduced expression has also been reported in postmortem tissues from AD cases [201].

The precise mechanism that leads to protein mislocalization is not clear, but one significant finding has demonstrated a link between aberrant tau accumulation and NPC impairment in AD [40]. In particular, that study provided evidence that AD-related tau disrupts nuclear pore function in AD and that Nups can trigger tau to aggregate. Specifically, one component of the NPC, Nup98, interacts with tau, facilitating its aggregation. In addition, in AD brains and tau-mutated mouse models, Nup98 is mislocalized into the cytoplasm [40]. This Nup98 mislocalization and NCT impairment can be rescued by solubilizing tau aggregates in mice. Hyperphosphorylated and mislocalized tau protein has been found also in FTD-MAPT neurons to lead to microtubules impairment and to damage the nuclear membrane [202, 203].

Finally, the Musashi (MSI) family of RNA-binding proteins has been recently investigated using a tau-inducible HEK model. Tau co-localizes with MSI proteins in the cytoplasm and nucleus, altering the nuclear transport of MSI and inducing structural changes to LaminB1, leading to nuclear instability [204].

Allgrove syndrome

A dysfunctional NCT system has been proposed to operate in triple-A syndrome, also known as Allgrove syndrome [11, 184, 205]. This rare disorder is caused by mutations in a gene that encodes the Nup Aladin. Affected fibroblasts exhibit an impaired importin-α–mediated import pathway and weak nuclear import of Aprataxin and DNA ligase I [159, 206]. As a result, cells become more susceptible to oxidative stress, accumulate damaged DNA, and undergo cell death [159]. Of interest, patients with Allgrove syndrome display a peculiar brain multisystem neurodegenerative involvement, affecting motor neurons, Purkinje cells, striatal neurons, autonomic system, peripheral nerves, and endocrine system at the same time [207]. This disease could serve as an example model to study how the same mechanism (an NPC impairment) could underlie neurodegeneration of different neuronal systems.

Aging

In cells with a peculiar characteristic of being long-lived and non-dividing, the integrity of the genome is a fundamental issue [208]. Neurons are such cells: they do not rely on mitosis to renew their DNA and are strictly dependent for their entire lifespan on the mechanisms that guarantee the protection of DNA from external (toxins) and internal (ROS) damage. In this attempt, the machinery that provides a proper nuclear supply of DNA repair proteins (i.e., Aprataxin) and ROS scavengers must work efficiently [209]. The progressively reduced efficiency of this process and the accumulation of ROS and DNA damage lead to physiological cellular aging [210, 211].

Because the NCT is responsible for the supply role, an impaired NPC function may accelerate the alterations observed in the aging process. An example of accelerated aging from NCT deficiency is Hutchinson–Gilford progeria syndrome. This inherited laminopathy causes premature, rapid aging shortly after birth. In this syndrome, a mutant form of Lamin A leads to a dysmorphic nuclear lamina structure with the consequent alteration of the physiological RanGDP/GTP gradient, essential for a proper NCT functionality [212, 213].

Several pieces of evidence link neuronal aging with progressive NPC leaking. Aging-related progressive reduced density of NPCs has been found in the dentate gyrus of hippocampal rat neurons [214] Other studies have demonstrated altered Nups composition and turnover in aged rat neurons [215, 216]. Moreover, in iPSC-derived neurons and directly reprogrammed induced neurons from young and old humans, RanBP17 is reduced in older cells [217]. The reduced levels of this nuclear import receptor have a direct impact on the ability of aged neurons to maintain proper nuclear compartmentalization [217]. For these reasons, more elucidation of the function of Nups and their relationship with aging will significantly affect our understanding of the association of neuronal frailty with aging, opening the way to targeting the underlying molecular mechanisms.

Conclusions

The evidence clearly indicates that NCT rely on NPCs as key factors in neuronal health and vitality. Many alterations lead to a broad spectrum of neurodegenerative patterns and to neuronal aging. The most important aspects seem to involve the maintenance of a proper RanGDP/GTP gradient and the shuttling of transcription factors and proteins involved in protein and RNA conservation. However, most of the studies investigating these mechanisms are focused on ALS and FTD models, and a deep investigation of the link of NCT, NPC and neuronal cell death is needed in other neurodegenerative diseases, especially the most common AD and PD. Moreover, to confirm the observations from animal experimental models, additional studies on human iPSC-derived neurons and neuropathological samples are needed. The reason specific neuronal subsets are variably susceptible to specific damage that leads to different phenotypes of neurodegenerative diseases remains unclear. The study of emblematic but rare diseases that display an NPC-related impairment in different neuronal types, such as Allgrove syndrome, may shed light on common mechanisms underlying these neurodegenerative processes. Moreover, targeting NCT deficiency as a common pathway of several neurodegenerative diseases and neuronal aging may represent a unique therapeutic opportunity for these incurable disorders.

Availability of data and materials

Not applicable.

Abbreviations

AD:

Alzheimer’s disease

ADAR2:

Enzyme adenosine deaminase acting on RNA 2

ALS:

Amyotrophic lateral sclerosis

AOA:

Ataxia with oculomotor apraxia

CAS:

Cellular apoptosis susceptibility protein

CRM1:

Chromosome region maintenance 1

DM1:

Myotonic dystrophy type 1

DPR:

Dipeptide repeat proteins

FG:

Phenylalanine-glycine

FTD:

Frontotemporal dementia

GA:

Glycine-alanine

GP:

Glycine-proline

GR:

Glycine-arginine

HD:

Huntington’s disease

HTT:

Huntingtin

Kaps:

Karyopherins

MSI:

Musashi

NCT:

Nucleo–cytoplasmic transport

NES:

Nuclear export signal

NLS:

Nuclear localization signal

NPC:

Nuclear pore complex

NTF2:

Nuclear transport factor 2

NTRs:

Nuclear transport receptors

Nups:

Nucleoporins

NXF1:

Nuclear export factor 1

PA:

Proline-alanine

PD:

Parkinson’s disease

PR:

Proline-arginine

RAN:

repeat–associated non-ATG

Ran:

Ras-related nuclear protein

RCC1:

Regulator of chromosome condensation 1

ROS:

Radical oxygen species

SCA:

Spinocerebellar ataxia

TREX:

Transcription and export complex

XPO-t:

Exportin-t

XPO5:

Exportin-5

References

  1. Taylor JP, Hardy J, Fischbeck KH. Toxic proteins in neurodegenerative disease. Science. 2002;296:1991–5.

    Article  CAS  PubMed  Google Scholar 

  2. Nouspikel T, Hanawalt PC. When parsimony backfires: neglecting DNA repair may doom neurons in Alzheimer’s disease. BioEssays News Rev Mol Cell Dev Biol. 2003;25:168–73.

    Article  CAS  Google Scholar 

  3. Bender A, Krishnan KJ, Morris CM, Taylor GA, Reeve AK, Perry RH, et al. High levels of mitochondrial DNA deletions in substantia nigra neurons in aging and Parkinson disease. Nat Genet. 2006;38:515–7.

    Article  CAS  PubMed  Google Scholar 

  4. Katyal S, McKinnon PJ. DNA strand breaks, neurodegeneration and aging in the brain. Mech Ageing Dev. 2008;129:483–91.

    Article  CAS  PubMed  Google Scholar 

  5. Wente SR, Rout MP. The nuclear pore complex and nuclear transport. Cold Spring Harb Perspect Biol. 2010;2:a000562.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Strambio-De-Castillia C, Niepel M, Rout MP. The nuclear pore complex: bridging nuclear transport and gene regulation. Nat Rev Mol Cell Biol. 2010;11:490–501.

    Article  CAS  PubMed  Google Scholar 

  7. Kulkarni A, Wilson DM. The involvement of DNA-damage and -repair defects in neurological dysfunction. Am J Hum Genet. 2008;82:539–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Hoeijmakers JHJ. DNA damage, aging, and cancer. N Engl J Med. 2009;361:1475–85.

    Article  CAS  PubMed  Google Scholar 

  9. Boehringer A, Bowser R. RNA nucleocytoplasmic transport defects in neurodegenerative diseases. In: Sattler R, Donnelly CJ, editors. RNA Metab Neurodegener Dis. Cham: Springer International Publishing; 2018. p. 85–101. Available from: http://link.springer.com/10.1007/978-3-319-89689-2_4. [cited 2018 Sep 16].

    Chapter  Google Scholar 

  10. Ori A, Banterle N, Iskar M, Andrés-Pons A, Escher C, Khanh Bui H, et al. Cell type-specific nuclear pores: a case in point for context-dependent stoichiometry of molecular machines. Mol Syst Biol. 2013;9:648.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Cronshaw JM, Matunis MJ. The nuclear pore complex protein ALADIN is mislocalized in triple A syndrome. Proc Natl Acad Sci. 2003;100:5823–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Cronshaw JM, Krutchinsky AN, Zhang W, Chait BT, Matunis MJ. Proteomic analysis of the mammalian nuclear pore complex. J Cell Biol. 2002;158:915–27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Wozniak R, Burke B, Doye V. Nuclear transport and the mitotic apparatus: an evolving relationship. Cell Mol Life Sci CMLS. 2010;67:2215–30.

    Article  CAS  PubMed  Google Scholar 

  14. Weis K. Regulating access to the genome: nucleocytoplasmic transport throughout the cell cycle. Cell. 2003;112:441–51.

    Article  CAS  PubMed  Google Scholar 

  15. Cronshaw JM, Matunis MJ. The nuclear pore complex: disease associations and functional correlations. Trends Endocrinol Metab. 2004;15:34–9.

    Article  CAS  PubMed  Google Scholar 

  16. Izaurralde E, Mattaj IW. RNA export. Cell. 1995;81:153–9.

    Article  CAS  PubMed  Google Scholar 

  17. Moore MS, Blobel G. A G protein involved in nucleocytoplasmic transport: the role of Ran. Trends Biochem Sci. 1994;19:211–6.

    Article  CAS  PubMed  Google Scholar 

  18. Koonin EV, Aravind L. Comparative genomics, evolution and origins of the nuclear envelope and nuclear pore complex. Cell Cycle Georget Tex. 2009;8:1984–5.

    Article  CAS  Google Scholar 

  19. Bapteste E, Charlebois RL, MacLeod D, Brochier C. The two tempos of nuclear pore complex evolution: highly adapting proteins in an ancient frozen structure. Genome Biol. 2005;6:R85.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  20. Pan L, Penney J, Tsai L-H. Chromatin regulation of DNA damage repair and genome integrity in the central nervous system. J Mol Biol. 2014;426:3376–88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. R M, L P, Lh T. DNA damage and its links to neurodegeneration. Neuron. 2014; [cited 2020 May 23]. Available from: https://pubmed.ncbi.nlm.nih.gov/25033177/.

  22. Finkel T, Holbrook NJ. Oxidants, oxidative stress and the biology of ageing. Nature. 2000;408:239–47.

    Article  CAS  PubMed  Google Scholar 

  23. Mattson MP, Magnus T. Ageing and neuronal vulnerability. Nat Rev Neurosci. 2006;7:278–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Lodato MA, Rodin RE, Bohrson CL, Coulter ME, Barton AR, Kwon M, et al. Aging and neurodegeneration are associated with increased mutations in single human neurons. Science. 2018;359:555–9.

    Article  CAS  PubMed  Google Scholar 

  25. Pj M. DNA repair deficiency and neurological disease. Nat Rev Neurosci. 2009; Available from: https://pubmed.ncbi.nlm.nih.gov/19145234/. [cited 2020 May 23].

  26. Kim HJ, Taylor JP. Lost in transportation: nucleocytoplasmic transport defects in ALS and other neurodegenerative diseases. Neuron. 2017;96:285–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Patel VP, Chu CT. Nuclear transport, oxidative stress, and neurodegeneration. Int J Clin Exp Pathol. 2011;4:215–29.

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Robijns J, Houthaeve G, Braeckmans K, De Vos WH. Loss of nuclear envelope integrity in aging and disease. Int Rev Cell Mol Biol. 2018:205–22 Available from: https://linkinghub.elsevier.com/retrieve/pii/S1937644817300886. [cited 2018 Sep 16].

  29. Xie Y, Ren Y. Mechanisms of nuclear mRNA export: a structural perspective. Traffic Cph Den. 2019;20:829–40.

    Article  CAS  Google Scholar 

  30. Schwenzer H, Jühling F, Chu A, Pallett LJ, Baumert TF, Maini M, et al. Oxidative stress triggers selective tRNA retrograde transport in human cells during the integrated stress response. Cell Rep. 2019;26:3416–3428.e5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Sotelo C, Dusart I. Intrinsic versus extrinsic determinants during the development of Purkinje cell dendrites. Neuroscience. 2009;162:589–600.

    Article  CAS  PubMed  Google Scholar 

  32. Madabhushi R, Kim T-K. Emerging themes in neuronal activity-dependent gene expression. Mol Cell Neurosci. 2018;87:27–34.

    Article  CAS  PubMed  Google Scholar 

  33. Van Driesche SJ, Martin KC. New frontiers in RNA transport and local translation in neurons. Dev Neurobiol. 2018;78:331–9.

    Article  PubMed  CAS  Google Scholar 

  34. Chou C-C, Zhang Y, Umoh ME, Vaughan SW, Lorenzini I, Liu F, et al. TDP-43 pathology disrupts nuclear pore complexes and nucleocytoplasmic transport in ALS/FTD. Nat Neurosci. 2018;21:228–39.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Taylor JP, Brown RH, Cleveland DW. Decoding ALS: from genes to mechanism. Nature. 2016;539:197–206.

    Article  PubMed  PubMed Central  Google Scholar 

  36. Gasset-Rosa F, Lu S, Yu H, Chen C, Melamed Z, Guo L, et al. Cytoplasmic TDP-43 de-mixing independent of stress granules drives inhibition of nuclear import, loss of nuclear TDP-43, and cell death. Neuron. 2019;102:339–357.e7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Grima JC, Daigle JG, Arbez N, Cunningham KC, Zhang K, Ochaba J, et al. Mutant huntingtin disrupts the nuclear pore complex. Neuron. 2017;94:93–107.e6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Cornett J, Cao F, Wang C-E, Ross CA, Bates GP, Li S-H, et al. Polyglutamine expansion of huntingtin impairs its nuclear export. Nat Genet. 2005;37:198–204.

    Article  CAS  PubMed  Google Scholar 

  39. Li N, Lagier-Tourenne C. Nuclear pores: the gate to neurodegeneration. Nat Neurosci. 2018;21:156–8.

    Article  CAS  PubMed  Google Scholar 

  40. Eftekharzadeh B, Daigle JG, Kapinos LE, Coyne A, Schiantarelli J, Carlomagno Y, et al. Tau protein disrupts nucleocytoplasmic transport in Alzheimer’s disease. Neuron. 2018;99:925–940.e7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Sakuma S, D’Angelo MA. The roles of the nuclear pore complex in cellular dysfunction, aging and disease. Semin Cell Dev Biol. 2017;68:72–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Ciechanover A, Kwon YT. Degradation of misfolded proteins in neurodegenerative diseases: therapeutic targets and strategies. Exp Mol Med. 2015;47:e147.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. De Magistris P, Antonin W. The dynamic nature of the nuclear envelope. Curr Biol. 2018;28:R487–97.

    Article  PubMed  CAS  Google Scholar 

  44. de las Heras JI, Meinke P, Batrakou DG, Srsen V, Zuleger N, Kerr AR, et al. Tissue specificity in the nuclear envelope supports its functional complexity. Nucleus. 2013;4:460–77.

    Article  PubMed Central  Google Scholar 

  45. Amendola M, van Steensel B. Mechanisms and dynamics of nuclear lamina-genome interactions. Curr Opin Cell Biol. 2014;28:61–8.

    Article  CAS  PubMed  Google Scholar 

  46. D’Angelo MA, Hetzer MW. Structure, dynamics and function of nuclear pore complexes. Trends Cell Biol. 2008;18:456–66.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  47. D’Angelo MA. Nuclear pore complexes as hubs for gene regulation. Nucleus. 2018;9:142–8.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. Brohawn SG, Partridge JR, Whittle JRR, Schwartz TU. The nuclear pore complex has entered the atomic age. Struct Lond Engl 1993. 2009;17:1156–68.

    CAS  Google Scholar 

  49. Paine PL. Nucleocytoplasmic movement of fluorescent tracers microinjected into living salivary gland cells. J Cell Biol. 1975;66:652–7.

    Article  CAS  PubMed  Google Scholar 

  50. Reichelt R, Holzenburg A, Buhle EL Jr, Jarnik M, Engel A, Aebi U. Correlation between structure and mass distribution of the nuclear pore complex and of distinct pore complex components. J Cell Biol. 1990;110:883–94.

    Article  CAS  PubMed  Google Scholar 

  51. Rout MP, Aitchison JD, Suprapto A, Hjertaas K, Zhao Y, Chait BT. The yeast nuclear pore complex: composition, architecture, and transport mechanism. J Cell Biol. 2000;148:635–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. De Robertis EM, Longthorne RF, Gurdon JB. Intracellular migration of nuclear proteins in Xenopus oocytes. Nature. 1978;272:254–6.

    Article  PubMed  Google Scholar 

  53. Dingwall C, Sharnick SV, Laskey RA. A polypeptide domain that specifies migration of nucleoplasmin into the nucleus. Cell. 1982;30:449–58.

    Article  CAS  PubMed  Google Scholar 

  54. Hayama R, Rout MP, Fernandez-Martinez J. The nuclear pore complex core scaffold and permeability barrier: variations of a common theme. Curr Opin Cell Biol. 2017;46:110–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Suntharalingam M, Wente SR. Peering through the pore: nuclear pore complex structure, assembly, and function. Dev Cell. 2003;4:775–89.

    Article  CAS  PubMed  Google Scholar 

  56. Raveh B, Karp JM, Sparks S, Dutta K, Rout MP, Sali A, et al. Slide-and-exchange mechanism for rapid and selective transport through the nuclear pore complex. Proc Natl Acad Sci. 2016;113:E2489–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Beck M, Hurt E. The nuclear pore complex: understanding its function through structural insight. Nat Rev Mol Cell Biol. 2017;18:73–89.

    Article  CAS  PubMed  Google Scholar 

  58. Zilman A. Aggregation, phase separation and spatial morphologies of the assemblies of FG nucleoporins. J Mol Biol. 2018;430:4730–40.

    Article  CAS  PubMed  Google Scholar 

  59. Panté N, Kann M. Nuclear pore complex is able to transport macromolecules with diameters of about 39 nm. Mol Biol Cell. 2002;13:425–34.

    Article  PubMed  PubMed Central  Google Scholar 

  60. Keminer O, Peters R. Permeability of single nuclear pores. Biophys J. 1999;77:217–28.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Frey S, Görlich D. A saturated FG-repeat hydrogel can reproduce the permeability properties of nuclear pore complexes. Cell. 2007;130:512–23.

    Article  CAS  PubMed  Google Scholar 

  62. Mohr D, Frey S, Fischer T, Güttler T, Görlich D. Characterisation of the passive permeability barrier of nuclear pore complexes. EMBO J. 2009;28:2541–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Jovanovic-Talisman T, Zilman A. Protein transport by the nuclear pore complex: simple biophysics of a complex biomachine. Biophys J. 2017;113:6–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Pemberton LF, Paschal BM. Mechanisms of receptor-mediated nuclear import and nuclear export. Traffic Cph Den. 2005;6:187–98.

    Article  CAS  Google Scholar 

  65. Tran EJ, Wente SR. Dynamic nuclear pore complexes: life on the edge. Cell. 2006;125:1041–53.

    Article  CAS  PubMed  Google Scholar 

  66. Sorokin AV, Kim ER, Ovchinnikov LP. Nucleocytoplasmic transport of proteins. Biochem Biokhimiia. 2007;72:1439–57.

    Article  CAS  Google Scholar 

  67. Fukuda M, Asano S, Nakamura T, Adachi M, Yoshida M, Yanagida M, et al. CRM1 is responsible for intracellular transport mediated by the nuclear export signal. Nature. 1997;390:308–11.

    Article  CAS  PubMed  Google Scholar 

  68. Görlich D, Panté N, Kutay U, Aebi U, Bischoff FR. Identification of different roles for RanGDP and RanGTP in nuclear protein import. EMBO J. 1996;15:5584–94.

    Article  PubMed  PubMed Central  Google Scholar 

  69. Bayliss R, Littlewood T, Strawn LA, Wente SR, Stewart M. GLFG and FxFG nucleoporins bind to overlapping sites on importin-beta. J Biol Chem. 2002;277:50597–606.

    Article  CAS  PubMed  Google Scholar 

  70. Bayliss R, Leung SW, Baker RP, Quimby BB, Corbett AH, Stewart M. Structural basis for the interaction between NTF2 and nucleoporin FxFG repeats. EMBO J. 2002;21:2843–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Fu X, Liang C, Li F, Wang L, Wu X, Lu A, et al. The rules and functions of nucleocytoplasmic shuttling proteins. Int J Mol Sci. 2018;19:1445.

    Article  PubMed Central  CAS  Google Scholar 

  72. Moore MS. Nuclear pores: David and Goliath in nuclear transport. Curr Biol. 1995;5:1339–41.

    Article  CAS  PubMed  Google Scholar 

  73. Izaurralde E, Kutay U, von Kobbe C, Mattaj IW, Görlich D. The asymmetric distribution of the constituents of the ran system is essential for transport into and out of the nucleus. EMBO J. 1997;16:6535–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Wen W, Meinkotht JL, Tsien RY, Taylor SS. Identification of a signal for rapid export of proteins from the nucleus. Cell. 1995;82:463–73.

    Article  CAS  PubMed  Google Scholar 

  75. Bischoff FR, Ponstingl H. Catalysis of guanine nucleotide exchange on Ran by the mitotic regulator RCC1. Nature. 1991;354:80–2.

    Article  CAS  PubMed  Google Scholar 

  76. la Cour T, Kiemer L, Mølgaard A, Gupta R, Skriver K, Brunak S. Analysis and prediction of leucine-rich nuclear export signals. Protein Eng Des Sel PEDS. 2004;17:527–36.

    Article  PubMed  CAS  Google Scholar 

  77. Mahajan R, Delphin C, Guan T, Gerace L, Melchior F. A small ubiquitin-related polypeptide involved in targeting RanGAP1 to nuclear pore complex protein RanBP2. Cell. 1997;88:97–107.

    Article  CAS  PubMed  Google Scholar 

  78. Taagepera S, McDonald D, Loeb JE, Whitaker LL, McElroy AK, Wang JY, et al. Nuclear-cytoplasmic shuttling of C-ABL tyrosine kinase. Proc Natl Acad Sci U S A. 1998;95:7457–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Paraskeva E, Izaurralde E, Bischoff FR, Huber J, Kutay U, Hartmann E, et al. CRM1-mediated recycling of snurportin 1 to the cytoplasm. J Cell Biol. 1999;145:255–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Kubitscheck U, Siebrasse J-P. Kinetics of transport through the nuclear pore complex. Semin Cell Dev Biol. 2017;68:18–26.

    Article  CAS  PubMed  Google Scholar 

  81. Bourne HR, Sanders DA, McCormick F. The GTPase superfamily: conserved structure and molecular mechanism. Nature. 1991;349:117–27.

    Article  CAS  PubMed  Google Scholar 

  82. Kutay U, Lipowsky G, Izaurralde E, Bischoff FR, Schwarzmaier P, Hartmann E, et al. Identification of a tRNA-specific nuclear export receptor. Mol Cell. 1998;1:359–69.

    Article  CAS  PubMed  Google Scholar 

  83. Bohnsack MT, Czaplinski K, Gorlich D. Exportin 5 is a RanGTP-dependent dsRNA-binding protein that mediates nuclear export of pre-miRNAs. RNA N Y N. 2004;10:185–91.

    Article  CAS  Google Scholar 

  84. Arts GJ, Kuersten S, Romby P, Ehresmann B, Mattaj IW. The role of exportin-t in selective nuclear export of mature tRNAs. EMBO J. 1998;17:7430–41.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Kim VN. MicroRNA precursors in motion: exportin-5 mediates their nuclear export. Trends Cell Biol. 2004;14:156–9.

    Article  CAS  PubMed  Google Scholar 

  86. Okamura M, Inose H, Masuda S. RNA export through the NPC in eukaryotes. Genes. 2015;6:124–49.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Wickramasinghe VO, Laskey RA. Control of mammalian gene expression by selective mRNA export. Nat Rev Mol Cell Biol. 2015;16:431–42.

    Article  CAS  PubMed  Google Scholar 

  88. Strässer K, Masuda S, Mason P, Pfannstiel J, Oppizzi M, Rodriguez-Navarro S, et al. TREX is a conserved complex coupling transcription with messenger RNA export. Nature. 2002;417:304–8.

    Article  PubMed  CAS  Google Scholar 

  89. Hough LE, Dutta K, Sparks S, Temel DB, Kamal A, Tetenbaum-Novatt J, et al. The molecular mechanism of nuclear transport revealed by atomic-scale measurements. eLife. 2015; Available from: https://elifesciences.org/articles/10027. [cited 2018 Dec 31].

  90. Lari A, Arul Nambi Rajan A, Sandhu R, Reiter T, Montpetit R, Young BP, et al. A nuclear role for the DEAD-box protein Dbp5 in tRNA export. eLife. 2019;8:e48410.

    Article  PubMed  PubMed Central  Google Scholar 

  91. Swinnen B, Robberecht W. The phenotypic variability of amyotrophic lateral sclerosis. Nat Rev Neurol. 2014;10:661–70.

    Article  PubMed  Google Scholar 

  92. Lomen-Hoerth C, Anderson T, Miller B. The overlap of amyotrophic lateral sclerosis and frontotemporal dementia. Neurology. 2002;59:1077–9.

    Article  PubMed  Google Scholar 

  93. Hodges JR, Davies RR, Xuereb JH, Casey B, Broe M, Bak TH, et al. Clinicopathological correlates in frontotemporal dementia. Ann Neurol. 2004;56:399–406.

    Article  PubMed  Google Scholar 

  94. Mackenzie IRA, Neumann M. Molecular neuropathology of frontotemporal dementia: insights into disease mechanisms from postmortem studies. J Neurochem. 2016;138(Suppl 1):54–70.

    Article  CAS  PubMed  Google Scholar 

  95. Ling S-C, Polymenidou M, Cleveland DW. Converging mechanisms in ALS and FTD: disrupted RNA and protein homeostasis. Neuron. 2013;79:416–38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Barber SC, Shaw PJ. Oxidative stress in ALS: key role in motor neuron injury and therapeutic target. Free Radic Biol Med. 2010;48:629–41.

    Article  CAS  PubMed  Google Scholar 

  97. Buratti E, Brindisi A, Giombi M, Tisminetzky S, Ayala YM, Baralle FE. TDP-43 binds heterogeneous nuclear ribonucleoprotein A/B through its C-terminal tail: an important region for the inhibition of cystic fibrosis transmembrane conductance regulator exon 9 splicing. J Biol Chem. 2005;280:37572–84.

    Article  CAS  PubMed  Google Scholar 

  98. Birsa N, Bentham MP, Fratta P. Cytoplasmic functions of TDP-43 and FUS and their role in ALS. Semin Cell Dev Biol. 2019.

  99. Diaper DC, Adachi Y, Sutcliffe B, Humphrey DM, Elliott CJH, Stepto A, et al. Loss and gain of Drosophila TDP-43 impair synaptic efficacy and motor control leading to age-related neurodegeneration by loss-of-function phenotypes. Hum Mol Genet. 2013;22:1539–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Feiguin F, Godena VK, Romano G, D’Ambrogio A, Klima R, Baralle FE. Depletion of TDP-43 affects Drosophila motoneurons terminal synapsis and locomotive behavior. FEBS Lett. 2009;583:1586–92.

    Article  CAS  PubMed  Google Scholar 

  101. Ederle H, Funk C, Abou-Ajram C, Hutten S, Funk EBE, Kehlenbach RH, et al. Nuclear egress of TDP-43 and FUS occurs independently of Exportin-1/CRM1. Sci Rep. 2018;8 Available from: http://www.nature.com/articles/s41598-018-25007-5. [cited 2018 Sep 16].

  102. Archbold HC, Jackson KL, Arora A, Weskamp K, Tank EM-H, Li X, et al. TDP43 nuclear export and neurodegeneration in models of amyotrophic lateral sclerosis and frontotemporal dementia. Sci Rep. 2018;8 Available from: http://www.nature.com/articles/s41598-018-22858-w. [cited 2018 Sep 16].

  103. Sugai A, Kato T, Koyama A, Koike Y, Kasahara S, Konno T, et al. Robustness and vulnerability of the autoregulatory system that maintains nuclear TDP-43 levels: a trade-off hypothesis for ALS pathology based on in silico data. Front Neurosci. 2018;12:28.

    Article  PubMed  PubMed Central  Google Scholar 

  104. Xia Q, Wang G, Wang H, Hu Q, Ying Z. Folliculin, a tumor suppressor associated with Birt-Hogg-Dubé (BHD) syndrome, is a novel modifier of TDP-43 cytoplasmic translocation and aggregation. Hum Mol Genet. 2016;25:83–96.

    Article  CAS  PubMed  Google Scholar 

  105. Nishimura AL, Zupunski V, Troakes C, Kathe C, Fratta P, Howell M, et al. Nuclear import impairment causes cytoplasmic trans-activation response DNA-binding protein accumulation and is associated with frontotemporal lobar degeneration. Brain J Neurol. 2010;133:1763–71.

    Article  Google Scholar 

  106. Svahn AJ, Don EK, Badrock AP, Cole NJ, Graeber MB, Yerbury JJ, et al. Nucleo-cytoplasmic transport of TDP-43 studied in real time: impaired microglia function leads to axonal spreading of TDP-43 in degenerating motor neurons. Acta Neuropathol (Berl). 2018;136:445–59.

    Article  CAS  Google Scholar 

  107. Aizawa H, Yamashita T, Kato H, Kimura T, Kwak S. Impaired nucleoporins are present in sporadic amyotrophic lateral sclerosis motor neurons that exhibit mislocalization of the 43-kDa TAR DNA-binding protein. J Clin Neurol Seoul Korea. 2019;15:62–7.

    Article  Google Scholar 

  108. Gopal PP, Nirschl JJ, Klinman E, Holzbaur ELF. Amyotrophic lateral sclerosis-linked mutations increase the viscosity of liquid-like TDP-43 RNP granules in neurons. Proc Natl Acad Sci U S A. 2017;114:E2466–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Solomon DA, Stepto A, Au WH, Adachi Y, Diaper DC, Hall R, et al. A feedback loop between dipeptide-repeat protein, TDP-43 and karyopherin-α mediates C9orf72-related neurodegeneration. Brain. 2018;141:2908–24.

    Article  PubMed  PubMed Central  Google Scholar 

  110. Giampetruzzi A, Danielson EW, Gumina V, Jeon M, Boopathy S, Brown RH, et al. Modulation of actin polymerization affects nucleocytoplasmic transport in multiple forms of amyotrophic lateral sclerosis. Nat Commun. 2019;10:3827.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  111. Shelkovnikova TA. Modelling FUSopathies: focus on protein aggregation. Biochem Soc Trans. 2013;41:1613–7.

    Article  CAS  PubMed  Google Scholar 

  112. Kwiatkowski TJ, Bosco DA, Leclerc AL, Tamrazian E, Vanderburg CR, Russ C, et al. Mutations in the FUS/TLS gene on chromosome 16 cause familial amyotrophic lateral sclerosis. Science. 2009;323:1205–8.

    Article  CAS  PubMed  Google Scholar 

  113. Kwong LK, Neumann M, Sampathu DM, Lee VM-Y, Trojanowski JQ. TDP-43 proteinopathy: the neuropathology underlying major forms of sporadic and familial frontotemporal lobar degeneration and motor neuron disease. Acta Neuropathol (Berl). 2007;114:63–70.

    Article  CAS  Google Scholar 

  114. Sreedharan J, Blair IP, Tripathi VB, Hu X, Vance C, Rogelj B, et al. TDP-43 mutations in familial and sporadic amyotrophic lateral sclerosis. Science. 2008;319:1668–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Rogelj B, Easton LE, Bogu GK, Stanton LW, Rot G, Curk T, et al. Widespread binding of FUS along nascent RNA regulates alternative splicing in the brain. Sci Rep. 2012;2:603.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  116. De Santis R, Santini L, Colantoni A, Peruzzi G, de Turris V, Alfano V, et al. FUS mutant human motoneurons display altered transcriptome and microRNA pathways with implications for ALS pathogenesis. Stem Cell Rep. 2017;9:1450–62.

    Article  CAS  Google Scholar 

  117. Wang W-Y, Pan L, Su SC, Quinn EJ, Sasaki M, Jimenez JC, et al. Interaction of FUS and HDAC1 regulates DNA damage response and repair in neurons. Nat Neurosci. 2013;16:1383–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Rulten SL, Rotheray A, Green RL, Grundy GJ, Moore DAQ, Gómez-Herreros F, et al. PARP-1 dependent recruitment of the amyotrophic lateral sclerosis-associated protein FUS/TLS to sites of oxidative DNA damage. Nucleic Acids Res. 2014;42:307–14.

    Article  CAS  PubMed  Google Scholar 

  119. Brelstaff J, Lashley T, Holton JL, Lees AJ, Rossor MN, Bandopadhyay R, et al. Transportin1: a marker of FTLD-FUS. Acta Neuropathol (Berl). 2011;122:591–600.

    Article  CAS  Google Scholar 

  120. Darovic S, Prpar Mihevc S, Župunski V, Gunčar G, Štalekar M, Lee Y-B, et al. Phosphorylation of C-terminal tyrosine residue 526 in FUS impairs its nuclear import. J Cell Sci. 2015;128:4151–9.

    Article  CAS  PubMed  Google Scholar 

  121. Tischbein M, Baron DM, Lin Y-C, Gall KV, Landers JE, Fallini C, et al. The RNA-binding protein FUS/TLS undergoes calcium-mediated nuclear egress during excitotoxic stress and is required for GRIA2 mRNA processing. J Biol Chem. 2019;294:10194–210.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Bosco DA, Lemay N, Ko HK, Zhou H, Burke C, Kwiatkowski TJ, et al. Mutant FUS proteins that cause amyotrophic lateral sclerosis incorporate into stress granules. Hum Mol Genet. 2010;19:4160–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Dormann D, Madl T, Valori CF, Bentmann E, Tahirovic S, Abou-Ajram C, et al. Arginine methylation next to the PY-NLS modulates Transportin binding and nuclear import of FUS. EMBO J. 2012;31:4258–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Qamar S, Wang G, Randle SJ, Ruggeri FS, Varela JA, Lin JQ, et al. FUS phase separation is modulated by a molecular chaperone and methylation of arginine cation-π interactions. Cell. 2018;173:720–734.e15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Hoch NC, Hanzlikova H, Rulten SL, Tétreault M, Komulainen E, Ju L, et al. XRCC1 mutation is associated with PARP1 hyperactivation and cerebellar ataxia. Nature. 2017;541:87–91.

    Article  CAS  PubMed  Google Scholar 

  126. Wang H, Guo W, Mitra J, Hegde PM, Vandoorne T, Eckelmann BJ, et al. Mutant FUS causes DNA ligation defects to inhibit oxidative damage repair in amyotrophic lateral sclerosis. Nat Commun. 2018;9:3683.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  127. Naumann M, Pal A, Goswami A, Lojewski X, Japtok J, Vehlow A, et al. Impaired DNA damage response signaling by FUS-NLS mutations leads to neurodegeneration and FUS aggregate formation. Nat Commun. 2018;9 Available from: http://www.nature.com/articles/s41467-017-02299-1. [cited 2018 Sep 16].

  128. Steyaert J, Scheveneels W, Vanneste J, Van Damme P, Robberecht W, Callaerts P, et al. FUS-induced neurotoxicity in Drosophila is prevented by downregulating nucleocytoplasmic transport proteins. Hum Mol Genet. 2018;27:4103–16.

    CAS  PubMed  PubMed Central  Google Scholar 

  129. DeJesus-Hernandez M, Mackenzie IR, Boeve BF, Boxer AL, Baker M, Rutherford NJ, et al. Expanded GGGGCC hexanucleotide repeat in noncoding region of C9ORF72 causes chromosome 9p-linked FTD and ALS. Neuron. 2011;72:245–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Renton AE, Majounie E, Waite A, Simón-Sánchez J, Rollinson S, Gibbs JR, et al. A hexanucleotide repeat expansion in C9ORF72 is the cause of chromosome 9p21-linked ALS-FTD. Neuron. 2011;72:257–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Yuva-Aydemir Y, Almeida S, Gao F-B. Insights into C9ORF72 -related ALS/FTD from drosophila and iPSC models. Trends Neurosci. 2018;41:457–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Babić Leko M, Župunski V, Kirincich J, Smilović D, Hortobágyi T, Hof PR, et al. Molecular mechanisms of neurodegeneration related to C9orf72 hexanucleotide repeat expansion. Behav Neurol. 2019;2019 Available from: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6350563/. [cited 2020 Feb 8].

  133. Burberry A, Suzuki N, Wang J-Y, Moccia R, Mordes DA, Stewart MH, et al. Loss-of-function mutations in the C9ORF72 mouse ortholog cause fatal autoimmune disease. Sci Transl Med. 2016;8:347ra93.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  134. Xiao S, MacNair L, McGoldrick P, McKeever PM, McLean JR, Zhang M, et al. Isoform-specific antibodies reveal distinct subcellular localizations of C9orf72 in amyotrophic lateral sclerosis. Ann Neurol. 2015;78:568–83.

    Article  CAS  PubMed  Google Scholar 

  135. C9orf72 is required for proper macrophage and microglial function in mice | Science [Internet]. Available from: http://science.sciencemag.org/content/351/6279/1324. [cited 2018 Dec 1].

  136. Gijselinck I, Van Langenhove T, van der Zee J, Sleegers K, Philtjens S, Kleinberger G, et al. A C9orf72 promoter repeat expansion in a Flanders-Belgian cohort with disorders of the frontotemporal lobar degeneration-amyotrophic lateral sclerosis spectrum: a gene identification study. Lancet Neurol. 2012;11:54–65.

    Article  CAS  PubMed  Google Scholar 

  137. Mohan A, Goodwin M, Swanson MS. RNA–protein interactions in unstable microsatellite diseases. Brain Res. 2014;1584:3–14.

    Article  CAS  PubMed  Google Scholar 

  138. Haeusler AR, Donnelly CJ, Periz G, Simko EAJ, Shaw PG, Kim M-S, et al. C9orf72 nucleotide repeat structures initiate molecular cascades of disease. Nature. 2014;507:195–200.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. O’Rourke JG, Bogdanik L, Muhammad AKMG, Gendron TF, Kim KJ, Austin A, et al. C9orf72 BAC transgenic mice display typical pathologic features of ALS/FTD. Neuron. 2015;88:892–901.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  140. Zu T, Gibbens B, Doty NS, Gomes-Pereira M, Huguet A, Stone MD, et al. Non-ATG–initiated translation directed by microsatellite expansions. Proc Natl Acad Sci. 2011;108:260–5.

    Article  CAS  PubMed  Google Scholar 

  141. Chai N, Gitler AD. Yeast screen for modifiers of C9orf72 poly(glycine-arginine) dipeptide repeat toxicity. FEMS Yeast Res. 2018;18 Available from: https://academic.oup.com/femsyr/article/doi/10.1093/femsyr/foy024/4925064. [cited 2018 Sep 16].

  142. Zu T, Liu Y, Bañez-Coronel M, Reid T, Pletnikova O, Lewis J, et al. RAN proteins and RNA foci from antisense transcripts in C9ORF72 ALS and frontotemporal dementia. Proc Natl Acad Sci. 2013;110:E4968–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Donnelly CJ, Zhang P-W, Pham JT, Haeusler AR, Mistry NA, Vidensky S, et al. RNA toxicity from the ALS/FTD C9ORF72 expansion is mitigated by antisense intervention. Neuron. 2013;80:415–28.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Targeting RNA Foci in iPSC-Derived Motor Neurons from ALS Patients with a C9ORF72 Repeat Expansion | Science Translational Medicine [Internet]. Available from: http://stm.sciencemag.org/content/5/208/208ra149. [cited 2018 Dec 1].

  145. Almeida S, Gascon E, Tran H, Chou HJ, Gendron TF, DeGroot S, et al. Modeling key pathological features of frontotemporal dementia with C9ORF72 repeat expansion in iPSC-derived human neurons. Acta Neuropathol (Berl). 2013;126:385–99.

    Article  CAS  Google Scholar 

  146. Liu Y, Pattamatta A, Zu T, Reid T, Bardhi O, Borchelt DR, et al. C9orf72 BAC mouse model with motor deficits and neurodegenerative features of ALS/FTD. Neuron. 2016;90:521–34.

    Article  CAS  PubMed  Google Scholar 

  147. Jiang J, Zhu Q, Gendron TF, Saberi S, McAlonis-Downes M, Seelman A, et al. Gain of toxicity from ALS/FTD-linked repeat expansions in C9ORF72 is alleviated by antisense oligonucleotides targeting GGGGCC-containing RNAs. Neuron. 2016;90:535–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Mizielinska S, Grönke S, Niccoli T, Ridler CE, Clayton EL, Devoy A, et al. C9orf72 repeat expansions cause neurodegeneration in drosophila through arginine-rich proteins. Science. 2014;345:1192–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Boeynaems S, Bogaert E, Michiels E, Gijselinck I, Sieben A, Jovičić A, et al. Drosophila screen connects nuclear transport genes to DPR pathology in c9ALS/FTD. Sci Rep. 2016;6:20877.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Jovičić A, Mertens J, Boeynaems S, Bogaert E, Chai N, Yamada SB, et al. Modifiers of C9orf72 dipeptide repeat toxicity connect nucleocytoplasmic transport defects to FTD/ALS. Nat Neurosci. 2015;18:1226–9.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  151. Lin Y, Mori E, Kato M, Xiang S, Wu L, Kwon I, et al. Toxic PR poly-dipeptides encoded by the C9orf72 repeat expansion target LC domain polymers. Cell. 2016;167:789–802.e12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Lee K-H, Zhang P, Kim HJ, Mitrea DM, Sarkar M, Freibaum BD, et al. C9orf72 dipeptide repeats impair the assembly, dynamics, and function of membrane-less organelles. Cell. 2016;167:774–788.e17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Freibaum BD, Lu Y, Lopez-Gonzalez R, Kim NC, Almeida S, Lee K-H, et al. GGGGCC repeat expansion in C9orf72 compromises nucleocytoplasmic transport. Nature. 2015;525:129–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Zhang K, Donnelly CJ, Haeusler AR, Grima JC, Machamer JB, Steinwald P, et al. The C9orf72 repeat expansion disrupts nucleocytoplasmic transport. Nature. 2015;525:56–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Moore S, Alsop E, Lorenzini I, Starr A, Rabichow BE, Mendez E, et al. ADAR2 mislocalization and widespread RNA editing aberrations in C9orf72-mediated ALS/FTD. Acta Neuropathol (Berl). 2019;138:49–65.

    Article  CAS  Google Scholar 

  156. Rosen DR, Siddique T, Patterson D, Figlewicz DA, Sapp P, Hentati A, et al. Mutations in Cu/Zn superoxide dismutase gene are associated with familial amyotrophic lateral sclerosis. Nature. 1993;362:59–62.

    Article  CAS  PubMed  Google Scholar 

  157. Chang Y, Kong Q, Shan X, Tian G, Ilieva H, Cleveland DW, et al. Messenger RNA oxidation occurs early in disease pathogenesis and promotes motor neuron degeneration in ALS. PLoS One. 2008;3:e2849.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  158. Zhang J, Ito H, Wate R, Ohnishi S, Nakano S, Kusaka H. Altered distributions of nucleocytoplasmic transport-related proteins in the spinal cord of a mouse model of amyotrophic lateral sclerosis. Acta Neuropathol (Berl). 2006;112:673–80.

    Article  CAS  Google Scholar 

  159. Kinoshita Y, Ito H, Hirano A, Fujita K, Wate R, Nakamura M, et al. Nuclear contour irregularity and abnormal transporter protein distribution in anterior horn cells in amyotrophic lateral sclerosis. J Neuropathol Exp Neurol. 2009;68:1184–92.

    Article  CAS  PubMed  Google Scholar 

  160. al A-RA et. Inositol hexakisphosphate and Gle1 activate the DEAD-box protein Dbp5 for nuclear mRNA export. - PubMed - NCBI [Internet]. [cited 2018 Dec 2]. Available from: https://www.ncbi.nlm.nih.gov/pubmed/16783363.

  161. Nousiainen HO, Kestilä M, Pakkasjärvi N, Honkala H, Kuure S, Tallila J, et al. Mutations in mRNA export mediator GLE1 result in a fetal motoneuron disease. Nat Genet. 2008;40:155–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  162. Kaneb HM, Folkmann AW, Belzil VV, Jao L-E, Leblond CS, Girard SL, et al. Deleterious mutations in the essential mRNA metabolism factor, hGle1, in amyotrophic lateral sclerosis. Hum Mol Genet. 2015;24:1363–73.

    Article  CAS  PubMed  Google Scholar 

  163. Folkmann AW, Collier SE, Zhan X, Null A, Ohi MD, Wente SR. Gle1 functions during mRNA export in an oligomeric complex that is altered in human disease. Cell. 2013;155:582–93.

    Article  CAS  PubMed  Google Scholar 

  164. Hautbergue GM. RNA nuclear export: from neurological disorders to cancer. In: El-Khamisy S, editor. Pers Med. Cham: Springer International Publishing; 2017. p. 89–109. Available from: http://link.springer.com/10.1007/978-3-319-60733-7_6. [cited 2018 Sep 16].

    Chapter  Google Scholar 

  165. Landles C, Bates GP. Huntingtin and the molecular pathogenesis of Huntington’s disease. EMBO Rep. 2004;5:958–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Jimenez-Sanchez M, Licitra F, Underwood BR, Rubinsztein DC. Huntington’s disease: mechanisms of pathogenesis and therapeutic strategies. Cold Spring Harb Perspect Med. 2017;7.

  167. Maiuri T, Woloshansky T, Xia J, Truant R. The huntingtin N17 domain is a multifunctional CRM1 and Ran-dependent nuclear and cilial export signal. Hum Mol Genet. 2013;22:1383–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Polyglutamine-expanded huntingtin exacerbates age-related disruption of nuclear integrity and nucleocytoplasmic transport - ScienceDirect [Internet]. Available from: https://www.sciencedirect.com/science/article/pii/S0896627317302398. [cited 2018 Dec 2].

  169. Kumar A, Ratan RR. Oxidative stress and Huntington’s disease: the good, the bad, and the ugly. J Huntingt Dis. 2016;5:217–37.

    Article  Google Scholar 

  170. Bañez-Coronel M, Ayhan F, Tarabochia AD, Zu T, Perez BA, Tusi SK, et al. RAN translation in Huntington disease. Neuron. 2015;88:667–77.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  171. Manto M-U. The wide spectrum of spinocerebellar ataxias (SCAs). Cerebellum. 2005;4:2.

    Article  CAS  PubMed  Google Scholar 

  172. Sowa AS, Martin E, Martins IM, Schmidt J, Depping R, Weber JJ, et al. Karyopherin α-3 is a key protein in the pathogenesis of spinocerebellar ataxia type 3 controlling the nuclear localization of ataxin-3. Proc Natl Acad Sci. 2018;115:E2624–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Bennett CL, Dastidar SG, Ling S-C, Malik B, Ashe T, Wadhwa M, et al. Senataxin mutations elicit motor neuron degeneration phenotypes and yield TDP-43 mislocalization in ALS4 mice and human patients. Acta Neuropathol (Berl). 2018;136:425–43.

    Article  CAS  Google Scholar 

  174. Nguyen L, Cleary JD, Ranum LPW. Repeat associated non-ATG translation: molecular mechanisms and contribution to neurologic disease. Annu Rev Neurosci. 2019;42:227–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Hagerman R, Hagerman P. Advances in clinical and molecular understanding of the FMR1 premutation and fragile X-associated tremor/ataxia syndrome. Lancet Neurol. 2013;12:786–98.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Cohen S, Masyn K, Adams J, Hessl D, Rivera S, Tassone F, et al. Molecular and imaging correlates of the fragile X-associated tremor/ataxia syndrome. Neurology. 2006;67:1426–31.

    Article  CAS  PubMed  Google Scholar 

  177. Edens BM, Vissers C, Su J, Arumugam S, Xu Z, Shi H, et al. FMRP modulates neural differentiation through m6A-dependent mRNA nuclear export. Cell Rep. 2019;28:845–854.e5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Arnold FJ, Pluciennik A, Merry DE. Impaired nuclear export of polyglutamine-expanded androgen receptor in spinal and bulbar muscular atrophy. Sci Rep. 2019;9 Available from: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6333819/. [cited 2020 Feb 6].

  179. Nagafuchi S, Yanagisawa H, Ohsaki E, Shirayama T, Tadokoro K, Inoue T, et al. Structure and expression of the gene responsible for the triplet repeat disorder, dentatorubral and pallidoluysian atrophy (DRPLA). Nat Genet. 1994;8:177–82.

    Article  CAS  PubMed  Google Scholar 

  180. Takahashi H, Egawa S, Piao YS, Hayashi S, Yamada M, Shimohata T, et al. Neuronal nuclear alterations in dentatorubral-pallidoluysian atrophy: ultrastructural and morphometric studies of the cerebellar granule cells. Brain Res. 2001;919:12–9.

    Article  CAS  PubMed  Google Scholar 

  181. Moreira M-C, Klur S, Watanabe M, Németh AH, Ber IL, Moniz J-C, et al. Senataxin, the ortholog of a yeast RNA helicase, is mutant in ataxia-ocular apraxia 2. Nat Genet. 2004;36:225–7.

    Article  CAS  PubMed  Google Scholar 

  182. Hirano M, Quinzii CM, Mitsumoto H, Hays AP, Roberts JK, Richard P, et al. Senataxin mutations and amyotrophic lateral sclerosis. Amyotroph Lateral Scler Off Publ World Fed Neurol Res Group Mot Neuron Dis. 2011;12:223–7.

    CAS  Google Scholar 

  183. Moreira M-C, Barbot C, Tachi N, Kozuka N, Uchida E, Gibson T, et al. The gene mutated in ataxia-ocular apraxia 1 encodes the new HIT/Zn-finger protein aprataxin. Nat Genet. 2001;29:189–93.

    Article  CAS  PubMed  Google Scholar 

  184. Hirano M, Furiya Y, Asai H, Yasui A, Ueno S. ALADINI482S causes selective failure of nuclear protein import and hypersensitivity to oxidative stress in triple a syndrome. Proc Natl Acad Sci. 2006;103:2298–303.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Zanni G, De Magistris P, Nardella M, Bellacchio E, Barresi S, Sferra A, et al. Biallelic variants in the nuclear pore complex protein NUP93 are associated with non-progressive congenital ataxia. Cerebellum Lond Engl. 2019;18:422–32.

    Article  CAS  Google Scholar 

  186. Lotharius J, Brundin P. Pathogenesis of Parkinson’s disease: dopamine, vesicles and α-synuclein. Nat Rev Neurosci. 2002;3:932–42.

    Article  CAS  PubMed  Google Scholar 

  187. Pinho R, Paiva I, Jercic KG, Fonseca-Ornelas L, Gerhardt E, Fahlbusch C, et al. Nuclear localization and phosphorylation modulate pathological effects of alpha-synuclein. Hum Mol Genet. 2019;28:31–50.

    Article  CAS  PubMed  Google Scholar 

  188. Dias V, Junn E, Mouradian MM. The role of oxidative stress in Parkinson’s disease. J Park Dis. 2013;3:461–91.

    CAS  Google Scholar 

  189. Hunot S, Brugg B, Ricard D, Michel PP, Muriel M-P, Ruberg M, et al. Nuclear translocation of NF-κB is increased in dopaminergic neurons of patients with Parkinson disease. Proc Natl Acad Sci U S A. 1997;94:7531–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  190. Chalovich EM, Zhu J, Caltagarone J, Bowser R, Chu CT. Functional repression of cAMP response element in 6-hydroxydopamine-treated neuronal cells. J Biol Chem. 2006;281:17870–81.

    Article  CAS  PubMed  Google Scholar 

  191. Um JW, Min DS, Rhim H, Kim J, Paik SR, Chung KC. Parkin ubiquitinates and promotes the degradation of RanBP2. J Biol Chem. 2006;281:3595–603.

    Article  CAS  PubMed  Google Scholar 

  192. Petrucelli L, O’Farrell C, Lockhart PJ, Baptista M, Kehoe K, Vink L, et al. Parkin protects against the toxicity associated with mutant α-Synuclein: proteasome dysfunction selectively affects catecholaminergic neurons. Neuron. 2002;36:1007–19.

    Article  CAS  PubMed  Google Scholar 

  193. Beyer K, Domingo-Sàbat M, Humbert J, Carrato C, Ferrer I, Ariza A. Differential expression of alpha-synuclein, parkin, and synphilin-1 isoforms in Lewy body disease. Neurogenetics. 2008;9:163–72.

    Article  CAS  PubMed  Google Scholar 

  194. Dawson TM. Parkin and defective ubiquitination in Parkinson’s disease. In: Riederer P, Reichmann H, Youdim MBH, Gerlach M, editors. Park Dis Relat Disord. Vienna: Springer; 2006. p. 209–13.

    Chapter  Google Scholar 

  195. Langston JW. The MPTP story. J Park Dis. 2017;7:S11–9.

    Google Scholar 

  196. Cho K-I, Searle K, Webb M, Yi H, Ferreira PA. Ranbp2 haploinsufficiency mediates distinct cellular and biochemical phenotypes in brain and retinal dopaminergic and glia cells elicited by the parkinsonian neurotoxin, 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP). Cell Mol Life Sci CMLS. 2012;69:3511–27.

    Article  CAS  PubMed  Google Scholar 

  197. Scheltens P, Blennow K, Breteler MMB, de Strooper B, Frisoni GB, Salloway S, et al. Alzheimer’s disease. Lancet. 2016;388:505–17.

    Article  CAS  PubMed  Google Scholar 

  198. Metuzals J, Robitaille Y, Houghton S, Gauthier S, Leblanc R. Paired helical filaments and the cytoplasmic-nuclear interface in Alzheimer’s disease. J Neurocytol. 1988;17:827–33.

    Article  CAS  PubMed  Google Scholar 

  199. Mulvihill P, Perry G. Immunoaffinity demonstration that paired helical filaments of Alzheimer disease share epitopes with neurofilaments, MAP2 and tau. Brain Res. 1989;484:150–6.

    Article  CAS  PubMed  Google Scholar 

  200. Sheffield LG, Miskiewicz HB, Tannenbaum LB, Mirra SS. Nuclear pore complex proteins in Alzheimer disease. J Neuropathol Exp Neurol. 2006;65:45–54.

    Article  CAS  PubMed  Google Scholar 

  201. Mastroeni D, Chouliaras L, Grover A, Liang WS, Hauns K, Rogers J, et al. Reduced RAN expression and disrupted transport between cytoplasm and nucleus; A Key Event in Alzheimer’s Disease Pathophysiology. PLoS One. 2013;8:e53349.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  202. Tripathi T, Kalita J. Abnormal microtubule dynamics impair the nuclear-cytoplasmic transport in dementia. ACS Chem Neurosci. 2019;10:1133–4.

    Article  CAS  PubMed  Google Scholar 

  203. Paonessa F, Evans LD, Solanki R, Larrieu D, Wray S, Hardy J, et al. Microtubules deform the nuclear membrane and disrupt nucleocytoplasmic transport in Tau-mediated frontotemporal dementia. Cell Rep. 2019;26:582–593.e5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  204. Montalbano M, McAllen S, Sengupta U, Puangmalai N, Bhatt N, Ellsworth A, et al. Tau oligomers mediate aggregation of RNA-binding proteins Musashi1 and Musashi2 inducing Lamin alteration. Aging Cell. 2019;18 Available from: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6826126/. [cited 2020 Feb 6].

  205. Allgrove J, Clayden GS, Grant DB, Macaulay JC. Familial glucocorticoid deficiency with achalasia of the cardia and deficient tear production. Lancet Lond Engl. 1978;1:1284–6.

    Article  CAS  Google Scholar 

  206. Prasad R, Metherell LA, Clark AJ, Storr HL. Deficiency of ALADIN impairs redox homeostasis in human adrenal cells and inhibits steroidogenesis. Endocrinology. 2013;154:3209–18.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  207. Bitetto G, Ronchi D, Bonato S, Pittaro A, Compagnoni GM, Bordoni A, et al. Loss of the nucleoporin Aladin in central nervous system and fibroblasts of Allgrove syndrome. Hum Mol Genet. 2019. https://doi.org/10.1093/hmg/ddz236.

  208. Liguori I, Russo G, Curcio F, Bulli G, Aran L, Della-Morte D, et al. Oxidative stress, aging, and diseases. Clin Interv Aging. 2018;13:757–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  209. Wang X, Michaelis EK. Selective neuronal vulnerability to oxidative stress in the brain. Front Aging Neurosci. 2010;2 Available from: https://www.frontiersin.org/articles/10.3389/fnagi.2010.00012/full. [cited 2020 Feb 7].

  210. Beckman KB, Ames BN. The free radical theory of aging matures. Physiol Rev. 1998;78:547–81.

    Article  CAS  PubMed  Google Scholar 

  211. Chandrasekaran A, Idelchik MDPS, Melendez JA. Redox control of senescence and age-related disease. Redox Biol. 2017;11:91–102.

    Article  CAS  PubMed  Google Scholar 

  212. Ferri G, Storti B, Bizzarri R. Nucleocytoplasmic transport in cells with progerin-induced defective nuclear lamina. Biophys Chem. 2017;229:77–83.

    Article  CAS  PubMed  Google Scholar 

  213. Dworak N, Makosa D, Chatterjee M, Jividen K, Yang C-S, Snow C, et al. A nuclear lamina-chromatin-Ran GTPase axis modulates nuclear import and DNA damage signaling. Aging Cell. 2019;18:e12851.

    Article  PubMed  CAS  Google Scholar 

  214. Fifková E, Tonks M, Cullen-Dockstader K. Changes in the nuclear pore complexes of the dentate granule cells in aged rats. Exp Neurol. 1987;95:755–62.

    Article  PubMed  Google Scholar 

  215. D’Angelo MA, Raices M, Panowski SH, Hetzer MW. Age-dependent deterioration of nuclear pore complexes causes a loss of nuclear integrity in postmitotic cells. Cell. 2009;136:284–95.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  216. Toyama BH, Savas JN, Park SK, Harris MS, Ingolia NT, Yates JR, et al. Identification of long-lived proteins reveals exceptional stability of essential cellular structures. Cell. 2013;154:971–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  217. Mertens J, Paquola ACM, Ku M, Hatch E, Böhnke L, Ladjevardi S, et al. Directly reprogrammed human neurons retain aging-associated transcriptomic signatures and reveal age-related nucleocytoplasmic defects. Cell Stem Cell. 2015;17:705–18.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors would like to acknowledge the Fresco Institute for support.

Funding

None.

Author information

Authors and Affiliations

Authors

Contributions

GB, ADF: reviewing the literature, writing the manuscript. The author(s) read and approved the final manuscript.

Corresponding author

Correspondence to Alessio Di Fonzo.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bitetto, G., Di Fonzo, A. Nucleo–cytoplasmic transport defects and protein aggregates in neurodegeneration. Transl Neurodegener 9, 25 (2020). https://doi.org/10.1186/s40035-020-00205-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40035-020-00205-2

Keywords