Phase to Phase with TDP-43 - Biochemistry (ACS Publications)

Jan 23, 2017 - TDP-43 is a dimeric nuclear protein that plays a central role in RNA metabolism. In recent years, this protein has become a focal point...
0 downloads 10 Views 3MB Size
Subscriber access provided by UB + Fachbibliothek Chemie | (FU-Bibliothekssystem)

Current Topic/Perspective

Phase to phase with TDP-43 Yulong Sun, and Avijit Chakrabartty Biochemistry, Just Accepted Manuscript • DOI: 10.1021/acs.biochem.6b01088 • Publication Date (Web): 23 Jan 2017 Downloaded from http://pubs.acs.org on January 24, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Biochemistry is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

Phase to phase with TDP-43 Funding Source Statement: Canadian Consortium of Neurodegeneration and Aging (CCNA), Canadian Institute of Health Research (CIHR), ALS Society of Canada (ALS Canada), Alzheimer Society of Canada (ASC) Yulong Sun1 and Avijit Chakrabartty1, 2* 1: Department of Medical Biophysics, University of Toronto. 2: Department of Biochemistry, University of Toronto. *To whom correspondence may be addressed: Dr. Avi Chakrabartty. University Health Network. Princess Margaret Cancer Research Tower. 101 College Street, Room 4-305. Toronto, Ontario, M5G1L7. Phone/Fax: (416)-581-7554. Email: [email protected]

1 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

LIST OF ABBREVIATIONS Aβ, amyloid-beta; AD, Alzheimer's disease; ALS, amyotrophic lateral sclerosis; bvFTLD, behavioral variant FTLD; CDK6, cyclin-dependent kinase 6; CFTR, cystic fibrosis transmembrane regulator; CTD, C-terminal domain; fALS, familial ALS; FTD, frontotemporal dementia; FTLD, frontotemporal lobar degeneration; FUS/TSL, Fused in Sarcoma/Translocated in Sarcoma; HIV-1, human immunodeficiency virus type 1; hnRNP, heterogeneous nuclear ribonucleoprotein; LLPS, liquid-liquid phase separation; lncRNA, long non-coding RNA; LTR, long terminal repeat; ncRNA, non-coding RNA; NES, nuclear export signal; NLS, nuclear localization signal; NMD, nonsense mediated decay; NTD, N-terminal domain; PLAAC, prionlike amino acid composition; PNFA, primary non-fluent aphasia; pRb, retinoblastoma protein; PrLD, prion-like domain; PrP, prion protein; RNP, ribonucleoprotein; RRM, RNA recognition motif; sALS, sporadic ALS; SAXS, small angle X-ray scattering; SD, semantic dementia; SG, stress granule; TAR, transactive response; TDP-43, TAR element DNA binding protein of 43 kDa; TTR, transthyretin

2 ACS Paragon Plus Environment

Page 2 of 62

Page 3 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

ABSTRACT TDP-43 is a dimeric nuclear protein that plays a central role in RNA metabolism. In recent years, this protein has become a focal point of research in the amyotrophic lateral sclerosis and frontotemporal dementia (ALS/FTD) disease spectrum, as pathognomonic inclusions within affected neurons contain post-translationally modified TDP-43. A key question in TDP-43 research involves determining the mechanisms and triggers that cause TDP-43 to form pathological aggregates. This review gives a brief overview of the physiological and pathological roles of TDP-43 and focuses on the structural features of its protein domains and how they may contribute to normal protein function and to disease. A special emphasis is placed on the C-terminal prion-like region thought to be implicated in pathology, as it is where nearly all ALS/FTD-associated mutations reside. Recent structural studies on this domain revealed its crucial role in the formation of phase-separated liquid droplets through a partially populated α-helix. This new discovery provides further support for the theory that liquid droplets such as stress granules may be precursors to pathological aggregates, linking environmental effects such as stress to the potential etiology of the disease. The transition of TDP-43 between soluble, droplet, and aggregate phases and the implications of these transitions on pathological aggregation are summarized and discussed.

3 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

INTRODUCTION Recent advances in modern medicine have led to dramatically longer lifespans in the human population. Ironically, through this increase in life expectancy, the imperfections in our evolutionary mechanisms that maintain the robustness of proteins appear to have been exposed. Molecular evolution has produced tremendously sturdy proteins that have little to no turnover in an organism’s lifetime such as proteins of the lens. However, even these proteins can accumulate modifications and aggregate in individuals of advanced age, leading to the disruption of optical properties of the lens and senile cataracts.1 As suggested by Christopher Dobson, the rapid increase in human life expectancy may have out-paced molecular evolution, giving rise to numerous diseases that are caused by protein misfolding and aggregation.2,3 Many of these diseases, such as senile cardiac amyloidosis caused by the deposition of the protein transthyretin (TTR) in heart tissue, and certain forms of cancer where the key regulatory protein p53 is known to aggregate, have only emerged in the past century.4–6 Neurons seem particularly vulnerable to late-onset protein misfolding diseases, possibly due to the lack of neuronal cellular turnover and age-dependent deficits in protein quality control. Indeed, the abnormal accumulation of protein in affected neurons has emerged as a common hallmark of neurodegenerative diseases. Identification and characterization of major protein components of these aggregates has often lead to transformative breakthroughs in uncovering the mechanisms of disease pathogenesis. Classic examples include the discovery of the prion protein (PrP) and formulation of the proteinonly hypothesis of prion disease and the identification and study of the amyloid-beta (Aβ) peptide in formulating the amyloid cascade model of Alzheimer’s disease (AD).7,8 This review will focus on the molecular mechanisms of misfolding and aggregation of transactive response (TAR) element DNA binding protein of 43 kDa (TDP-43), a key protein found in neuronal

4 ACS Paragon Plus Environment

Page 4 of 62

Page 5 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

inclusions of patients with amyotrophic lateral sclerosis (ALS) and frontotemporal lobar degeneration (FTLD). ALS/FTD is a spectrum disorder ALS, also known as Lou Gehrig’s disease, is a devastating neurological disorder characterized by the progressive loss of lower motor neurons in the anterior horn of the spinal cord and upper motor neurons in the motor cortex, leading to initial paralysis of extremities followed by fatal paralysis of the diaphragm.9,10 ALS has a prevalence of 3.9 cases per 100,000 in the United States.11 The disease is incurable and progresses rapidly, resulting in an average life expectancy of 3-5 years after disease onset, usually occurring in mid-adult life.10,12 90% of ALS cases are sporadic (sALS) with unknown etiology, while 10% of cases have family history (fALS), although the two are clinically indistinguishable. Distinct from ALS, FTLD, presented clinically as frontotemporal dementia (FTD), belongs to a broad range of disorders leading to progressive cognitive, behavioral, and/or language deficits.13 It is the second most common form of dementia in people younger than 65 years of age after AD.14 Clinically, FTD can present as behavioral variant FTLD (bvFTLD) with predominantly behavioral changes, primary non-fluent aphasia (PNFA), affecting speech, or semantic dementia (SD), affecting comprehension.15 Approximately 30-50% of FTD cases show family history, with the majority of cases caused by mutations in three major genes: microtubule protein tau (MAPT), progranulin (GRN) and chromosome 9 open reading frame 72 (C9ORF72).16–21 To date, a large number of genes have been identified to be causative for ALS and FTD (reviewed in 22). Some of them are related to clinically pure ALS or FTD, but a large portion of genes are found in both diseases, suggesting a common disease mechanism. Additionally, despite distinction in disease presentation, co-occurrences of ALS and FTD have been widely reported.

5 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

About 15% of FTD patients develop ALS and about 50% of ALS patients show some signs of cognitive impairment, meeting diagnostic criteria of FTD in about 5% cases, indicating significant clinical overlap.15,23 A major breakthrough linking the two diseases occurred in 2006, where TDP-43 was found to be ubiquitinated, hyperphosphorylated, and fragmented in neuronal inclusions of patients with sporadic and familial forms of ALS and FTD.24,25 Due to their clinical, genetic and pathological overlaps, it is now believed that the two diseases belong to a spectrum of disorders termed TDP-43 proteinopathies and that the disease phenotype arises from differences in the primary sites of neurodegeneration: motor neurons in ALS and cortical neurons in FTD. Despite the numerous genes involved in the disease spectrum, the fact that aggregates of TDP-43 have now been found in 97% of cases of ALS and 45% of cases of FTD suggests that it is directly linked to the disease mechanism.26 The study of TDP-43’s folding and aggregation is therefore invaluable to determining the cause of ALS/FTD. The identification of TDP-43 as a major component of ALS/FTD pathology catapulted the investigation of the protein’s structure, function (native and pathological), and biophysical characteristics. Since the discovery of its involvement in ALS/FTD ten years ago, over 1700 publications on TDP-43 have been produced. The biophysical features of this multi-domain dimeric protein and how the behavior of these domains may contribute to disease pathogenesis are discussed herein.

PHYSIOLOGICAL FUNCTIONS AND PATHOBIOLOGY OF TDP-43 Physiological functions of TDP-43 TDP-43 was first discovered as a ubiquitously expressed cellular factor that binds to the TAR element, an element in the long terminal repeat (LTR) region of human immunodeficiency virus type 1 (HIV-1) which is critical for the control of gene expression in the virus.27 The 43 kDa protein was named for its function as TAR DNA binding protein of 43 kDa upon its initial

6 ACS Paragon Plus Environment

Page 6 of 62

Page 7 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

discovery. The role of the protein in human disease was first studied in the early 2000, where it was found to regulate splicing of the gene coding for the cystic fibrosis transmembrane regulator (CFTR) protein exon 9 by binding to a region near its 3’-splice site consisting of repeats of UG nucleotides.28 TDP-43 binds preferentially to TG-rich and UG-rich sequences (vide infra) and appears to be involved in a number of roles in splicing and transcription.28–33 To date, TDP-43 has been found to participate in a large number of nuclear and cytoplasmic functions as it is shuttled between the two cellular milieus.34 In brief, TDP-43 is known to be involved in premRNA processing and splicing, microRNA processing and regulation, control of long noncoding RNA (lncRNA) and non-coding RNA (ncRNA) expression, mRNA transport, mRNA stability through recruitment into stress granules (SGs) and mRNA translation (reviewed in 35). TDP-43 is also involved in various aspects of cell proliferation and apoptosis. It regulates the phosphorylation of retinoblastoma protein (pRb), a tumour suppressor dysfunctional in several major cancers, through the repression of cyclin-dependent kinase 6 (CDK6) expression.36 Mutant forms of TDP-43 are also more prone to cause neural apoptosis in chick embryos.37 Disruption to cell cycle and apoptotic proteins by TDP-43 mutations may implicate the protein in neuronal cancers. Due to its many functions, TDP-43 levels are tightly regulated through a negative feedback loop. TDP-43 binds to the 3’UTR of its own mRNA, leading to nonsense mediated decay (NMD)-independent mRNA degradation and reduction in TDP-43 production.38–40 Recent findings also suggest that RNA/DNA binding modulates TDP-43 solubility.41–43 In cells, TDP-43 localizes to the nucleus in both diffuse and speckled distributions.44 During stress response to heat shock or sodium arsenite, TDP-43 coalesces into SGs in the cytoplasm and modulates SG assembly and dynamics.45–49 Alterations to these SG processes have been suggested to play a key role in TDP-43 aggregation and pathology.50–52 The involvement of SGs also links

7 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

environmental effects to protein aggregation, which may provide an explanation for the mostly sporadic nature of ALS (vide infra). Given these crucial RNA-related functions of TDP-43, it is not surprising that homozygous knockouts of TDP-43 are embryonically lethal in mouse models while heterozygous mice are not as affected, possibly due to the ability for TDP-43 to tightly control its expression levels through negative feedback.53,54 The gene encoding TDP-43 protein (TARDBP) is also highly conserved in human, mouse, Drosophila melanogaster, and Caenorhabditis elegans, with very low rates of divergence between the four eukaryotes, suggesting that TDP-43 likely has crucial roles as a gene regulator.55 TDP-43’s many functions have led to the suggestion that disruptions to TDP-43 expression level and function is at least partially responsible for neurotoxicity. Pathological functions of TDP-43 Aggregation of wild type TDP-43 primarily in the cytoplasm of neurons is a prominent feature in ALS/FTD, and research on the mechanistic relationship between aggregation and disease is ongoing. In pathognomonic, cytoplasmic aggregates, TDP-43 is aberrantly ubiquitinated, phosphorylated, acetylated, sumoylated, and cleaved into C-terminal fragments.24,25,56,57 The nature of how these post translational modifications relate to disease pathology is still under investigation. Unlike AD and prion disease, the aggregates in ALS/FTD neurons are amorphous and non-amyloid, and TDP-43 aggregates created in vitro and in vivo often share this property.43,58,59 Cytoplasmic aggregation is usually accompanied by depletion of native TDP-43 from the nucleus as well as sequestration of other RNA-binding proteins into these aggregates.60– 63

Whether depletion of TDP-43 and other RNA-binding proteins can cause RNA disruption and

the extent of this disruption on neurotoxicity still remains unclear. Current evidence suggests that the cytoplasmic aggregates themselves are toxic to cells and cause cell death through a toxic gain

8 ACS Paragon Plus Environment

Page 8 of 62

Page 9 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

of function, although alternative theories of TDP-43 aggregates as cytoprotective structures do exist in Drosophila models.58,64–68 The current consensus is that the disease likely arises from a combination of loss of TDP-43 native function and gain of toxic function from aggregation.69 A number of factors can contribute to the aggregation of the protein, including cytoplasmic accumulation, changes in TDP-43 expression levels, aberrant cleavage and fragmentation, loss of native state binding partners, or the production of truncated isoforms through alternative splicing.39,43,70–72 Environmental factors such as stress have long been suspected as a contributing factor in ALS/FTD pathogenesis, and evidence for this suspicion has recently been growing. One mechanism mammalian cells use to cope with environmental stress is to transiently repress the translation of mRNAs for proteins not essential to survival by organizing these arrested mRNAs and their RNA binding proteins into small (≤ 5 µm) non-membrane bound cytoplasmic domains called SGs. The assembly of SGs can be induced by oxidative, genotoxic, osmotic and thermal stresses.73,74 SG assembly and disassembly are dynamic processes mediated by a number of proteins, including TDP-43.48,75,76 Knockdown of TDP-43 reduces the expression of levels of G3BP and increases TIA-1 levels, proteins known to affect SG assembly, causing SGs to form slower, take more time to reach the average size of normal SGs, and dissipates more quickly.49 The disruption to SG regulation and persistence is predicted to cause cytoplasmic inclusions similar to those observed in ALS/FTD neurons.47 How TDP-43 mutations impact the in vivo response to stress in motor neurons is disease relevant and remains to be explored. It has been suggested that a predisposing event that enriches a population of cytoplasmic, aggregation-prone TDP-43 (through mutation or other events) followed by chronic environmental stress and persistence of SGs can cause normally reversible, functional TDP-43 aggregation in SGs to form irreversible aggregates as seen in disease. However, it is still unclear whether SGs are direct

9 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

precursors to TDP-43 aggregates, or whether they are formed independently and recruited to SGs afterwards.52 The assembly of these membraneless organelles, also known as ribonucleoprotein (RNP) granules, occurs through the physical process of liquid-liquid phase separation (LLPS).77 Current research into the formation of these RNP granules has made the SG precursor hypothesis increasingly popular, as it serves as a long missing link between environmental effects and ALS/FTD pathology. Recent structural studies of TDP-43 have shed light on the potentially detailed molecular mechanisms of how TDP-43 diverges from folding into reversible RNP granule assemblies versus pathological aggregates.

INSIGHTS INTO TDP-43 AGGREGATION FROM STRUCTURAL STUDIES Domain structure of TDP-43 TDP-43 is 414 amino acid residues in length, and is comprised of an N-terminal domain (NTD: 1-102) that includes a predicted nuclear localization signal (NLS: 82-98), two RNA recognition motifs (RRMs) composed of residues 106-177 (RRM1), and residues 192-259 (RRM2) which includes a nuclear export signal (NES) from residues 239-250, and a C-terminal domain (CTD: 274-414) (Figure 1A).34,78 In vitro biophysical characterization and crosslinking studies in cell culture and mouse brains all suggest that TDP-43 is intrinsically a dimeric protein and that dimer formation may be mediated by a number of regions across the entirety TDP-43, including the NTD, RRM2, and/or the CTD.43,44,79–81 The CTD is particularly disease relevant, as it is where nearly all ALS/FTD-associated mutations are found.37,82–123 This is a flexible region containing only a transient α-helical structure, and contains QN-rich residues implicated in aggregation (Figure 1D). The structural study of full length TDP-43 has been difficult due to its high aggregation propensity and difficulty of purification, as well as its flexible CTD.64 No crystal or 10 ACS Paragon Plus Environment

Page 10 of 62

Page 11 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

NMR structures have been produced for the protein in its entirety but structures of the individual domains of TDP-43 have been solved. The ubiquitin-like fold of the NTD The N-terminus of TDP-43 contains a ubiquitin-like fold from residues 1-77 and a nonstructured region from residues 78-102, but depending on experimental conditions, NMR spectroscopy of residues 1-77 from two groups have reported conflicting results on the stability of this domain. While Qin et al. reports that it is at an equilibrium with an unstructured state, Mompean et al. reports a single stable structure at high resolution.124,125 In the latter condition, residues 1-77 appear to adopt a well-folded structure that consists of six β-strands and an α-helix arranged in β1β2α1β3β4β5β6 topology (Figure 1B, Figure 2A). The low resolution structure by Qin et al. reports a similar conformation except β4 and β5 were not observed and appeared as a single β strand. Strands β1β2β3 and β6 form one β sheet that is similar to ubiquitin, while a second smaller sheet composed of β4 and β5 appears to be a novel feature unique to TDP-43 NTD, resembling the structure of the C-terminal Dix domain of the scaffolding protein axin 1.125 The remaining residues (78-102) in the N-terminus is rich in positively charged amino acids. This region appears to bind non-specifically to DNA only at pH 4.0, likely due to charge-charge interactions.44,124,125 The relative position of the NTD to the rest of TDP-43 is unclear, but small angle X-ray scattering (SAXS) data suggests that it may dock closely to the tandem RRMs (Figure 2).126 The N-terminus is involved in aggregation and splicing The NTD of TDP-43 appears essential to TDP-43 normal function, but is also required for pathological aggregation. Cytosolic localization of ectopically expressed TDP-43 without a nuclear localization signal caused the formation of TDP-43 inclusions in HEK293T cells, but

11 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

expression of the same construct without the first 10 N-terminal residues, or constructs containing mutations of residues 6-9 (RVTE) to glycine residues showed diffuse cytoplasmic localization, suggesting that the N-terminus is required for aggregation.79 Additionally, the same mutations resulted in loss of TDP-43 splicing activity when the mutants were expressed in conjunction with knockdown of endogenous TDP-43 in cell culture.79 This suggests that loss or mutation of these first 10 residues may adversely affect the ability for TDP-43 to form its native dimeric structure and its ability to recruit proteins needed in the splicing machinery. This is expected since residues 6-8 also form the first β-strand in the N-terminal fold, and residues 6-9 are involved in stabilizing the first β-sheet of the N-terminal domain (Figure 1B). Mutations of these residues would disrupt the structure and dimerization of NTD, which agrees with predictions made by computer simulations.79 In a cellular aggregation model of TDP-43 where exogenous TDP-43 protein containing additional 12 tandem repeats of its aggregation-prone QNrich sequence at its C-terminus was expressed, aggregates resembling pathological inclusions formed and sequestered endogenous TDP-43, causing loss of TDP-43 exon skipping function.127 However, when the same construct without the N-terminal 75 residues was expressed, aggregates formed but without sequestration of native TDP-43 or loss of splicing function.128 Furthermore, the N-terminal TDP-43 fragment containing residues 1-105 also appears to oligomerize into larger species in a concentration dependent manner and the constructs containing the NTD plus RRM domains shows improved DNA binding activity compared to the tandem RRMs alone.81 Taken together, these evidence suggest that the NTD of TDP-43, and specifically the β-sheet structural motif, contributes to both native TDP-43 function as well as aggregation. This region of the protein is required for the initial dimerization of TDP-43 and recruitment of other RNA binding proteins, an event required for RNA splicing and perhaps the formation of RNP granules

12 ACS Paragon Plus Environment

Page 12 of 62

Page 13 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

such as SGs. The N-terminal region effectively increases the local concentration of the TDP-43 and other RNA binding proteins, which enhances RNA binding and splicing functions. On the other hand, this very mechanism of congregating proteins to close proximities may also serve as a prerequisite for aggregation of the protein or recruitment of native proteins into established aggregates.128 The only ALS/FTD-associated mutation in this region is A90V located at the predicted nuclear localization signal at residues 83-98.84,107 The location in the nuclear localization signal suggests a possible mechanism of pathology through disruption to nuclear localization and cytosolic accumulation. Tandem RRMs contain canonical folds but are uniquely arranged TDP-43 contains two RRMs in tandem, separated by a 15-amino acid linker. Initial co-crystal structures of individual RRM1 and RRM2 domains of TDP-43 bound to single-stranded DNA demonstrated that the structures of these RRMs and the molecular interactions involved in oligonucleotide binding are congruent with typical RRM domains.80,129,130 Structurally, they consist of a β-sheet sandwiched between two α-helices arranged in the β1α1β2β3α2β4β5 topology, where the β4 strand can also be referred to as β-hairpin.131 Two segments of 6 and 8 amino acids rich in aromatic residues on the β1 and β3 strands form the typical interacting surface on the β-sheet for nucleotide binding through direct stacking interactions with the ribonucleic bases, while a few amino acids on the loop regions between β1 and α1 (Loop1) as well as between β2 and β3 (Loop3) provide hydrogen bonding interactions (Figure 1B, 2B).80,132 Both RRMs share this canonical structure, except that RRM1 possesses a longer Loop3 region than RRM2, which is thought to contribute to RRM1’s higher affinity for targets due to the more numerous amino acid-DNA interactions generated from this longer loop region (Figure 2B).129

13 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The two RRMs are individually capable of binding to relatively short poly-UG RNA sequences. RRM1 binds 6 UG repeats (Kd = 65.2 nM) while RRM2 binds to 3 UG repeats (Kd = 379 nM), but both RRM domains are required for high affinity, synergistic binding to sequences with greater than 6 UG repeats (Kd = 14.2 nM).30,80 Indeed, the RNA binding targets of TDP-43 are very numerous, and not all possess short UG repeats. The 3’UTR sequence through which TDP43 modulates autoregulation contains 34 nucleotides and some targets of TDP-43 can extend up to 100 nucleotides in length.133,134 Recent NMR studies have produced a structure of tandem RRM domains interacting with a single RNA strand with the sequence 5′-GUGUGAAUGAAU3′, termed AUG12, revealing the role of both RRMs in binding to this target. The tandem domains reside side-by-side upon RNA binding and use both of their hydrophobic β-sheet regions and its loops to generate an extended groove to accommodate the RNA molecule (Figure 2B). However, unlike usual tandem RRM domains in which RNA binds to the grove in a 3’-to-5’ direction from RRM1 to RRM2, the TDP-43 tandem RRMs are arranged in reverse. Subsequently, the linker between the two RRMs that canonically spans only 2 β-strands now spans across a larger area, across 4 β-strands, which allows for this linker region to participate in more extensive interactions with RNA targets as well as the RRMs themselves and potentially other regions of the protein such as the N-terminus (Figure 2B).131 The structural study also reveals that a degenerate consensus sequence of (5′-GNGUGNNUGN-3′) is recognized by the tandem RRMs, unlike other RRMs that require a continuous stretch of 6-9 nucleotides for high affinity binding.132 It is possible that this extended inter-RRM linker region provides the basis for TDP-43’s ability to participate in a large number of RNA processing functions due to its ability to recognize a large number of specific sequences as well as potentially interacting with other RNA binding molecules.131

14 ACS Paragon Plus Environment

Page 14 of 62

Page 15 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

Role of RRM domains in aggregation The role of RRM domains are most often associated with the native RNA processing functions of the protein. In vitro studies have shown, however, that binding of oligonucleotide targets such as 12 TG repeat DNA sequences to TDP-43 through its RRMs prevents TDP-43 aggregation, suggesting some involvement of RRMs in aggregation, directly or indirectly.42,43 Only two ALSassociated mutants have been identified in the RRM domains. The recently identified P112H mutation resides on the Loop1 region of RRM1 between β1 and α1, which may affect RNA binding interactions (Figure 1C). However due to the novelty of the study, the effect of this mutation on TDP-43 structure or function is not well-characterized.122 The other mutant caused by the mutation D169G, located at a short loop region between β4 and β5 of RRM1, shares the same overall structure as the wild type protein and actually has slightly higher binding affinity to oligonucleotide targets, suggesting that it is unlikely to disrupt normal binding functions. Furthermore, this mutant increased the thermal stability of RRM1 due to increased hydrophobic interactions from the D to G mutation. Interestingly, this mutant is more susceptible to caspase 3 cleavage between D208 and V209, which effectively separates α1 from β2 in RRM2, and potentially exposes one side of the β-sheet within the RRM to aberrant protein-protein or proteinDNA interactions, while causing cytoplasmic mislocalization though loss of the nuclear localization signal (Figure 2B). The effect of this cleavage reaction may contribute to aggregation or disruption to native protein function.130 The C-terminal region has a dynamic structure Perhaps the most rigorously studied region of the protein is the CTD (274-414). Since nearly all ALS-associated mutations are found in this region, it has been implicated as an important contributor to pathogenesis. 35 kDa and 25 kDa C-terminal fragments of TDP-43 can be

15 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

generated by caspase 3 cleavage or alternative splicing and are found in pathological inclusions. The exact mechanism as to how these fragments are generated in disease is still a subject of debate.70,71,135,136 This low-complexity region is glycine-rich and resembles sequences of yeast prions.137. The CTD appears to adopt transient and dynamic secondary structures ranging from αhelix to β-strand conformations. The structural study of the CTD is difficult due to its flexible nature, but current studies have focused primarily on a segment of the CTD approximately between residues 318 to 369. This region is considered to be the amyloidogenic core of the protein, as it contains the QN-rich segment at residues 331-369 that is capable of forming amyloid-like β-sheet structures implicated in aggregation, although TDP-43 aggregates formed in vitro and found in patients do not stain positively with amyloid-specific dyes such as Congo Red and Thioflavin S.58,59,64,138–140 Expression of 12 tandem repeats (12×QN) within this region is sufficient to induce the formation of phosphorylated and ubiquitinated inclusions in a cell culture model.127 The residues 321-366 have also been implicated in protein-protein interactions and specifically binding to heterogeneous nuclear ribonucleoprotein (hnRNP) A2/B1 and several other members of the hnRNP family.141,142 Structurally, the amyloidogenic core can be largely divided into two segments, one approximately between residues 320-343, which is capable of forming a transient helix-turn-helix structure and the remaining residues 341-366 which are predicted to form two antiparallel β-sheet structures in molecular dynamics simulations.138,140,143,144 Interestingly, the α-helical region can also undergo α-to-β secondary structure transitions as measured by CD spectroscopy.138 Recent studies have implied that the key α-helix formed cooperatively from residues 321-330 is required for in the formation of RNP granules such as SGs through LLPS, which may be a key mechanism of how TDP-43 performs its native and pathological functions.143,145

16 ACS Paragon Plus Environment

Page 16 of 62

Page 17 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

THE C-TERMINAL DOMAIN: PRIONS, DROPLETS AND AGGREGATION The C-terminal domain resembles yeast prions The most studied region of TDP-43 is its CTD due to its direct involvement in aggregation and pathology. This region can be considered a prion-like domain (PrLD) due to its sequence similarity to yeast prions. Recent findings suggest that this PrLD is not entirely disordered, but can fold dynamically into α-helices or β-strands, which may govern both native protein function in reaction to environmental stress as well as pathological aggregation. The PrLD is not a unique property of TDP-43, and may trace its evolutionary history to early eukaryotes. Although identified as pathological, infectious agents in prion disease in humans, prions in yeast play a major role in yeast metabolism and may confer selective advantages.7,146 Classically, yeast protein Sup35 can misfold into its prion state Ψ+ via its N-terminal domains into the typical amyloid structures composed of cross β-sheets.146,147 Sup35 is a translation termination factor in yeast, but upon folding into its prion conformation, it loses this function and leads to readthrough of nonsense mutations. In the laboratory it was demonstrated that in yeast strains harboring a premature stop codon in their ADE1 gene, cells without the capacity to form Sup35 prion state (Ψ- strain) become auxotrophic for adenine, whereas Ψ+ cells can grow on media lacking adenine. The Ψ+ state can be propagated through template-directed misfolding during mitosis, and amyloid aggregates in the diploid cells are segregated in the four spores during meiosis, leading to non-Mendelian propagation of the Ψ+ state to yeast progeny and retention of this selective advantage in future generations. Another yeast prion Mot3 is a transcription factor that regulates mating, carbon metabolism and stress response under its native state, but the prion state [MOT3+] allows for facultative multicellular growth phenotypes.148 These examples show 17 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

that under certain environmental conditions, the ability to form prions can confer a survival advantage. However, these prion phenotypes are not without disadvantages, since prion formation triggers the increased activity of HSP40/70 in yeast and Ψ+ strains have reduced growth rates compared to their Ψ- counterparts, implying that the presence of prions induces a certain degree of cellular stress.149 Yeast prions appear to be bet-hedging mechanisms that allow adaptive response to environmental stress albeit with the risk of gaining pathology.77,148,150,151 Human proteins containing PrLD form membraneless organelles In addition to TDP-43, PrLD are also found in other human proteins. 70% of human proteins predicted to contain PrLDs by the PLAAC (Prion-Like Amino Acid Composition) search algorithm have molecular functions related to RNA/DNA binding, transcription factor activity or mRNA processing.77 Several of these proteins are implicated in human neurodegenerative disease, such as ataxin 1 and ataxin 2 in spinocerebellar ataxias, and more significantly, hnRNPA1, hnRNPA2/B1, TDP-43 and the RNA binding protein FUS (FUS/TLS; Fused in Sarcoma/Translocated in Sarcoma), which are proteins whose mutations are known to cause ALS/FTD.116,120,152–154 These proteins all share the common feature of having a disordered PrLD and RRMsto mediate RNA binding. In the case of TDP-43, regions within the PrLD are often considered the amyloidogenic core of the protein, responsible for TDP-43 aggregation 64,138,139. Despite the association of these PrLD with disease, there would be no selective pressure to retain these domains if they only confer detrimental effects of protein misfolding in the form of neurodegenerative diseases, yet regions of the PrLD such as the α-helical segment of TDP-43 PrLD is well-conserved in vertebrates.143,155 Thus, it is likely that in humans, the PrLD of proteins have functions that may confer selective advantages with risk of pathology, similar to yeast prions.

18 ACS Paragon Plus Environment

Page 18 of 62

Page 19 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

One possible functional advantage of PrLDs is their role in the formation of membraneless organelles or RNP granules. The lack of a lipid-rich barrier to enclose its constituents is advantageous because it allows environmental changes to rapidly alter the internal equilibrium of the organelle.156 The physical properties of RNP granules were initially studied in germline P granules in C. elegans embryos, where P granules showed classic liquid droplet properties such as spherical morphology, fusion, dissolution, and concentration-dependent condensation, strongly implicating LLPS as their mechanism of formation.157 Many other RNP granules have since been reported such as processing bodies (P-bodies) which are sites of mRNA decay, and SGs which are assemblies of translationally stalled ribosomal subunits and its associated mRNAs that form during cellular stress.158,159 Proteins such as hnRNPA1 and TDP-43 are known to be recruited to SGs, and TDP-43 also modulates SG formation and dynamics.160,161 Membraneless organelles such as SGs allow for transient and reversible aggregation of unneeded transcripts and allows for cell survival under stress conditions.158 Recently, RNA binding proteins containing PrLD, such as FUS, hnRNPA1 and TDP-43, have been reported to undergo LLPS through their PrLD, which is thought to be the underlying mechanism of the formation RNP granules or bodies.143,145,155,160,162–165 The structural changes that occur within these proteins during LLPS, however, are not well understood and remain an active area of research. While a NMR study of FUS droplets suggests that the PrLD remains entirely disordered, a study using mass spectrometry-based footprinting of the PrLD in hnRNPA2 droplets asserts that it adopts a structure rich in cross-β sheets.163,166 It also remains uncertain whether liquid droplets formed in vitro are reflective of the phase architecture of membraneless organelles. TDP-43 undergoes phase transitions through its PrLD

19 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The component of TDP-43 responsible for phase separation into liquid droplets, emerged very recently. Unlike FUS, which appears to have no apparent structure at its PrLD in the droplet state, NMR studies on a TDP-43 PrLD construct consisting of residues 267-414 revealed that it can assemble into liquid droplets through a cooperatively folded, partially populated α-helix, initiated by the addition of 150 mM NaCl or addition of yeast RNA extract.143 Although NMR spectroscopy of the entire CTD indicates that the region is almost entirely disordered, this αhelix formed by residues 321-330 appears to be populated 50% of the time, and interacts with helices from other TDP-43 CTD molecules during liquid droplet formation.143 This apparent RRM-independent RNA interaction may be mediated through a RRG motif on the C-terminus.167 ALS/FTD-associated mutations A321G, Q331K and M337V disrupted phase separation and encouraged conversion to aggregates.143 Although residues Q331 and M337 do not reside within the transiently formed α-helix, structural studies indicate that these residues belong to the helixloop-helix region formed by resides 319-341, a region perfectly conserved among vertebrates and rich in aliphatic residues.168 Additionally, in a cell culture system where the full length TDP43 construct was modified by replacing its RRMs with a GFP reporter, the expressed construct formed nuclear droplets with “bubbles” containing nuclear millieu.145 In agreement with the NMR studies, ALS-causing mutation M337V altered the dynamics of these droplets, whereas mutations N345K and A382T outside the α-helical region did not have as significant of an effect.145 This suggests that the α-helical segment of the CTD and its inter-molecular interactions are critical for the formation of liquid droplets. It is unclear how the N-terminus biophysically affects TDP-43 droplet formation, as LLPS through the C-terminal critical residues have occurred with or without an N-terminal component.143,145 It is possible that the environment within a test tube allows for much higher than physiological achievable concentrations of TDP-

20 ACS Paragon Plus Environment

Page 20 of 62

Page 21 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

43, thus the N-terminus that would normally be required for oligomerization and consolidating TDP-43 would no longer be necessary. In addition to liquid droplets, FUS, TDP-43 and hnRNPA1/A2 can also form amyloid-like folds rich in β-structure.140,144,154,169 TDP-43 is predicted to form β-rich structures through its QN-rich region consisting of residues 341-366, while hnRNPA1 is intrinsically prone to forming irreversible amyloid fibrils.140,144,154 FUS, hnRNPA2, and TDP-43 PrLD are also capable of phase separating into a gel phase known as “hydrogels”.163,165,166,169,170 Unlike the liquid droplet phase, the structural basis of hydrogels appears to be distinctly amyloid-like.166,169 Electron microscopy and X-ray diffraction of FUS and hnRNPA2 hydrogels reveal that they are composed of cross-β, amyloid-like folds. But unlike typical, irreversible amyloid, these hydrogels are readily solubilized by SDS or mild heating.169 The relationship between hydrogels and liquid droplets is not well defined, but recent studies suggest that unlike the present model of SGs where they behave as purely liquid compartments, SGs may contain “cores” of denser material held together by strong interactions within the PrLD.171 These core particles can be purified by conventional centrifugation. They are surrounded by a liquid shell in SGs that allows for the free exchange of materials with other liquid compartments, while exchange of material between the core and the shell is an ATP-dependent process that involves ribonucleoprotein remodeling mechanisms.171 It remains uncertain whether these cores are formed first, followed by recruitment of the outer shell, or whether the SG is formed, followed by condensation of the liquid phase into more stable core structure. We speculate that these core structures may be held together by β-rich, amyloid-like interactions, similar to those found in hydrogels, which PrLDs are capable of forming. It is however uncertain whether TDP-43 forms β-rich core structures in SGs as they were not found in the centrifugation-based purification procedure, but it is becoming

21 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

clear that like other proteins with PrLDs, TDP-43 is capable of forming both liquid droplets mediated by α-helical interactions and amyloid-like hydrogel structures through β-rich folds. While the liquid states appear to be a part of physiological SG function, the conversion of these structures into irreversible aggregates may be associated with pathology. LLPS as drivers for aggregation The formation and maintenance of liquid droplets appears to be a delicate process and a number of factors can cause their transition into aggregates. In fact, liquid droplets tend to become less dynamic and less reversible over time and appear to have an intrinsic propensity to aggregate.51,52 For instance, constructs containing the PrLD of several RNA binding proteins including FUS and hnRNPA1 all produce liquid droplets in vitro under low salt conditions. The PrLD of different proteins can be recruited into the same droplet, forming a heterogeneous mixture, similar to the environment within SGs. However, over timespans of 24 hours, these droplets lose their liquid-like properties and change from dynamic spheres into more stable, irregular or filamentous structures.172 ThT-positive fibrils eventually form on the surface of full length hnRNPA1 liquid droplets after 24 hours and time-dependent “maturation” of SGs that lead to stable and β-rich SG core structures or amyloid-like fibrillization are also observed.160,171 In the case of TDP-43, droplets formed by the PrLD construct only remain stable in the timescale of hours before conversion to aggregates, suggesting that these droplets are transient structures that inevitably enter an irreversible state over time unless otherwise maintained.143 This property of TDP-43 liquid droplets may be attributed to transient nature of the α-helix segment of the PrLD, which readily undergoes α-β transitions, and ALS/FTD mutations that disrupt LLPS encouraged this conversion process.138,143

22 ACS Paragon Plus Environment

Page 22 of 62

Page 23 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

In SGs, the β-rich core structures may be a means of liquid droplet maintenance to sequester proteins entering the aggregating phase and convert them to the droplet state through active, ATP-dependent remodeling mechanisms. SGs normally only persist for durations of hours, and it is possible that prolonged or repeated environmental stresses can overwhelm the remodeling mechanisms and cause irreversible aggregates to form.52,173 The conversion of cellular liquid droplets such as SGs into aggregates is a plausible pathway through which TDP-43 can form pathological inclusions. It follows that any factors that enhances TDP-43’s propensity to enter the aggregate state such as disruption of the α-helix interactions, increase in the propensity to form β-folds, or alterations to RNA content of SGs would accelerate the process of dropletaggregate conversion. Notably, a significant number of ALS/FTD-associated mutations in FUS and hnRNPA1, as well as nearly all disease-associated mutations in TDP-43 occur in their PrLD, and these mutations cause defects in droplet formation leading to decreased reversibility and higher propensity to convert into fibrils.154,160,163,165,174 In FUS, ALS/FTD-associated mutations reduces the speed at which FUS traverses through the liquid droplet and causes formation of starburst-shaped aggregates after in vitro aging.164 In TDP-43, ALS/FTD mutations that occur at the α-helix segment cause disruptions to the intermolecular interactions of the PrLD transient αhelices and altered droplet formation and reversibility.143 Although the effect of mutations outside the α-helix segment on the properties of liquid droplets remains to be tested, we speculate that these mutations may affect LLPS formation through altered protein-protein interactions. Under physiological conditions, membraneless organelles are not purely protein droplets, but contain a collection of different proteins and their associated RNA targets, and the biophysical properties of these droplets such as viscosity and droplet fusion rates, can be altered depending on the identity of the RNA to which the proteins are bound.175 Thus mutations on the

23 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

PrLD that can potentially cause aberrant recruitment of RNA or RNA/protein complexes can contribute to aggregate formation in this manner. The specific occurrence of ALS/FTD mutations on the PrLD suggests that the disease pathology is intimately linked to the formation and aggregation of SGs. Spreading and propagation of ALS/FTD An unanswered question of ALS/FTD is the mechanism of cell to cell spread of the disease. It is tempting to attribute the spreading agent to the aggregation prone, β-rich fold of the TDP-43 PrLD, as prions spread by template-directed misfolding in prion disease. In vitro aggregates of recombinant TDP-43 can seed the aggregation of endogenous TDP-43 when transduced into HEK293T cells and multiple TDP-43 CTD fragments containing residues 287-322 are capable of forming amyloid that possesses the same seeding properties as classic amyloid fibrils.139,176,177 Moreover, transduction of insoluble fractions from ALS/FTD patient brain lysates into SHSY5Y neuronal cells can lead to the formation of phosphorylated and ubiquitinated aggregates.67 In order for cell-to-cell transmission to occur, the aggregation-prone form of TDP-43 must be expelled from the affected neuron, whether through endosomal pathways or cell death, and cross the cell membrane into the extracellular milieu and be taken up by a subsequent neuron. One proposed mechanism of membrane entry is through the microvesicle/exosome pathway through axonal terminals, while others have proposed that aggregates can rupture unstructured macropinosomes through “membrane ruffling” to gain cell entry.178,179 Recent findings suggest that segments containing the α-helical region of the PrLD (311-343) can interact with DMPC/DHPC bicelles, suggesting the possibility of membrane disruption through membranehelix interactions.139,180 The formation of α-helices may also play a more direct role in cell-tocell spreading, since α-helices are also known to form large supramolecular structures through

24 ACS Paragon Plus Environment

Page 24 of 62

Page 25 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

helix-helix interactions between varying numbers of helices such as 3-helix micelles which are used as nanocarrier polymers that can cross a multitude of biological barriers for the delivery of its contents.181 Further studies to identify the mechanism and agent of cell-to-cell spread can open new avenues of research for the treatment of ALS/FTD by potentially blocking the entry of these agents from the affected neuron to neighboring healthy neurons, or sequester these agents using conformation-specific antibodies before they can gain cellular entry.

CONCLUSIONS Recent structural findings in the TDP-43 CTD have produced unprecedented insights into the molecular mechanism of TDP-43 function and aggregation. In particular, the observation of TDP-43 forming liquid droplets through its PrLD has provided a strong line of evidence for the intimate link between SG formation and pathological aggregation. The molecular mechanisms of pathogenesis are starting to come into focus (summarized in Figure 3). The PrLD containing the amyloidogeneic core plays a central role in TDP-43 phase changes and there is a clear correlation between secondary structure of the PrLD and the propensity to form functional versus pathological states. The conversion between these conformations may be a key step in pathogenesis. TDP-43 PrLD appears to wobble at a cusp of the protein folding energy landscape, where any number of factors can tip the balance and trigger the conversion from its native function to its pathological one. This structural conversion could be template-directed in nature, and its occurrence in soluble cytosolic TDP-43 and subsequent formation of pathological aggregates cannot be ruled out. However, as a more exquisite alternative, the conversion may occur within liquid droplets such as SGs, where the increased protein density can accelerate the propagation of β-rich folds. SGs have an intrinsic propensity to aggregate, and mutations that render TDP-43 more prone to adopt a β-fold at the PrLD, that alter concentration through

25 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

cytoplasmic mislocalization, or that can alter protein-protein interactions and in turn alter RNA composition, can all affect SG assembly and dynamics. Additionally, environmental factors such as persistent or chronic stress due to exposure to extreme temperatures, toxins or physical harm, can also accelerate this intrinsic property of SGs. The clearance of these β-rich structures may eventually become stagnant due to the inevitable age-dependent decline of the protein quality control mechanisms, leading to accumulated aggregates and neurotoxicity.

26 ACS Paragon Plus Environment

Page 26 of 62

Page 27 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

ACKNOWLEDGEMENTS This study was supported in whole or in parts by the Canadian Consortium of Neurodegeneration and Aging (CCNA), the Canadian Institute of Health Research (CIHR), the ALS Society of Canada (ALS Canada) and the Alzheimer Society of Canada (ASC). The authors would like to thank Kevin C. Hadley for helpful discussions and critically reading the manuscript.

27 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

FIGURE LEGENDS Figure 1: Domain arrangement and secondary structures of TDP-43. A) Schematic representation of the TDP-43 domains. Domain boundaries are numbered according to full length TDP-43. TDP-43 consists of a N-terminal domain (brown) which contains the nuclear localization signal (grey), tandem RNA recognition motifs (dark yellow) containing predicted nuclear export signal (grey) and the C-terminal domain (blue). Nearly all known ALS/FTD associated mutation occur in the CTD (green box).37,82–123 B) Secondary structure of the NTD. Residues 1-77 contain a ubiquitin-like fold consisting of six β-strands (orange) and one α-helix (green). The nuclear localization sequence is underlined. C) Secondary structure of the tandem RRM domains of TDP-43. Both RRMs share similar secondary structure consisting of five βstrands (dark yellow) and helices α1 (light blue) and α2 (dark blue). D) Secondary structure of Cterminal domain. Only sequences containing observed or simulated secondary structures are shown. These include the helix-turn-helix motif (dark blue) and predicted β-strands (underlined).

Figure 2: Combined molecular structure of TDP-43 bound to a single strand of AUG12 RNA. Separate PDB files were joined together for a conceptual representation of the TDP-43 molecule. Dotted lines represent gaps in the sequence containing unknown structure. The speculative position of the N-terminal domain relative to the tandem RRM domains is based on small angle X-ray scattering data.126 NLS and NES sequences are shown in purple boxes. The color scheme of the secondary structures is consistent with that of Figure 1. A) The N-terminal ubiquitin-like fold consists of residues 1-77, derived from the NMR structure 2N4P.125 β-strands 1, 2, 3, and 6 form sheet 1 (dark orange), while strands β4 and β5 form sheet 2 (light orange). The α1 helix is colored in green. B) Structure of tandem TDP-43 RRM domains bound to

28 ACS Paragon Plus Environment

Page 28 of 62

Page 29 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

AUG12 (pink line drawing) generated from the NMR structure 4BS2.131 Color scheme of the secondary structures in each domain is consistent, with β-sheet consisting of strands 1-5 (yellow) sandwiched between helices α1 (cyan) and α2 (light blue). The (’) notation denotes the matching secondary structures in RRM2. Key loop regions loop 1 (pink), loop 3 (orange) and the loop joining the two RRMs (green) are indicated. These loop regions combined with the extensive βsheet surface of both RRMs creates the binding surface for RNA target AUG12 (5′GUGUGAAUGAAU-3′). C) The secondary structure of TPD-43 amyloidogenic region derived from NMR structure 2N3X.168 This region contains extensive loops of unstructured regions except for residues 320-343 which consists of a helix-turn-helix motif (dark blue).

Figure 3: Graphical representation of TDP-43 aggregation model. TDP-43 is represented by a round blue circle (N-terminus) and two orange squares (tandem RRMs) bound to RNA (red half-ladders), attached to a C-terminal helix (blue lines) in its native state as a dimer. In the stress granule, TDP-43 C-terminal helix interact to form liquid droplets containing other proteins (green circles) and RNA (green half-ladders). The structural transition of the prion-like domain of TDP-43 to β-rich folds is represented with pink antiparallel arrows. The fibril-like arrangement of the antiparallel arrows reflect amyloid-like folds associated with stable stress granule cores or irreversible aggregates in the cytoplasm. Post-translational modifications to these aggregates are represented by yellow circles labeled P for phosphorylation, or purple circles labeled U to represent ubiquitination. The pink-blue gradient on the left represents the conversion of α-helix to β-rich folds in the prion-like domain, reflecting the transition from native to pathological TDP-43 states. The schematic presents both stress granule dependent and independent TDP-43 aggregation pathways.

29 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

REFERENCES (1) Swamy, M. S., and Abraham, E. C. (1987) Lens protein composition, glycation and high molecular weight aggregation in aging rats. Investig. Ophthalmol. Vis. Sci. 28, 1693–1701. (2) Dobson, C. M. (2002) Getting out of shape. Nature 418, 729–730. (3) Dobson, C. M. (2003) Protein folding and misfolding. Nature 426, 884–890. (4) Galant, N. J., Bugyei-Twum, A., Rakhit, R., Walsh, P., Sharpe, S., Arslan, P. E., Westermark, P., Higaki, J. N., Torres, R., Tapia, J., and Chakrabartty, A. (2016) Substoichiometric inhibition of transthyretin misfolding by immune-targeting sparsely populated misfolding intermediates: a potential diagnostic and therapeutic for TTR amyloidoses. Sci. Rep. 6, 25080. (5) Ishimaru, D., Ano Bom, A. P. D., Lima, L. M. T. R., Quesado, P. a, Oyama, M. F. C., de Moura Gallo, C. V, Cordeiro, Y., and Silva, J. L. (2009) Cognate DNA stabilizes the tumor suppressor p53 and prevents misfolding and aggregation. Biochemistry 48, 6126–6135. (6) Soragni, A., Janzen, D. M., Johnson, L. M., Lindgren, A. G., Thai-Quynh Nguyen, A., Tiourin, E., Soriaga, A. B., Lu, J., Jiang, L., Faull, K. F., Pellegrini, M., Memarzadeh, S., and Eisenberg, D. S. (2016) A Designed Inhibitor of p53 Aggregation Rescues p53 Tumor Suppression in Ovarian Carcinomas. Cancer Cell 29, 90–103. (7) Bolton, D. C., McKinley, M. P., and Prusiner, S. B. (1982) Identification of a protein that purifies with the scrapie prion. Science 218, 1309–1311. (8) Glenner, G. G., and Wong, C. W. (1984) Alzheimer’s disease: Initial report of the purification and characterization of a novel cerebrovascular amyloid protein. Biochem. Biophys. Res. Commun. 120, 885–890.

30 ACS Paragon Plus Environment

Page 30 of 62

Page 31 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

(9) Brooks, B. R., Miller, R. G., Swash, M., and Munsat, T. L. (2000) El Escorial revisited: revised criteria for the diagnosis of amyotrophic lateral sclerosis. Amyotroph. Lateral Scler. Other Motor Neuron Disord. 1, 293–299. (10) Tandan, R., and Bradley, W. G. (1985) Amyotrophic lateral sclerosis: Part 1. Clinical features, pathology, and ethical issues in management. Ann. Neurol. 18, 271–280. (11) Mehta, P., Kaye, W., Bryan, L., Larson, T., Copeland, T., Wu, J., Muravov, O., and Horton, K. (2016) Prevalence of Amyotrophic Lateral Sclerosis - United States, 2012-2013. MMWR. Surveill. Summ. 65, 1–12. (12) Majoor-Krakauer, D., Willems, P. J., and Hofman, a. (2003) Genetic epidemiology of amyotrophic lateral sclerosis. Clin. Genet. 63, 83–101. (13) Mackenzie, I. R. A., and Neumann, M. (2016) Molecular neuropathology of frontotemporal dementia: insights into disease mechanisms from postmortem studies. J. Neurochem. 138, 54–70. (14) Neary, D., Snowden, J. S., Gustafson, L., Passant, U., Stuss, D., Black, S., Freedman, M., Kertesz, A., Robert, P. H., Albert, M., Boone, K., Miller, B. L., Cummings, J., and Benson, D. F. (1998) Frontotemporal lobar degeneration: a consensus on clinical diagnostic criteria. Neurology 51, 1546–1554. (15) Cruts, M., Gijselinck, I., Van Langenhove, T., van der Zee, J., and Van Broeckhoven, C. (2013) Current insights into the C9orf72 repeat expansion diseases of the FTLD/ALS spectrum. Trends Neurosci. 36, 450–459. (16) Hutton, M., Lendon, C. L., Rizzu, P., Baker, M., Froelich, S., Houlden, H., PickeringBrown, S., Chakraverty, S., Isaacs, a, Grover, a, Hackett, J., Adamson, J., Lincoln, S., Dickson,

31 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

D., Davies, P., Petersen, R. C., Stevens, M., de Graaff, E., Wauters, E., van Baren, J., Hillebrand, M., Joosse, M., Kwon, J. M., Nowotny, P., Che, L. K., Norton, J., Morris, J. C., Reed, L. a, Trojanowski, J., Basun, H., Lannfelt, L., Neystat, M., Fahn, S., Dark, F., Tannenberg, T., Dodd, P. R., Hayward, N., Kwok, J. B., Schofield, P. R., Andreadis, a, Snowden, J., Craufurd, D., Neary, D., Owen, F., Oostra, B. a, Hardy, J., Goate, a, van Swieten, J., Mann, D., Lynch, T., and Heutink, P. (1998) Association of missense and 5’-splice-site mutations in tau with the inherited dementia FTDP-17. Nature 393, 702–705. (17) Cruts, M., Gijselinck, I., van der Zee, J., Engelborghs, S., Wils, H., Pirici, D., Rademakers, R., Vandenberghe, R., Dermaut, B., Martin, J. J., van Duijn, C., Peeters, K., Sciot, R., Santens, P., De Pooter, T., Mattheijssens, M., Van den Broeck, M., Cuijt, I., Vennekens, K., De Deyn, P. P., Kumar-Singh, S., and Van Broeckhoven, C. (2006) Null mutations in progranulin cause ubiquitin-positive frontotemporal dementia linked to chromosome 17q21. Nature 442, 920–924. (18) Baker, M., Mackenzie, I. R., Pickering-Brown, S. M., Gass, J., Rademakers, R., Lindholm, C., Snowden, J., Adamson, J., Sadovnick, a D., Rollinson, S., Cannon, A., Dwosh, E., Neary, D., Melquist, S., Richardson, A., Dickson, D., Berger, Z., Eriksen, J., Robinson, T., Zehr, C., Dickey, C. a, Crook, R., McGowan, E., Mann, D., Boeve, B., Feldman, H., and Hutton, M. (2006) Mutations in progranulin cause tau-negative frontotemporal dementia linked to chromosome 17. Nature 442, 916–919. (19) DeJesus-Hernandez, M., Mackenzie, I. R. R., Boeve, B. F. F., Boxer, A. L. L., Baker, M., Rutherford, N. J. J., Nicholson, A. M. M., Finch, N. A. a, Flynn, H., Adamson, J., Kouri, N., Wojtas, A., Sengdy, P., Hsiung, G.-Y. R. R., Karydas, A., Seeley, W. W. W., Josephs, K. A. a, Coppola, G., Geschwind, D. H. H., Wszolek, Z. K. K., Feldman, H., Knopman, D. S. S., Petersen, R. C. C., Miller, B. L. L., Dickson, D. W. W., Boylan, K. B. B., Graff-Radford, N. R. 32 ACS Paragon Plus Environment

Page 32 of 62

Page 33 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

R., and Rademakers, R. (2011) Expanded GGGGCC Hexanucleotide Repeat in Noncoding Region of C9ORF72 Causes Chromosome 9p-Linked FTD and ALS. Neuron 72, 245–256. (20) Renton, A. E., Majounie, E., Waite, A., Simón-Sánchez, J., Rollinson, S., Gibbs, J. R., Schymick, J. C., Laaksovirta, H., van Swieten, J. C., Myllykangas, L., Kalimo, H., Paetau, A., Abramzon, Y., Remes, A. M., Kaganovich, A., Scholz, S. W., Duckworth, J., Ding, J., Harmer, D. W., Hernandez, D. G., Johnson, J. O., Mok, K., Ryten, M., Trabzuni, D., Guerreiro, R. J., Orrell, R. W., Neal, J., Murray, A., Pearson, J., Jansen, I. E., Sondervan, D., Seelaar, H., Blake, D., Young, K., Halliwell, N., Callister, J. B., Toulson, G., Richardson, A., Gerhard, A., Snowden, J., Mann, D., Neary, D., Nalls, M. A., Peuralinna, T., Jansson, L., Isoviita, V.-M. M., Kaivorinne, A.-L. L., Hölttä-Vuori, M., Ikonen, E., Sulkava, R., Benatar, M., Wuu, J., Chiò, A., Restagno, G., Borghero, G., Sabatelli, M., Heckerman, D., Rogaeva, E., Zinman, L., Rothstein, J. D., Sendtner, M., Drepper, C., Eichler, E. E., Alkan, C., Abdullaev, Z., Pack, S. D., Dutra, A., Pak, E., Hardy, J., Singleton, A., Williams, N. M., Heutink, P., Pickering-Brown, S., Morris, H. R., Tienari, P. J., and Traynor, B. J. (2011) A hexanucleotide repeat expansion in C9ORF72 is the cause of chromosome 9p21-linked ALS-FTD. Neuron 72, 257–268. (21) Lashley, T., Rohrer, J. D., Mead, S., and Revesz, T. (2015) Review: An update on clinical, genetic and pathological aspects of frontotemporal lobar degenerations. Neuropathol. Appl. Neurobiol. 41, 858–881. (22) Weishaupt, J. H., Hyman, T., and Dikic, I. (2016) Common Molecular Pathways in Amyotrophic Lateral Sclerosis and Frontotemporal Dementia. Trends Mol. Med. 22, 769–783. (23) Mackenzie, I. R. A. (2007) The neuropathology of FTD associated With ALS. Alzheimer Dis. Assoc. Disord. 21, S44–S49.

33 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(24) Arai, T., Hasegawa, M., Akiyama, H., Ikeda, K., Nonaka, T., Mori, H., Mann, D., Tsuchiya, K., Yoshida, M., Hashizume, Y., and Oda, T. (2006) TDP-43 is a component of ubiquitinpositive tau-negative inclusions in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Biochem. Biophys. Res. Commun. 351, 602–611. (25) Neumann, M., Sampathu, D. M., Kwong, L. K., Truax, A. C., Micsenyi, M. C., Chou, T. T., Bruce, J., Schuck, T., Grossman, M., Clark, C. M., McCluskey, L. F., Miller, B. L., Masliah, E., Mackenzie, I. R., Feldman, H., Feiden, W., Kretzschmar, H. a, Trojanowski, J. Q., and Lee, V. M.-Y. (2006) Ubiquitinated TDP-43 in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Science 314, 130–133. (26) Ling, S. C., Polymenidou, M., and Cleveland, D. W. (2013) Converging mechanisms in als and FTD: Disrupted RNA and protein homeostasis. Neuron 79, 416–438. (27) Ou, S. H., Wu, F., Harrich, D., García-Martínez, L. F., and Gaynor, R. B. (1995) Cloning and characterization of a novel cellular protein, TDP-43, that binds to human immunodeficiency virus type 1 TAR DNA sequence motifs. J. Virol. 69, 3584–3596. (28) Buratti, E., Dörk, T., Zuccato, E., Pagani, F., Romano, M., and Baralle, F. E. (2001) Nuclear factor TDP-43 and SR proteins promote in vitro and in vivo CFTR exon 9 skipping. EMBO J. 20, 1774–1784. (29) Acharya, K. K., Govind, C. K., Shore, A. N., Stoler, M. H., and Reddi, P. P. (2006) cisrequirement for the maintenance of round spermatid-specific transcription. Dev. Biol. 295, 781– 790. (30) Buratti, E., Brindisi, A., Pagani, F., and Baralle, F. E. (2004) Nuclear factor TDP-43 binds to the polymorphic TG repeats in CFTR intron 8 and causes skipping of exon 9: a functional link 34 ACS Paragon Plus Environment

Page 34 of 62

Page 35 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

with disease penetrance. Am. J. Hum. Genet. 74, 1322–1325. (31) Mercado, P. A., Ayala, Y. M., Romano, M., Buratti, E., and Baralle, F. E. (2005) Depletion of TDP 43 overrides the need for exonic and intronic splicing enhancers in the human apoA-II gene. Nucleic Acids Res. 33, 6000–6010. (32) Disset, A., Michot, C., Harris, A., Buratti, E., Claustres, M., and Tuffery-Giraud, S. (2005) A T3 allele in the CFTR gene exacerbates exon 9 skipping in vas deferens and epididymal cell lines and is associated with Congenital Bilateral Absence of Vas Deferens (CBAVD). Hum. Mutat. 25, 72–81. (33) Buratti, E., and Baralle, F. E. (2001) Characterization and functional implications of the RNA binding properties of nuclear factor TDP-43, a novel splicing regulator of CFTR exon 9. J. Biol. Chem. 276, 36337–36343. (34) Ayala, Y. M., Zago, P., D’Ambrogio, A., Xu, Y.-F., Petrucelli, L., Buratti, E., and Baralle, F. E. (2008) Structural determinants of the cellular localization and shuttling of TDP-43. J. Cell Sci. 121, 3778–3785. (35) Ratti, A., and Buratti, E. (2016) Physiological Functions and Pathobiology of TDP-43 and FUS/TLS proteins. J. Neurochem. 138, 95–111. (36) Ayala, Y. M., Misteli, T., and Baralle, F. E. (2008) TDP-43 regulates retinoblastoma protein phosphorylation through the repression of cyclin-dependent kinase 6 expression. Proc. Natl. Acad. Sci. U. S. A. 105, 3785–3789. (37) Sreedharan, J., Blair, I. P., Tripathi, V. B., Hu, X., Vance, C., Rogelj, B., Ackerley, S., Durnall, J. C., Williams, K. L., Buratti, E., Baralle, F., de Belleroche, J., Mitchell, J. D., Leigh, P.

35 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

N., Al-Chalabi, A., Miller, C. C., Nicholson, G., and Shaw, C. E. (2008) TDP-43 mutations in familial and sporadic amyotrophic lateral sclerosis. Science 319, 1668–1672. (38) Ayala, Y. M., De Conti, L., Avendaño-Vázquez, S. E., Dhir, A., Romano, M., D’Ambrogio, A., Tollervey, J., Ule, J., Baralle, M., Buratti, E., and Baralle, F. E. (2011) TDP-43 regulates its mRNA levels through a negative feedback loop. EMBO J. 30, 277–288. (39) Budini, M., and Buratti, E. (2011) TDP-43 autoregulation: implications for disease. J. Mol. Neurosci. 45, 473–479. (40) Bembich, S., Herzog, J. S., De Conti, L., Stuani, C., Avendaño-Vázquez, S. E., Buratti, E., Baralle, M., and Baralle, F. E. (2014) Predominance of spliceosomal complex formation over polyadenylation site selection in TDP-43 autoregulation. Nucleic Acids Res. 42, 3362–3371. (41) Pesiridis, G. S., Tripathy, K., Tanik, S., Trojanowski, J. Q., and Lee, V. M.-Y. (2011) A “two-hit” hypothesis for inclusion formation by carboxyl-terminal fragments of TDP-43 protein linked to RNA depletion and impaired microtubule-dependent transport. J. Biol. Chem. 286, 18845–18855. (42) Huang, Y.-C., Lin, K.-F., He, R.-Y., Tu, P.-H., Koubek, J., Hsu, Y.-C., and Huang, J. J.-T. (2013) Inhibition of TDP-43 Aggregation by Nucleic Acid Binding. PLoS One 8, e64002. (43) Sun, Y., Arslan, P. E., Won, A., Yip, C. M., and Chakrabartty, A. (2014) Binding of TDP43 to the 3’UTR of Its Cognate mRNA Enhances Its Solubility. Biochemistry 53, 5885–5894. (44) Wang, I.-F., Chang, H.-Y., Hou, S.-C., Liou, G.-G., Way, T.-D., and James Shen, C.-K. (2012) The self-interaction of native TDP-43 C terminus inhibits its degradation and contributes to early proteinopathies. Nat. Commun. 3, 766.

36 ACS Paragon Plus Environment

Page 36 of 62

Page 37 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

(45) Udan-Johns, M., Bengoechea, R., Bell, S., Shao, J., Diamond, M. I., True, H. L., Weihl, C. C., and Baloh, R. H. (2014) Prion-like nuclear aggregation of TDP-43 during heat shock is regulated by HSP40/70 chaperones. Hum. Mol. Genet. 23, 157–170. (46) Parker, S. J., Meyerowitz, J., James, J. L., Liddell, J. R., Crouch, P. J., Kanninen, K. M., and White, A. R. (2012) Endogenous TDP-43 localized to stress granules can subsequently form protein aggregates. Neurochem. Int. 60, 415–424. (47) Dewey, C. M., Cenik, B., Sephton, C. F., Dries, D. R., Mayer, P., Good, S. K., Johnson, B. A., Herz, J., and Yu, G. (2011) TDP-43 is directed to stress granules by sorbitol, a novel physiological osmotic and oxidative stressor. Mol. Cell. Biol. 31, 1098–1108. (48) Aulas, A., Stabile, S., and Vande Velde, C. (2012) Endogenous TDP-43, but not FUS, contributes to stress granule assembly via G3BP. Mol. Neurodegener. 7, 54. (49) McDonald, K. K., Aulas, A., Destroismaisons, L., Pickles, S., Beleac, E., Camu, W., Rouleau, G. a, and Vande Velde, C. (2011) TAR DNA-binding protein 43 (TDP-43) regulates stress granule dynamics via differential regulation of G3BP and TIA-1. Hum. Mol. Genet. 20, 1400–1410. (50) Wolozin, B. (2012) Regulated protein aggregation: stress granules and neurodegeneration. Mol. Neurodegener. 7, 56. (51) Li, Y. R., King, O. D., Shorter, J., and Gitler, A. D. (2013) Stress granules as crucibles of ALS pathogenesis. J. Cell Biol. 201, 361–372. (52) Bentmann, E., Haass, C., and Dormann, D. (2013) Stress granules in neurodegeneration-lessons learnt from TAR DNA binding protein of 43 kDa and fused in sarcoma. FEBS J. 280,

37 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

4348–4370. (53) Wu, L.-S., Cheng, W.-C., Hou, S.-C., Yan, Y.-T., Jiang, S.-T., and Shen, C.-K. J. (2010) TDP-43, a neuro-pathosignature factor, is essential for early mouse embryogenesis. Genesis 48, 56–62. (54) Sephton, C. F., Good, S. K., Atkin, S., Dewey, C. M., Mayer, P., Herz, J., and Yu, G. (2010) TDP-43 is a developmentally regulated protein essential for early embryonic development. J. Biol. Chem. 285, 6826–6834. (55) Wang, H.-Y., Wang, I.-F., Bose, J., and Shen, C.-K. J. (2004) Structural diversity and functional implications of the eukaryotic TDP gene family. Genomics 83, 130–139. (56) Seyfried, N. T., Gozal, Y. M., Dammer, E. B., Xia, Q., Duong, D. M., Cheng, D., Lah, J. J., Levey, A. I., and Peng, J. (2010) Multiplex SILAC analysis of a cellular TDP-43 proteinopathy model reveals protein inclusions associated with SUMOylation and diverse polyubiquitin chains. Mol. Cell. Proteomics 9, 705–718. (57) Cohen, T. J., Hwang, A. W., Restrepo, C. R., Yuan, C.-X., Trojanowski, J. Q., and Lee, V. M. Y. (2015) An acetylation switch controls TDP-43 function and aggregation propensity. Nat. Commun. 6, 5845. (58) Capitini, C., Conti, S., Perni, M., Guidi, F., Cascella, R., De Poli, A., Penco, A., Relini, A., Cecchi, C., and Chiti, F. (2014) TDP-43 inclusion bodies formed in bacteria are structurally amorphous, non-amyloid and inherently toxic to neuroblastoma cells. PLoS One 9, e86720. (59) Kerman, A., Liu, H.-N., Croul, S., Bilbao, J., Rogaeva, E., Zinman, L., Robertson, J., and Chakrabartty, A. (2010) Amyotrophic lateral sclerosis is a non-amyloid disease in which

38 ACS Paragon Plus Environment

Page 38 of 62

Page 39 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

extensive misfolding of SOD1 is unique to the familial form. Acta Neuropathol. 119, 335–344. (60) Dammer, E. B., Fallini, C., Gozal, Y. M., Duong, D. M., Rossoll, W., Xu, P., Lah, J. J., Levey, A. I., Peng, J., Bassell, G. J., and Seyfried, N. T. (2012) Coaggregation of RNA-binding proteins in a model of TDP-43 proteinopathy with selective RGG motif methylation and a role for RRM1 ubiquitination. PLoS One 7, e38658. (61) Collins, M., Riascos, D., Kovalik, T., An, J., Krupa, K., Krupa, K., Hood, B. L., Conrads, T. P., Renton, A. E., Traynor, B. J., and Bowser, R. (2012) The RNA-binding motif 45 (RBM45) protein accumulates in inclusion bodies in amyotrophic lateral sclerosis (ALS) and frontotemporal lobar degeneration with TDP-43 inclusions (FTLD-TDP) patients. Acta Neuropathol. 124, 717–732. (62) Giordana, M. T., Piccinini, M., Grifoni, S., De Marco, G., Vercellino, M., Magistrello, M., Pellerino, A., Buccinnà, B., Lupino, E., and Rinaudo, M. T. (2010) TDP-43 redistribution is an early event in sporadic amyotrophic lateral sclerosis. Brain Pathol. 20, 351–360. (63) Dormann, D., and Haass, C. (2011) TDP-43 and FUS: a nuclear affair. Trends Neurosci. 34, 339–348. (64) Johnson, B. S., Snead, D., Lee, J. J., McCaffery, J. M., Shorter, J., and Gitler, A. D. (2009) TDP-43 is intrinsically aggregation-prone, and amyotrophic lateral sclerosis-linked mutations accelerate aggregation and increase toxicity. J. Biol. Chem. 284, 20329–20339. (65) Igaz, L. M., Kwong, L. K., Chen-Plotkin, A., Winton, M. J., Unger, T. L., Xu, Y., Neumann, M., Trojanowski, J. Q., and Lee, V. M.-Y. (2009) Expression of TDP-43 C-terminal Fragments in Vitro Recapitulates Pathological Features of TDP-43 Proteinopathies. J. Biol. Chem. 284, 8516–8524. 39 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(66) Zhang, T., Mullane, P. C., Periz, G., and Wang, J. (2011) TDP-43 neurotoxicity and protein aggregation modulated by heat shock factor and insulin/IGF-1 signaling. Hum. Mol. Genet. 20, 1952–1965. (67) Nonaka, T., Masuda-Suzukake, M., Arai, T., Hasegawa, Y., Akatsu, H., Obi, T., Yoshida, M., Murayama, S., Mann, D. M. a, Akiyama, H., and Hasegawa, M. (2013) Prion-like properties of pathological TDP-43 aggregates from diseased brains. Cell Rep. 4, 124–134. (68) Langellotti, S., Romano, V., Romano, G., Klima, R., Feiguin, F., Cragnaz, L., Romano, M., and Baralle, F. E. (2016) A novel fly model of TDP-43 proteinopathies: N-terminus sequences combined with the Q/N domain induce protein functional loss and locomotion defects. Dis. Model. Mech. 659–669. (69) Lee, E. B., Lee, V. M.-Y., and Trojanowski, J. Q. (2012) Gains or losses: molecular mechanisms of TDP43-mediated neurodegeneration. Nat. Rev. Neurosci. 13, 38–50. (70) Nishimoto, Y., Ito, D., Yagi, T., Nihei, Y., Tsunoda, Y., and Suzuki, N. (2010) Characterization of alternative isoforms and inclusion body of the TAR DNA-binding protein43. J. Biol. Chem. 285, 608–619. (71) Xiao, S., Sanelli, T., Chiang, H., Sun, Y., Chakrabartty, A., Keith, J., Rogaeva, E., Zinman, L., and Robertson, J. (2015) Low molecular weight species of TDP-43 generated by abnormal splicing form inclusions in amyotrophic lateral sclerosis and result in motor neuron death. Acta Neuropathol. 130, 49–61. (72) Barmada, S. J., Skibinski, G., Korb, E., Rao, E. J., Wu, J. Y., and Finkbeiner, S. (2010) Cytoplasmic mislocalization of TDP-43 is toxic to neurons and enhanced by a mutation associated with familial amyotrophic lateral sclerosis. J. Neurosci. 30, 639–649. 40 ACS Paragon Plus Environment

Page 40 of 62

Page 41 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

(73) Anderson, P., and Kedersha, N. (2009) RNA granules: post-transcriptional and epigenetic modulators of gene expression. Nat. Rev. Mol. Cell Biol. 10, 430–436. (74) Henao-Mejia, J., and He, J. J. (2009) Sam68 relocalization into stress granules in response to oxidative stress through complexing with TIA-1. Exp. Cell Res. 315, 3381–3395. (75) Gilks, N., Kedersha, N., Ayodele, M., Shen, L., Stoecklin, G., Dember, L. M., and Anderson, P. (2004) Stress granule assembly is mediated by prion-like aggregation of TIA-1. Mol. Biol. Cell 15, 5383–5398. (76) Kedersha, N., Chen, S., Gilks, N., Li, W., Miller, I. J., Stahl, J., and Anderson, P. (2002) Evidence that ternary complex (eIF2-GTP-tRNA(i)(Met))-deficient preinitiation complexes are core constituents of mammalian stress granules. Mol. Biol. Cell 13, 195–210. (77) March, Z. M., King, O. D., and Shorter, J. (2016) Prion-like domains as epigenetic regulators, scaffolds for subcellular organization, and drivers of neurodegenerative disease. Brain Res. 1647, 1–14. (78) la Cour, T., Kiemer, L., Mølgaard, A., Gupta, R., Skriver, K., and Brunak, S. (2004) Analysis and prediction of leucine-rich nuclear export signals. Protein Eng. Des. Sel. 17, 527– 536. (79) Zhang, Y. J., Caulfield, T., Xu, Y. F., Gendron, T. F., Hubbard, J., Stetler, C., Sasaguri, H., Whitelaw, E. C., Cai, S., Lee, W. C., and Petrucelli, L. (2013) The dual functions of the extreme N-terminus of TDP-43 in regulating its biological activity and inclusion formation. Hum. Mol. Genet. 22, 3112–3122. (80) Kuo, P.-H., Doudeva, L. G., Wang, Y.-T., Shen, C.-K. J., and Yuan, H. S. (2009) Structural

41 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

insights into TDP-43 in nucleic-acid binding and domain interactions. Nucleic Acids Res. 37, 1799–1808. (81) Chang, C., Wu, T.-H., Wu, C.-Y., Chiang, M., Toh, E. K.-W., Hsu, Y.-C., Lin, K.-F., Liao, Y., Huang, T., and Huang, J. J.-T. (2012) The N-terminus of TDP-43 promotes its oligomerization and enhances DNA binding affinity. Biochem. Biophys. Res. Commun. 425, 219–224. (82) Kovacs, G. G., Murrell, J. R., Horvath, S., Haraszti, L., Majtenyi, K., Molnar, M. J., Budka, H., Ghetti, B., and Spina, S. (2009) TARDBP variation associated with frontotemporal dementia, supranuclear gaze palsy, and chorea. Mov. Disord. 24, 1843–1847. (83) Kabashi, E., Valdmanis, P. N., Dion, P., Spiegelman, D., McConkey, B. J., Vande Velde, C., Bouchard, J.-P., Lacomblez, L., Pochigaeva, K., Salachas, F., Pradat, P.-F., Camu, W., Meininger, V., Dupre, N., and Rouleau, G. a. (2008) TARDBP mutations in individuals with sporadic and familial amyotrophic lateral sclerosis. Nat. Genet. 40, 572–574. (84) Winton, M. J., Van Deerlin, V. M., Kwong, L. K., Yuan, W., Wood, E. M., Yu, C.-E., Schellenberg, G. D., Rademakers, R., Caselli, R., Karydas, A., Trojanowski, J. Q., Miller, B. L., and Lee, V. M.-Y. (2008) A90V TDP-43 variant results in the aberrant localization of TDP-43 in vitro. FEBS Lett. 582, 2252–2256. (85) Guerreiro, R. J., Schymick, J. C., Crews, C., Singleton, A., Hardy, J., and Traynor, B. J. (2008) TDP-43 is not a common cause of sporadic amyotrophic lateral sclerosis. PLoS One 3, e2450. (86) Kirby, J., Goodall, E. F., Smith, W., Highley, J. R., Masanzu, R., Hartley, J. A., Hibberd, R., Hollinger, H. C., Wharton, S. B., Morrison, K. E., Ince, P. G., McDermott, C. J., and Shaw, P. J. 42 ACS Paragon Plus Environment

Page 42 of 62

Page 43 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

(2010) Broad clinical phenotypes associated with TAR-DNA binding protein (TARDBP) mutations in amyotrophic lateral sclerosis. Neurogenetics 11, 217–225. (87) Corrado, L., Ratti, A., Gellera, C., Buratti, E., Castellotti, B., Carlomagno, Y., Ticozzi, N., Mazzini, L., Testa, L., Taroni, F., Baralle, F. E., Silani, V., and D’Alfonso, S. (2009) High frequency of TARDBP gene mutations in Italian patients with amyotrophic lateral sclerosis. Hum. Mutat. 30, 688–694. (88) Borroni, B., Bonvicini, C., Alberici, A., Buratti, E., Agosti, C., Archetti, S., Papetti, A., Stuani, C., Di Luca, M., Gennarelli, M., and Padovani, A. (2009) Mutation within TARDBP leads to frontotemporal dementia without motor neuron disease. Hum. Mutat. 30, E974–E983. (89) Van Deerlin, V. M., Leverenz, J. B., Bekris, L. M., Bird, T. D., Yuan, W., Elman, L. B., Clay, D., Wood, E. M., Chen-Plotkin, A. S., Martinez-Lage, M., Steinbart, E., McCluskey, L., Grossman, M., Neumann, M., Wu, I.-L., Yang, W.-S., Kalb, R., Galasko, D. R., Montine, T. J., Trojanowski, J. Q., Lee, V. M.-Y., Schellenberg, G. D., and Yu, C.-E. (2008) TARDBP mutations in amyotrophic lateral sclerosis with TDP-43 neuropathology: a genetic and histopathological analysis. Lancet. Neurol. 7, 409–416. (90) Pamphlett, R., Luquin, N., McLean, C., Jew, S. K., and Adams, L. (2009) TDP-43 neuropathology is similar in sporadic amyotrophic lateral sclerosis with or without TDP-43 mutations. Neuropathol. Appl. Neurobiol. 35, 222–225. (91) Luquin, N., Yu, B., Saunderson, R. B., Trent, R. J., and Pamphlett, R. (2009) Genetic variants in the promoter of TARDBP in sporadic amyotrophic lateral sclerosis. Neuromuscul. Disord. 19, 696–700. (92) Del Bo, R., Ghezzi, S., Corti, S., Pandolfo, M., Ranieri, M., Santoro, D., Ghione, I., Prelle, 43 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

A., Orsetti, V., Mancuso, M., Sorarù, G., Briani, C., Angelini, C., Siciliano, G., Bresolin, N., and Comi, G. P. (2009) TARDBP (TDP-43) sequence analysis in patients with familial and sporadic ALS: identification of two novel mutations. Eur. J. Neurol. 16, 727–732. (93) Williams, K. L., Durnall, J. C., Thoeng, A. D., Warraich, S. T., Nicholson, G. A., and Blair, I. P. (2009) A novel TARDBP mutation in an Australian amyotrophic lateral sclerosis kindred. J. Neurol. Neurosurg. Psychiatry 80, 1286–1288. (94) Benajiba, L., Le Ber, I., Camuzat, A., Lacoste, M., Thomas-Anterion, C., Couratier, P., Legallic, S., Salachas, F., Hannequin, D., Decousus, M., Lacomblez, L., Guedj, E., Golfier, V., Camu, W., Dubois, B., Campion, D., Meininger, V., Brice, A., and French Clinical and Genetic Research Network on Frontotemporal Lobar Degeneration/Frontotemporal Lobar Degeneration with Motoneuron Disease. (2009) TARDBP mutations in motoneuron disease with frontotemporal lobar degeneration. Ann. Neurol. 65, 470–473. (95) Gitcho, M. A., Baloh, R. H., Chakraverty, S., Mayo, K., Norton, J. B., Levitch, D., Hatanpaa, K. J., White, C. L., Bigio, E. H., Caselli, R., Baker, M., Al-Lozi, M. T., Morris, J. C., Pestronk, A., Rademakers, R., Goate, A. M., and Cairns, N. J. (2008) TDP-43 A315T mutation in familial motor neuron disease. Ann. Neurol. 63, 535–538. (96) Cairns, N. J., Perrin, R. J., Schmidt, R. E., Gru, A., Green, K. G., Carter, D., TaylorReinwald, L., Morris, J. C., Gitcho, M. A., and Baloh, R. H. (2010) TDP-43 proteinopathy in familial motor neurone disease with TARDBP A315T mutation: a case report. Neuropathol. Appl. Neurobiol. 36, 673–679. (97) Bäumer, D., Parkinson, N., and Talbot, K. (2009) TARDBP in amyotrophic lateral sclerosis: identification of a novel variant but absence of copy number variation. J. Neurol. Neurosurg. 44 ACS Paragon Plus Environment

Page 44 of 62

Page 45 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

Psychiatry 80, 1283–1285. (98) Rutherford, N. J., Zhang, Y.-J., Baker, M., Gass, J. M., Finch, N. a, Xu, Y.-F., Stewart, H., Kelley, B. J., Kuntz, K., Crook, R. J. P., Sreedharan, J., Vance, C., Sorenson, E., Lippa, C., Bigio, E. H., Geschwind, D. H., Knopman, D. S., Mitsumoto, H., Petersen, R. C., Cashman, N. R., Hutton, M., Shaw, C. E., Boylan, K. B., Boeve, B., Graff-Radford, N. R., Wszolek, Z. K., Caselli, R. J., Dickson, D. W., Mackenzie, I. R., Petrucelli, L., and Rademakers, R. (2008) Novel mutations in TARDBP (TDP-43) in patients with familial amyotrophic lateral sclerosis. PLoS Genet. 4, e1000193. (99) Tamaoka, A., Arai, M., Itokawa, M., Arai, T., Hasegawa, M., Tsuchiya, K., Takuma, H., Tsuji, H., Ishii, A., Watanabe, M., Takahashi, Y., Goto, J., Tsuji, S., and Akiyama, H. (2010) TDP-43 M337V mutation in familial amyotrophic lateral sclerosis in Japan. Intern. Med. 49, 331–334. (100) Ju, X., Liu, W., Li, X., Liu, N., Zhang, N., Liu, T., and Deng, M. (2016) Two distinct clinical features and cognitive impairment in amyotrophic lateral sclerosis patients with TARDBP gene mutations in the Chinese population. Neurobiol. Aging 38, 216.e1–216.e6. (101) Yokoseki, A., Shiga, A., Tan, C.-F., Tagawa, A., Kaneko, H., Koyama, A., Eguchi, H., Tsujino, A., Ikeuchi, T., Kakita, A., Okamoto, K., Nishizawa, M., Takahashi, H., and Onodera, O. (2008) TDP-43 mutation in familial amyotrophic lateral sclerosis. Ann. Neurol. 63, 538–542. (102) Kühnlein, P., Sperfeld, A.-D., Vanmassenhove, B., Van Deerlin, V., Lee, V. M.-Y., Trojanowski, J. Q., Kretzschmar, H. A., Ludolph, A. C., and Neumann, M. (2008) Two German kindreds with familial amyotrophic lateral sclerosis due to TARDBP mutations. Arch. Neurol. 65, 1185–1189. 45 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(103) Daoud, H., Valdmanis, P. N., Kabashi, E., Dion, P., Dupré, N., Camu, W., Meininger, V., and Rouleau, G. A. (2009) Contribution of TARDBP mutations to sporadic amyotrophic lateral sclerosis. J. Med. Genet. 46, 112–114. (104) Kamada, M., Maruyama, H., Tanaka, E., Morino, H., Wate, R., Ito, H., Kusaka, H., Kawano, Y., Miki, T., Nodera, H., Izumi, Y., Kaji, R., and Kawakami, H. (2009) Screening for TARDBP mutations in Japanese familial amyotrophic lateral sclerosis. J. Neurol. Sci. 284, 69– 71. (105) Chiò, A., Borghero, G., Pugliatti, M., Ticca, A., Calvo, A., Moglia, C., Mutani, R., Brunetti, M., Ossola, I., Marrosu, M. G., Murru, M. R., Floris, G., Cannas, A., Parish, L. D., Cossu, P., Abramzon, Y., Johnson, J. O., Nalls, M. A., Arepalli, S., Chong, S., Hernandez, D. G., Traynor, B. J., Restagno, G., and Italian Amyotrophic Lateral Sclerosis Genetic (ITALSGEN) Consortium. (2011) Large proportion of amyotrophic lateral sclerosis cases in Sardinia due to a single founder mutation of the TARDBP gene. Arch. Neurol. 68, 594–598. (106) Orrù, S., Manolakos, E., Orrù, N., Kokotas, H., Mascia, V., Carcassi, C., and Petersen, M. B. (2012) High frequency of the TARDBP p.Ala382Thr mutation in Sardinian patients with amyotrophic lateral sclerosis. Clin. Genet. 81, 172–178. (107) Chiang, H.-H., Andersen, P. M., Tysnes, O.-B., Gredal, O., Christensen, P. B., and Graff, C. (2012) Novel TARDBP mutations in Nordic ALS patients. J. Hum. Genet. 57, 316–319. (108) Xiong, H.-L., Wang, J.-Y., Sun, Y.-M., Wu, J.-J., Chen, Y., Qiao, K., Zheng, Q.-J., Zhao, G.-X., and Wu, Z.-Y. (2010) Association between novel TARDBP mutations and Chinese patients with amyotrophic lateral sclerosis. BMC Med. Genet. 11, 8. (109) Nozaki, I., Arai, M., Takahashi, K., Hamaguchi, T., Yoshikawa, H., Muroishi, T., 46 ACS Paragon Plus Environment

Page 46 of 62

Page 47 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

Noguchi-Shinohara, M., Ito, H., Itokawa, M., Akiyama, H., Kawata, A., and Yamada, M. (2010) Familial ALS with G298S mutation in TARDBP: a comparison of CSF tau protein levels with those in sporadic ALS. Intern. Med. 49, 1209–1212. (110) Huey, E. D., Ferrari, R., Moreno, J. H., Jensen, C., Morris, C. M., Potocnik, F., Kalaria, R. N., Tierney, M., Wassermann, E. M., Hardy, J., Grafman, J., and Momeni, P. (2012) FUS and TDP43 genetic variability in FTD and CBS. Neurobiol. Aging 33, 1016.e9–1016.e17. (111) Iida, A., Kamei, T., Sano, M., Oshima, S., Tokuda, T., Nakamura, Y., and Ikegawa, S. (2012) Large-scale screening of TARDBP mutation in amyotrophic lateral sclerosis in Japanese. Neurobiol. Aging 33, 786–790. (112) Zou, Z.-Y., Peng, Y., Wang, X.-N., Liu, M.-S., Li, X.-G., and Cui, L.-Y. (2012) Screening of the TARDBP gene in familial and sporadic amyotrophic lateral sclerosis patients of Chinese origin. Neurobiol. Aging 33, 2229.e11–2229.e18. (113) Lemmens, R., Race, V., Hersmus, N., Matthijs, G., Van Den Bosch, L., Van Damme, P., Dubois, B., Boonen, S., Goris, A., and Robberecht, W. (2009) TDP-43 M311V mutation in familial amyotrophic lateral sclerosis. J. Neurol. Neurosurg. Psychiatry 80, 354–355. (114) Fujita, Y., Ikeda, M., Yanagisawa, T., Senoo, Y., and Okamoto, K. (2011) Different clinical and neuropathologic phenotypes of familial ALS with A315E TARDBP mutation. Neurology 77, 1427–1431. (115) Ticozzi, N., LeClerc, A. L., van Blitterswijk, M., Keagle, P., McKenna-Yasek, D. M., Sapp, P. C., Silani, V., Wills, A.-M., Brown, R. H., and Landers, J. E. (2011) Mutational analysis of TARDBP in neurodegenerative diseases. Neurobiol. Aging 32, 2096–2099.

47 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(116) Tsai, C.-P., Soong, B.-W., Lin, K.-P., Tu, P.-H., Lin, J.-L., and Lee, Y.-C. (2011) FUS, TARDBP, and SOD1 mutations in a Taiwanese cohort with familial ALS. Neurobiol. Aging 32, 553.e13–553.e21. (117) Origone, P., Caponnetto, C., Bandettini Di Poggio, M., Ghiglione, E., Bellone, E., Ferrandes, G., Mancardi, G. L., and Mandich, P. (2010) Enlarging clinical spectrum of FALS with TARDBP gene mutations: S393L variant in an Italian family showing phenotypic variability and relevance for genetic counselling. Amyotroph. Lateral Scler. 11, 223–227. (118) Huang, R., Fang, D.-F., Ma, M.-Y., Guo, X.-Y., Zhao, B., Zeng, Y., Zhou, D., Yang, Y., and Shang, H.-F. (2012) TARDBP gene mutations among Chinese patients with sporadic amyotrophic lateral sclerosis. Neurobiol. Aging 33, 1015.e1–1015.e6. (119) Conforti, F. L., Sproviero, W., Simone, I. L., Mazzei, R., Valentino, P., Ungaro, C., Magariello, A., Patitucci, A., La Bella, V., Sprovieri, T., Tedeschi, G., Citrigno, L., Gabriele, A. L., Bono, F., Monsurrò, M. R., Muglia, M., Gambardella, A., and Quattrone, A. (2011) TARDBP gene mutations in south Italian patients with amyotrophic lateral sclerosis. J. Neurol. Neurosurg. Psychiatry 82, 587–588. (120) Millecamps, S., Salachas, F., Cazeneuve, C., Gordon, P., Bricka, B., Camuzat, A., GuillotNoël, L., Russaouen, O., Bruneteau, G., Pradat, P.-F., Le Forestier, N., Vandenberghe, N., Danel-Brunaud, V., Guy, N., Thauvin-Robinet, C., Lacomblez, L., Couratier, P., Hannequin, D., Seilhean, D., Le Ber, I., Corcia, P., Camu, W., Brice, A., Rouleau, G., LeGuern, E., and Meininger, V. (2010) SOD1, ANG, VAPB, TARDBP, and FUS mutations in familial amyotrophic lateral sclerosis: genotype-phenotype correlations. J. Med. Genet. 47, 554–560. (121) Van Blitterswijk, M., Van Es, M. A., Hennekam, E. A. M., Dooijes, D., Van Rheenen, W., 48 ACS Paragon Plus Environment

Page 48 of 62

Page 49 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

Medic, J., Bourque, P. R., Schelhaas, H. J., Van der Kooi, A. J., De Visser, M., De Bakker, P. I. W., Veldink, J. H., and van den Berg, L. H. (2012) Evidence for an oligogenic basis of amyotrophic lateral sclerosis. Hum. Mol. Genet. 21, 3776–3784. (122) Moreno, F., Rabinovici, G. D., Karydas, A., Miller, Z., Hsu, S. C., Legati, A., Fong, J., Schonhaut, D., Esselmann, H., Watson, C., Stephens, M. L., Kramer, J., Wiltfang, J., Seeley, W. W., Miller, B. L., Coppola, G., and Grinberg, L. T. (2015) A novel mutation P112H in the TARDBP gene associated with frontotemporal lobar degeneration without motor neuron disease and abundant neuritic amyloid plaques. Acta Neuropathol. Commun. 3, 19. (123) Borroni, B., Archetti, S., Del Bo, R., Papetti, A., Buratti, E., Bonvicini, C., Agosti, C., Cosseddu, M., Turla, M., Di Lorenzo, D., Pietro Comi, G., Gennarelli, M., and Padovani, A. (2010) TARDBP mutations in frontotemporal lobar degeneration: frequency, clinical features, and disease course. Rejuvenation Res. 13, 509–517. (124) Qin, H., Lim, L., Wei, Y., and Song, J. (2014) TDP-43 N terminus encodes a novel ubiquitin-like fold and its unfolded form in equilibrium that can be shifted by binding to ssDNA. Proc. Natl. Acad. Sci. U. S. A. 111, 18619–18624. (125) Mompean, M., Romano, V., Pantoja-Uceda, D., Stuani, C., Baralle, F. E., Buratti, E., and Laurents, D. V. (2016) The TDP-43 N-Terminal Domain Structure at High Resolution. FEBS J. 283, 1–19. (126) Wang, Y. T., Kuo, P. H., Chiang, C. H., Liang, J. R., Chen, Y. R., Wang, S., Shen, J. C. K., and Yuan, H. S. (2013) The truncated C-terminal RNA recognition motif of TDP-43 protein plays a key role in forming proteinaceous aggregates. J. Biol. Chem. 288, 9049–9057. (127) Budini, M., Buratti, E., Stuani, C., Guarnaccia, C., Romano, V., De Conti, L., and Baralle, 49 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

F. E. (2012) Cellular model of TAR DNA-binding protein 43 (TDP-43) aggregation based on its C-terminal Gln/Asn-rich region. J. Biol. Chem. 287, 7512–7525. (128) Budini, M., Romano, V., Quadri, Z., Buratti, E., and Baralle, F. E. (2015) TDP-43 loss of cellular function through aggregation requires additional structural determinants beyond its Cterminal Q/N prion-like domain. Hum. Mol. Genet. 24, 9–20. (129) Kuo, P. H., Chiang, C. H., Wang, Y. T., Doudeva, L. G., and Yuan, H. S. (2014) The crystal structure of TDP-43 RRM1-DNA complex reveals the specific recognition for UG- and TG-rich nucleic acids. Nucleic Acids Res. 42, 4712–4722. (130) Chiang, C.-H., Grauffel, C., Wu, L.-S., Kuo, P.-H., Doudeva, L. G., Lim, C., Shen, C.-K. J., and Yuan, H. S. (2016) Structural analysis of disease-related TDP-43 D169G mutation: linking enhanced stability and caspase cleavage efficiency to protein accumulation. Sci. Rep. 6, 21581. (131) Lukavsky, P. J., Daujotyte, D., Tollervey, J. R., Ule, J., Stuani, C., Buratti, E., Baralle, F. E., Damberger, F. F., and Allain, F. H.-T. (2013) Molecular basis of UG-rich RNA recognition by the human splicing factor TDP-43. Nat. Publ. Gr. 20, 1443–1449. (132) Maris, C., Dominguez, C., and Allain, F. H. T. (2005) The RNA recognition motif, a plastic RNA-binding platform to regulate post-transcriptional gene expression. FEBS J. 272, 2118–2131. (133) Tollervey, J. R. J., Curk, T., Rogelj, B., Briese, M., Cereda, M., Ule, J., Tollervey, J. R. J., Curk, T., and Rogelj, B. (2011) Characterising the RNA targets and position-dependent splicing regulation by TDP-43; implications for neurodegenerative diseases. October 14, 452–458.

50 ACS Paragon Plus Environment

Page 50 of 62

Page 51 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

(134) Xiao, S., Sanelli, T., Dib, S., Sheps, D., Findlater, J., Bilbao, J., Keith, J., Zinman, L., Rogaeva, E., and Robertson, J. (2011) RNA targets of TDP-43 identified by UV-CLIP are deregulated in ALS. Mol. Cell. Neurosci. 47, 167–180. (135) Zhang, Y.-J., Xu, Y.-F., Cook, C., Gendron, T. F., Roettges, P., Link, C. D., Lin, W.-L., Tong, J., Castanedes-Casey, M., Ash, P., Gass, J., Rangachari, V., Buratti, E., Baralle, F., Golde, T. E., Dickson, D. W., and Petrucelli, L. (2009) Aberrant cleavage of TDP-43 enhances aggregation and cellular toxicity. Proc. Natl. Acad. Sci. U. S. A. 106, 7607–7612. (136) Zhang, Y.-J., Xu, Y., Dickey, C. a, Buratti, E., Baralle, F., Bailey, R., Pickering-Brown, S., Dickson, D., and Petrucelli, L. (2007) Progranulin mediates caspase-dependent cleavage of TAR DNA binding protein-43. J. Neurosci. 27, 10530–10534. (137) Gitler, A. D., and Shorter, J. (2011) RNA-binding proteins with prion-like domains in ALS and FTLD-U. Prion 5, 179–187. (138) Jiang, L. L., Che, M. X., Zhao, J., Zhou, C. J., Xie, M. Y., Li, H. Y., He, J. H., and Hu, H. Y. (2013) Structural transformation of the amyloidogenic core region of TDP-43 protein initiates its aggregation and cytoplasmic inclusion. J. Biol. Chem. 288, 19614–19624. (139) Liu, G. C.-H., Chen, B. P.-W., Ye, N. T.-J., Wang, C.-H., Chen, W., Lee, H.-M., Chan, S. I., and Huang, J. J.-T. (2013) Delineating the membrane-disrupting and seeding properties of the TDP-43 amyloidogenic core. Chem. Commun. (Camb). 49, 11212–11214. (140) Mompeán, M., Hervás, R., Xu, Y., Tran, T. H., Guarnaccia, C., Buratti, E., Baralle, F. E., Tong, L., Carrión-Vázquez, M., McDermott, A. E., and Laurents, D. V. (2015) Structural Evidence of Amyloid Fibril Formation in the Putative Aggregation Domain of TDP-43. J. Phys. Chem. Lett. 6, 2608–2615. 51 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(141) D’Ambrogio, A., Buratti, E., Stuani, C., Guarnaccia, C., Romano, M., Ayala, Y. M., and Baralle, F. E. (2009) Functional mapping of the interaction between TDP-43 and hnRNP A2 in vivo. Nucleic Acids Res. 37, 4116–4126. (142) Buratti, E., Brindisi, A., Giombi, M., Tisminetzky, S., Ayala, Y. M., and Baralle, F. E. (2005) 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. 280, 37572–37584. (143) Conicella, A. E., Zerze, G. H., Mittal, J., and Fawzi, N. L. (2016) ALS Mutations Disrupt Phase Separation Mediated by α-Helical Structure in the TDP-43 Low-Complexity C-Terminal Domain. Structure 1–13. (144) Mompeán, M., Buratti, E., Guarnaccia, C., Brito, R. M. M., Chakrabartty, A., Baralle, F. E., and Laurents, D. V. (2014) “Structural characterization of the minimal segment of TDP-43 competent for aggregation”. Arch. Biochem. Biophys. 545, 53–62. (145) Schmidt, H. B., and Rohatgi, R. (2016) In Vivo Formation of Vacuolated Multi-phase Report In Vivo Formation of Vacuolated Multi-phase Compartments Lacking Membranes. CellReports 16, 1–9. (146) Cox, B. S. (1965) Ψ, A cytoplasmic suppressor of super-suppressor in yeast. Heredity (Edinb). 20, 505–521. (147) Tuite, M. F., Staniforth, G. L., and Cox, B. S. (2015) [PSI(+)] turns 50. Prion 9, 318–332. (148) Holmes, D. L., Lancaster, A. K., Lindquist, S., and Halfmann, R. (2013) Heritable remodeling of yeast multicellularity by an environmentally responsive prion. Cell 153, 153–165.

52 ACS Paragon Plus Environment

Page 52 of 62

Page 53 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

(149) Wickner, R. B., Edskes, H. K., Kryndushkin, D., McGlinchey, R., Bateman, D., and Kelly, A. (2011) Prion diseases of yeast: Amyloid structure and biology. Semin. Cell Dev. Biol. 22, 469–475. (150) Halfmann, R., and Lindquist, S. (2010) Epigenetics in the extreme: prions and the inheritance of environmentally acquired traits. Science 330, 629–632. (151) Wickner, R. B., Edskes, H. K., Gorkovskiy, A., Bezsonov, E. E., and Stroobant, E. E. (2016) Yeast and Fungal Prions: Amyloid-Handling Systems, Amyloid Structure, and Prion Biology. Adv. Genet. 93, 191-236. (152) Kwiatkowski, T. J., Bosco, D. A., Leclerc, a L., Tamrazian, E., Vanderburg, C. R., Russ, C., Davis, A., Gilchrist, J., Kasarskis, E. J., Munsat, T., Valdmanis, P., Rouleau, G. a, Hosler, B. A., Cortelli, P., de Jong, P. J., Yoshinaga, Y., Haines, J. L., Pericak-Vance, M. a, Yan, J., Ticozzi, N., Siddique, T., McKenna-Yasek, D., Sapp, P. C., Horvitz, H. R., Landers, J. E., and Brown, R. H. (2009) Mutations in the FUS/TLS gene on chromosome 16 cause familial amyotrophic lateral sclerosis. Science 323, 1205–1208. (153) Vance, C., Rogelj, B., Hortobágyi, T., De Vos, K. J., Nishimura, A. L., Sreedharan, J., Hu, X., Smith, B., Ruddy, D., Wright, P., Ganesalingam, J., Williams, K. L., Tripathi, V., Al-Saraj, S., Al-Chalabi, A., Leigh, P. N., Blair, I. P., Nicholson, G., de Belleroche, J., Gallo, J.-M., Miller, C. C., and Shaw, C. E. (2009) Mutations in FUS, an RNA Processing Protein, Cause Familial Amyotrophic Lateral Sclerosis Type 6. Science 323, 1208–1211. (154) Kim, H. J., Kim, N. C., Wang, Y.-D., Scarborough, E. A., Moore, J., Diaz, Z., MacLea, K. S., Freibaum, B., Li, S., Molliex, A., Kanagaraj, A. P., Carter, R., Boylan, K. B., Wojtas, A. M., Rademakers, R., Pinkus, J. L., Greenberg, S. A., Trojanowski, J. Q., Traynor, B. J., Smith, B. N., 53 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Topp, S., Gkazi, A.-S., Miller, J., Shaw, C. E., Kottlors, M., Kirschner, J., Pestronk, A., Li, Y. R., Ford, A. F., Gitler, A. D., Benatar, M., King, O. D., Kimonis, V. E., Ross, E. D., Weihl, C. C., Shorter, J., and Taylor, J. P. (2013) Mutations in prion-like domains in hnRNPA2B1 and hnRNPA1 cause multisystem proteinopathy and ALS. Nature 495, 467–473. (155) Chong, P. A., and Forman-Kay, J. D. (2016) A New Phase in ALS Research. Structure 24, 1435–1436. (156) Mitrea, D. M., and Kriwacki, R. W. (2016) Phase separation in biology; functional organization of a higher order. Cell Commun. Signal. 14, 1. (157) Brangwynne, C. P., Eckmann, C. R., Courson, D. S., Rybarska, A., Hoege, C., Gharakhani, J., Jülicher, F., and Hyman, A. A. (2009) Germline P granules are liquid droplets that localize by controlled dissolution/condensation. Science 324, 1729–1732. (158) Buchan, J. R. (2014) mRNP granules. Assembly, function, and connections with disease. RNA Biol. 11, 1019–1030. (159) Buchan, J. R., Nissan, T., and Parker, R. (2010) Analyzing P-bodies and stress granules in Saccharomyces cerevisiae. Methods Enzymol. 470, 619-640. (160) Molliex, A., Temirov, J., Lee, J., Coughlin, M., Kanagaraj, A. P., Kim, H. J., Mittag, T., and Taylor, J. P. (2015) Phase Separation by Low Complexity Domains Promotes Stress Granule Assembly and Drives Pathological Fibrillization. Cell 163, 123–133. (161) Liu-Yesucevitz, L., Bilgutay, A., Zhang, Y.-J., Vanderweyde, T., Vanderwyde, T., Citro, A., Mehta, T., Zaarur, N., McKee, A., Bowser, R., Sherman, M., Petrucelli, L., and Wolozin, B. (2010) Tar DNA binding protein-43 (TDP-43) associates with stress granules: analysis of

54 ACS Paragon Plus Environment

Page 54 of 62

Page 55 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

cultured cells and pathological brain tissue. PLoS One 5, e13250. (162) Elbaum-Garfinkle, S., and Brangwynne, C. P. (2015) Liquids, Fibers, and Gels: The Many Phases of Neurodegeneration. Dev. Cell 35, 531–532. (163) Burke, K. A., Janke, A. M., Rhine, C. L., and Fawzi, N. L. (2015) Residue-by-Residue View of In Vitro FUS Granules that Bind the C-Terminal Domain of RNA Polymerase II. Mol. Cell 60, 231–241. (164) Patel, A., Lee, H. O., Jawerth, L., Maharana, S., Jahnel, M., Hein, M. Y., Stoynov, S., Mahamid, J., Saha, S., Franzmann, T. M., Pozniakovski, A., Poser, I., Maghelli, N., Royer, L. A., Weigert, M., Myers, E. W., Grill, S., Drechsel, D., Hyman, A. A., and Alberti, S. (2015) A Liquid-to-Solid Phase Transition of the ALS Protein FUS Accelerated by Disease Mutation. Cell 162, 1066–1077. (165) Murakami, T., Qamar, S., Lin, J. Q., Schierle, G. S. K., Rees, E., Miyashita, A., Costa, A. R., Dodd, R. B., Chan, F. T. S., Michel, C. H., Kronenberg-Versteeg, D., Li, Y., Yang, S. P., Wakutani, Y., Meadows, W., Ferry, R. R., Dong, L., Tartaglia, G. G., Favrin, G., Lin, W. L., Dickson, D. W., Zhen, M., Ron, D., Schmitt-Ulms, G., Fraser, P. E., Shneider, N. A., Holt, C., Vendruscolo, M., Kaminski, C. F., and St George-Hyslop, P. (2015) ALS/FTD MutationInduced Phase Transition of FUS Liquid Droplets and Reversible Hydrogels into Irreversible Hydrogels Impairs RNP Granule Function. Neuron 88, 678–690. (166) Xiang, S., Kato, M., Wu, L. C., Lin, Y., Ding, M., Zhang, Y., Yu, Y., and McKnight, S. L. (2015) The LC Domain of hnRNPA2 Adopts Similar Conformations in Hydrogel Polymers, Liquid-like Droplets, and Nuclei. Cell 163, 829–839. (167) Phan, A. T., Kuryavyi, V., Darnell, J. C., Serganov, A., Majumdar, A., Ilin, S., Raslin, T., 55 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Polonskaia, A., Chen, C., Clain, D., Darnell, R. B., and Patel, D. J. (2011) Structure-function studies of FMRP RGG peptide recognition of an RNA duplex-quadruplex junction. Nat. Struct. Mol. Biol. 18, 796–804. (168) Jiang, L.-L., Zhao, J., Yin, X.-F., He, W.-T., Yang, H., Che, M.-X., and Hu, H.-Y. (2016) Two mutations G335D and Q343R within the amyloidogenic core region of TDP-43 influence its aggregation and inclusion formation. Sci. Rep. 6, 23928. (169) Kato, M., Han, T. W., Xie, S., Shi, K., Du, X., Wu, L. C., Mirzaei, H., Goldsmith, E. J., Longgood, J., Pei, J., Grishin, N. V., Frantz, D. E., Schneider, J. W., Chen, S., Li, L., Sawaya, M. R., Eisenberg, D., Tycko, R., and McKnight, S. L. (2012) Cell-free formation of RNA granules: Low complexity sequence domains form dynamic fibers within hydrogels. Cell 149, 753–767. (170) Saini, A., and Chauhan, V. S. (2011) Delineation of the core aggregation sequences of TDP-43 C-terminal fragment. Chembiochem 12, 2495–2501. (171) Jain, S., Wheeler, J. R., Walters, R. W., Agrawal, A., Barsic, A., and Parker, R. (2016) ATPase-Modulated Stress Granules Contain a Diverse Proteome and Substructure. Cell 164, 487–498. (172) Lin, Y., Protter, D. S. W., Rosen, M. K., and Parker, R. (2015) Formation and Maturation of Phase-Separated Liquid Droplets by RNA-Binding Proteins. Mol. Cell 60, 208–219. (173) Buchan, J. R., Yoon, J.-H., and Parker, R. (2011) Stress-specific composition, assembly and kinetics of stress granules in Saccharomyces cerevisiae. J. Cell Sci. 124, 228–239. (174) Lagier-Tourenne, C., Polymenidou, M., and Cleveland, D. W. (2010) TDP-43 and FUS/TLS: emerging roles in RNA processing and neurodegeneration. Hum. Mol. Genet. 19,

56 ACS Paragon Plus Environment

Page 56 of 62

Page 57 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

R46–R64. (175) Zhang, H., Elbaum-Garfinkle, S., Langdon, E. M., Taylor, N., Occhipinti, P., Bridges, A. A., Brangwynne, C. P., and Gladfelter, A. S. (2015) RNA Controls PolyQ Protein Phase Transitions. Mol. Cell 60, 220–230. (176) Furukawa, Y., Kaneko, K., Watanabe, S., Yamanaka, K., and Nukina, N. (2011) A seeding reaction recapitulates intracellular formation of Sarkosyl-insoluble transactivation response element (TAR) DNA-binding protein-43 inclusions. J. Biol. Chem. 286, 18664–18672. (177) Chen, A. K. H., Lin, R. Y. Y., Hsieh, E. Z. J., Tu, P. H., Chen, R. P. Y., Liao, T. Y., Chen, W., Wang, C. H., and Huang, J. J. T. (2010) Induction of amyloid fibrils by the C-terminal fragments of TDP-43 in amyotrophic lateral sclerosis. J. Am. Chem. Soc. 132, 1186–1187. (178) Feiler, M. S., Strobel, B., Freischmidt, A., Helferich, A. M., Kappel, J., Brewer, B. M., Li, D., Thal, D. R., Walther, P., Ludolph, A. C., Danzer, K. M., and Weishaupt, J. H. (2015) TDP-43 is intercellularly transmitted across axon terminals. J. Cell Biol. 211, 897–911. (179) Zeineddine, R., Pundavela, J. F., Corcoran, L., Stewart, E. M., Do-Ha, D., Bax, M., Guillemin, G., Vine, K. L., Hatters, D. M., Ecroyd, H., Dobson, C. M., Turner, B. J., Ooi, L., Wilson, M. R., Cashman, N. R., and Yerbury, J. J. (2015) SOD1 protein aggregates stimulate macropinocytosis in neurons to facilitate their propagation. Mol. Neurodegener. 10, 57. (180) Lim, L., Wei, Y., Lu, Y., and Song, J. (2016) ALS-Causing Mutations Significantly Perturb the Self-Assembly and Interaction with Nucleic Acid of the Intrinsically Disordered Prion-Like Domain of TDP-43. PLoS Biol. 14, e1002338. (181) Ang, J., Ma, D., Lund, R., Keten, S., and Xu, T. (2016) Internal Structure of 15 nm 3-Helix

57 ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Micelle Revealed by Small-Angle Neutron Scattering and Coarse-Grained MD Simulation. Biomacromolecules 17, 3262–3267.

58 ACS Paragon Plus Environment

Page 58 of 62

Page 59 of 62

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

Figure 1 A

ALS/FTD mutations

A90V 1

102

NTD

K263E N267S G287S G290A S292N G294A/V G295C/R/S G298S M311V

A315T/E A321G/V Q331K S332N G335D M337V Q343R N345K G348C/R/V

N352S/T G357R/S M359V R361S/T P363A G368S Y374X G376D N378D/S

S379C/P A382P/T I383V G384R W385G S387indel N390S/D S393L

P112H D169G 106

L

177 192

RRM1

259 274

RRM2

E

414

CTD

B NTD β1 β2 α1 β3 1 MSEY TED ENDEP PS EDDGTV PGAC NPVSQC β4 β5 β6 51 MRGV EG APDAGWGN L YPKD NKRKMDETDA SSAVKVKRAV NLS 101 QK

C

RRM1 β1 α1 106 GLPWKT T α2 β4 β5 156 R DG RRM2 β1 192 242

β2 TFGEVL

FTE

L

α1

RCTED M β4 β5 LCGED K G NES

β3 KDLKTGHSK G

β2 QYGD

β3 IPKPFRA

α2 ADD

N

D CTD 318 IN

helix-turn helix SS

QNQS GPSGNNQNQG NMQREPNQA simulated β-strands

ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2

ACS Paragon Plus Environment

Page 60 of 62

Page 61 of 62

Figure 3 α-helix

Nuclear TDP-43 Functional dimer; RNA/DNA binding

Stress; N-terminal oligomerization; α-helical interactions

Stress granule TDP-43 LLPS; Contains other RNA and RNPs

Cytoplasmic shuttling

PrLD secondary structure

fragmentation

α-helix to β-strand conversion

N-terminal oligomerization; C-terminal aggregation

Expulsion of infectious agent

β-strand

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

Disruption to SG dynamics Persistent SGs

Persistent or ”Aged” SGs

ATP-dependent remodeling

Irreversible aggregates

Stress granule “core”

Amorphous; Phosphorylated; Ubiquitinated

Hydrogel-like; β-rich

intracellular Spreading agent uncharacterized

extracellular

ACS Paragon Plus Environment

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

For ToC use only Į-helix

TDP-43 PrLD 2º structure /LTXLGGURSOHWV

1DWLYHQXFOHDU

ß-strand 3DWKRORJLFDO DJJUHJDWHV

ZLWKJHOOLNHFRUHV

ACS Paragon Plus Environment

Page 62 of 62