+ All Categories
Home > Documents > UNIVERSITY OF NOVA GORICA GRADUATE SCHOOLlibrary/doktorati/genetika/13Safini.pdfUNIVERSITY OF NOVA...

UNIVERSITY OF NOVA GORICA GRADUATE SCHOOLlibrary/doktorati/genetika/13Safini.pdfUNIVERSITY OF NOVA...

Date post: 29-Apr-2018
Category:
Upload: phungdieu
View: 215 times
Download: 1 times
Share this document with a friend
143
UNIVERSITY OF NOVA GORICA GRADUATE SCHOOL ROLE OF AUTOPHAGY AND ITS ENHANCING IN THE CLEARANCE OF TDP-43 AGGREGATES DISSERTATION Naj e t e Safini Mentor: Dr.Sergio G. Tisminetzky Nova Gorica, 2014
Transcript

UNIVERSITY OF NOVA GORICA GRADUATE SCHOOL

R O L E O F A U T OPH A G Y A ND I TS E N H A N C IN G IN T H E C L E A R A N C E O F T DP-43 A G G R E G A T ES

DISSERTATION

Najete Safini

Mentor: Dr.Sergio G. Tisminetzky

Nova Gorica, 2014

UNIVERZA V NOVI GORICI

FAKULTETA ZA PODIPLOMSKI !TUDIJ

VLOGA AVTOFAGIJE IN NJEN VPLIV NA IZBOLJ!ANJE

"I!"ENJA AGREGATOV TDP-43

DISERTACIJA

Najete Safini

Mentor: dr. Sergio G. Tisminetzky

Nova Gorica, 2014

2

A BST R A C T

Neurodegenerative diseases, such as Amyotrophic Lateral Sclerosis (ALS) and

Fronto-Temporal Lobar Degeneration (FTLD), are characterized by imbalance

between generation and degradation of misfolded TDP-43 protein, resulting in the

formation of cytoplasmic inclusions followed by a loss of nuclear TDP-43 in affected

neuronal cells. Previously established cell-based TDP-43 aggregation model, where

tandem repeats carrying the C-terminal Gln/Asn-rich region of TDP-43 (EGFP-

12XQ/N) were able to trigger the formation of phosphorylated and ubiquitinated

aggregates, have been used in this study to investigate the autophagy-endolysosomal

cell pathway response. Cotransfection with EGFP-12xQ/N and HcRed-hLC3 (LC3

Light chain 3is marker of autophagosome and autophagy induction), immunostaining

with anti-p62 (p62 is an autophagy adaptor) and anti-Lamp1 (Lamp1 is marker of

late endosome/lysosome marker) show that EGFP-12xQ/N is able to generate

cytoplasmic aggregates that have the ability to induce an autophagy-endolysosomal

pathway cell response. Observed autophagic response incidence in EGFP-12XQ/N

aggregates led us to try Dextran Sulfate 5000 Da (DS), previously associated with

the autophagy induction, as a therapeutic effector aimed at reducing the aggregation

phenomenon. We found that DS treatment enhances inclusions clearance and

significantly increase the proportion of the LC3-positive aggregates, suggesting an

improved autophagic response of the cell. In addition, it was shown that DS

decreases the amount of flag-TDP43 wild type in the insoluble fraction suggesting its

release from the aggregates sequestration. Observed DS-triggered clearance is in

correlation with the increased LC3-II/LC3-I ratio, as well as with the significant p62

decay in the treated cells. It has been demonstrated that DS-induced EGFP-12xQ/N

clearance was abolished in an autophagy-deficient cells confirming DS clearing

potential through the autophagy pathway activation. Furthermore, we have identified

the histone deacetylase-6 (HDAC6), as an important player of DS clearance action.

Indeed, DS treatment failed to clear EGFP-12xQ/N in the HDAC6 deficient cells,

while by re-introducing HDAC6 wild type protein DS-triggered clearance has been

restored. Moreover, it has been demonstrated that DS-induced EGFP-12xQ/N

product decay requires catalytically active HDAC6.

Our study suggests that DS may help the cell to clear TDP-43 aggregates, in this

model, through the autophagy resulting in the more efficient cargo degradation in

3

HDAC6-dependent manner, thus revealing DS therapeutic potential in TDP-43

proteinopathies.

K eywords: Dextran sulfate (DS), Autophagy, TDP-43 proteinopathies, Amyotrophic Lateral

Sclerosis (ALS) and Frontotemporal Lobar Degeneration (FTLD).

4

Izvleček

Nevrodegenerativna obolenja, kot so amiotrofična lateralna skleroza (ALS) in

frontotemporalna lobarna degeneracija (FTLD), so posledica neravnovesja med

tvorbo in degradacijo nepravilno zvitega TDP-43 proteina, kar vodi do kopičenja

TDP-43 v citoplazmi v obliki vključkov in izgube v jedrih prizadetih nevronskih

celic. V raziskavah sem za študij odgovora avtofagne-endolizosomalne celične poti

uporabila predhodno uveljavljen celični model tvorbe agregatov TDP-43, kjer

tandemske ponovitve vključujejo C-terminalno Gln/Asn-bogato regijo proteina TDP-

43 (EGFP-12XQ/N) in povzročijo tvorbo fosforiliranih in ubikvitiniranih agregatov.

Sočasna transfekcija EGFP-12xQ/N in HcRed-hLC3 ter imunooznačevanje s

protitelesi proti p62 in Lamp1 so pokazali, da je model EGFP-12xQ/N zmožen

tvorbe agregatov v citoplazmi, kar lahko sprožiavtofagno-endolizosomalni celični

odgovor. Zaradi opaženega avtofagnega odgovora v agregatih EGFP-12XQ/N smo

preverjali možen terapevtski vpliv dekstran sulfata 5000 Da (DS), ki je bil predhodno

povezan s pojavom avtofagije in za katerega smo predpostavljali, da lahko zmanjša

proces agregacije. Ugotovili smo, da tretiranje z DS povzroži degradacijo vključkov

in signifikantno poveča delež LC3-pozitivnih agregatov, kar nakazuje na izboljšan

avtofagni odgovor celic. V nadaljevanju smo pokazali, da DS zmanjša količino

proteinov flag-TDP43 divjega tipa v netopni frakciji, z možnim vzrokom sprostitve

le-teh iz nastajajočih agregatov. Ugotovili smo, da je opaženo zmanjšanje agregatov

zaradi aktivnosti DS v povezavi s povečanim razmerjem LC3-II/LC3-I ter

signifikantnim zmanjšanjem p62 v tretiranih celicah. Prav tako smo pokazali, da z

DS-inducirana degradacija agregatov ni prisotna v celicah nezmožnih avtofagije, kar

je potrdilo velik potencial DS pri zmanjšanju agregatov in aktivaciji avtofagne poti.

V nadaljevanju smo kot pomemben člen pri degradaciji vključkov z DS določili

histonsko deacetilazo 6 (HDAC6). Tretiranje z DS ni povzročilo izgube EGFP-

12xQ/N v celicah brez HDAC6 in ponovno izražanje proteina divjega tipa HDAC6

je povrnilo degradacijo vključkov z DS. Pokazali smo, da je za tovrstno aktivnost

potreben katalitično aktiven HDAC6.

Naše raziskave nakazujejo, da lahko DS pomembno pripomore pri uničenju

agregatov TDP-43. V predstavljenem modelu preko avtofagije povzroči učinkovito

5

degradacijo, odvisno od aktivnosti HDAC6, kar predstavlja pomemben terapevstki

potencial DS pri TDP-43 proteinopatijah.

Ključne besede: dekstran sulfat (DS), avtofagija, TDP-43 proteinopatije, amiotrofična lateralna

skleroza (ALS) in frontotemporalna lobarna degeneracija (FTLD).

6

A C K N O W L E D G M E N TS

I would like to express my gratitude to my Thesis mentor, Dr. Sergio Tisminetzky,

for allowing me to work on this project in his lab and for his great supervision and

support throughout my PhD experience.

A very special thank go out to Dr. Natasa Skoko, for her understanding, and

patience. I appreciate her vast knowledge and skill in many areas, and her close

assistance on the bench and also in writing reports (i.e., proposals, and this

thesis),which without her this work would have never happened.

Prof. Francesco Baralle, and Dr. Marco Baralle for allowing me the opportunity to

use their aggregation model (EGFP-12xQ/N construct), and for their very helpful

discussions and suggestions.

The members of the commity for reading carefully this thesis and offering their

precious time to make my dissertation better. I really appreciate the time you have

taken to evaluate this work.

I am deeply grateful to all the current biotechnology development group members for

their encouragement, support, and help.

I would like to thank Mojca Tajnik for the translation of my abstract to Slovinien,

Thank you so much (hvala).

Thanks to Dr. Maurizio Budini for the pEGFP-12xQ/N construct, for Dr. Francesca

Arnoldi for the siATG7, for Dr Francesca DeMarchi and Dr Isei Tanida for the

HcRed-LC3 plasmid.

I would like to extend my acknowledgments to all wonderful people and friends I

had the honor to meet them during my stay in ICGEB. Especially, mio Sylvia

DellaMea and her family, Fatemeh, Cristiana, Maureen, and Betul. Your support and

friendship have been priceless.

7

And finally, I would like to thank deeply my parent, and sisters for their

unconditional love and support, which made all of this possible. And also, for my ex-

boyfriend Mohammad, you were really a good person that marked my life and will

stay in my heart, thanks for your support, I will never forget that.

8

Abbreviations ALS Amyotrophic lateral sclerosis AD Alzheimer’s disease ALIS Aggresome-like inducible structures ALS Amyotrophic lateral sclerosis a-Syn alpha-Synuclein Atg Autophagy-related genes CMA Chaperone-mediated autophagy Cvt Cytoplasm to vacuole targeting ESCRT Endosomal sorting complex required for transport FIP200 Focal adhesion kinase family interacting protein of 200 kDa FUS/TLS fused in sarcoma/translocated in liposarcoma HD Huntington’s diseases HDAC6 Histone deacetylase 6 HDL high-density lipoprotein KO Knock out LAMP2A Lysosome-associated membrane protein type-2A LDL low-density lipoprotein LC3 Light chain 3 LIR LC3-interacting region mAtg Mammalian Autophagy geneMTOC Microtubule-organizing centre MVB Multivesicular body NBR1 Neighbour of BRCA1 gene 1 NES Nuclear export sequence ROS Oxygene species SCA Spinocerebellar ataxia SOD1 Cu/Zn superoxide dismutase 1 SQSTM1 Sequestome 1 TOR Target of rapamycin TORC1/2 TOR complex 1 or 2 TARDBP TDP-43 Tar DNA-binding protein UBA Ubiquitin-associated ULK1 Unc-51-like kinase 1 UPS Ubiquitin-proteasome system UVRAG Ultraviolet irradiation resistance-associated gene VCP Valosin-containing protein VLDL very low-density lipoprotein

9

C O N T E N TS

A BST R A C T ....................................................................................... 2

Izvleček ............................................................................................... 4

A C K N O W L E D G M E N TS ................................................................ 6

Abbreviations .................................................................................... 8

C O N T E N TS....................................................................................... 9

L ist of figures ................................................................................... 12

IN T R O DU C T I O N .......................................................................... 14

1.1. Protein degradation ................................................................. 14

1.1.1 UPS pathway:......................................................................... 15

1.1.2 Autophagy: ............................................................................. 19

1.1.2.1 Physiological role of Autophagy: ....................................... 28

1.1.2.2. Selective autophagy ........................................................... 30

1.1.2.3. Aggrephagy ......................................................................... 31

1.2. Autophagy and proteinopathies ............................................. 34

1.3 Amyotrophic Lateral Sclerosis ................................................ 37

1.3.1 G enetics of Amyotrophic Lateral Sclerosis ......................... 39

1.3.2 T DP-43 aggregation ............................................................... 43

1.4 T reatment of A LS ..................................................................... 49

1.5 A LS treatment through autophagy induction ....................... 53

1.6 Dextran sulfate as inducer of autophagy ................................ 55

2. M A T E RI A LS A ND M E T H O DS .......................................... 59

2.1. Chemical reagents: ................................................................ 59

2.2. Standard solutions: ................................................................ 59

10

2.3. Cells: ....................................................................................... 60

2.4. Cell growth media: ................................................................ 60

2.5. Preparation of bacter ial competent cells: ........................... 61

2.6. Amplification of selected DN A fragments (PC R): ............. 61

2.7. Enzymatic modifications of DN A : ....................................... 62

2.8. T ransformation of bacteria: ................................................. 63

2.9. RN A preparation from cultured cells: ................................ 64

2.10. Estimation of nucleic acid concentration: ........................... 64

2.11. cDN A preparation and R T-PC R: ........................................ 64

2.12. Mammalian cell line and cell culture conditions: ............... 65

2.13. Plasmids: ................................................................................ 66

2.14. T ransfection: .......................................................................... 66

2.15. G eneration of F lp-In™ T-Rex™ HEK 293 cells that

inducibly expressing E G FP-12xQ/N: ............................................ 67

2.16. Antibodies: ............................................................................. 68

2.17. RN A interference (siH D A C6) and adding-back exper iment

(H D A C6-W T ,H D A C6-D C): ........................................................... 68

2.18. SDS-PA G E : ............................................................................ 69

2.19. W estern blot analysis: ........................................................... 69

2.20. Immunofluorescence microscopy: ....................................... 70

2.21. C learance assay in stable H ek293 cells expressing E G FP-

12xQ/N ............................................................................................. 71

2.22. V iable cell count determination: .......................................... 71

3. R ESU L TS ............................................................................... 72

3.1. Cellular model of T DP-43 aggregation and autophagy-

lysosome pathway involvement ..................................................... 72

3.2. Dextran sulfate 5000 Da decreases the number of cells

containing E G FP-12xQ /N aggregates ........................................... 78

11

3.3 Positive trend in autophagic response to the aggregation

upon DS treatment .......................................................................... 80

3.4. Establishing stable cell model of E G FP-12xQ /N aggregation

82

3.5. Dextran sulfate 5000 Da enhances the clearance of E G FP-

12xQ/N aggregates in stable H E K 293 cell line ............................ 83

3.6. DS induces E G FP-12xQ /N aggregates clearance through the

autophagy pathway ......................................................................... 87

3.7. E G FP-12xQ/N aggregates are able to sequester T DP-43 wild

type in the cytoplasm ...................................................................... 91

3.8. DS decreases the capture of flag-T DP43 wild type in the

insoluble fraction ............................................................................. 95

3.9. DS-induced E G FP-12xQ /N clearance requires H D A C6 ...... 99

3.10. Mouse motor neuronal-like cell line NSC-34 model of

E G FP-12xQ/N aggregation .......................................................... 103

3.12. Dextran sulfate 5000 Da is not toxic for the motor

neuronal-like cell line NSC-34 ..................................................... 104

4. D ISC USSI O N ............................................................................ 106

5.C O N C L USI O NS ........................................................................ 119

6. B IB L I O G R APH Y .................................................................... 120

12

L ist of figures

F igure 1. Intracellular protein degradation via two main proteolytic systems: the ubiquitin-proteasome system (UPS) and the lysosomal autophagy pathway………18 F igure 2. Schematic representation of macroautophagy……………………..........20 F igure 3. Molecular pathway of autophagy initiation……………………………..23 F igure 4. Autophagy elongation………………………...........................................25 F igure 5. Atg12 and Atg8/LC3 ubiquitin-like conjugation pathways required for autophagosome formation……………………………………………………….....25 F igure 6. Molecules involved in the maturation step of autophagy………………..28 F igure 7. Aggrephagy main players ……………………………………………......34 F igure 8. TDP-43 proteinstructure…………………………..…………………….43 F igure 9. Chemical structure of dextran sulfate……………………………………56 F igure 10. EGFP-12x Q/N aggregates are labeled with LC3, p62 and Lamp1……75 F igure 11. Dextran sulfate 5000 Da decreases the number of cells containing EGFP-12xQ/N aggregates………………………………………………………………….78 F igure 12. Positive trend in autophagic response to the aggregation upon DS treatment………………………………………………………………….................80 F igure 13. EGFP-12xQ/N aggregates formed in HEK293 Flp-In T-Rex stable cell line after tetracycline induction…………………………………………………… 82 F igure 14. Dextran sulfate 5000 Da (DS) enhances the clearance of EGFP-12xQ/N aggregates…………………………………………………………………...............86 F igure 15. DS enhances the clearance of EGFP-12xQ/N aggregates…………….. 85 F igure 16. DS enhances the clearance of EGFP-12xQ/N aggregates through the autophagy……………………………………………………………………………87 F igure 17. Autophagy inhibition with siRNA against Atg7 counteracts the clearance action of DS…………………………………………………………………………89 F igure 18. EGFP-12xQ/N aggregates colocalize with autophagy marker LC3 in stable cell line. …………………………………………………………………… 90

13

F igure 19. EGFP-12xQ/N aggregates may entrap endogenous TDP-43 and flag-TDP-43 wild type…………………………………………………………………95 F igure 20. EGFP-12xQ/N aggregation model is able to produce truncation products of TDP-43…………………………………………………………………………94 F igure 21. DS decreases the capture of flag-TDP43 wild type in the insoluble fraction……………………………………………………………………………..95 F igure 22. DS enhances EGFP-12xQ/N aggregates clearance and decreases TDP-43 wild type retention in the cytoplasm………………………………………………97 F igure 23. DS is decreasing acetylated tubulin level in HEK293 cells forming EGFP-12xQ/N aggregates…………………………………………………………99 F igure 24. DS action on EGFP-12xQ/N clearance is HDAC6 dependent……….101 F igure 25. EGFP-12xQ/N aggregates are co-localizing with autophagosome marker LC3………………………………………………………………………………..102 F igure 26. Dextran sulfate 5000 Da is not toxic for the neuronal-like cell line NSC-34…………………………………………………………………………………..104 Tables Table 1.Some genes and loci for familial ALS ……………………………………40 Table 2.Chemical structure and mechanism of action of compounds reported to treat ALS disease ………………………………………………………………………..53

14

1. IN T R O DU C T I O N

1.1. Protein degradation

Eukaryotic cells are continuously synthesizing proteins in order to endure, to

survive, and to maintain proper homeostasis. In physiological conditions, cells need

to eliminate misfolded/unfolded proteins. However, an imbalance between synthesis

and mis/unfolded proteins degradation is frequently observed. Cells accumulating

unfolded proteins have an impaired internal homeostasis that leads subsequently to

their death. Thus, it is essential to identify the mechanisms behind protein formation

and degradation, in order to propose a specific therapeutic strategy to target disease

which is characterized by an abnormal accumulation of mis/unfolded proteins.

Since early 1898, Hahn and colleagues have described cellular proteolytic activities,

using yeast(Hahn 1898). Later on, De Duve and Novikoff were able to obtain the

first electron micrographs of the partially purified hepatic lysosomes and have shown

the localization of acid phosphatase activity in these organelles(Essner E. &

Novikoff 1961).In the following years, researchers have studied different types of

cells using electron microscope and discovered a wide variety of vesicles involved in

the proteolytic pathway of the cell. As some of the vesicles contained engulfed

cytoplasmic material, it was suggested that these particular vesicles were pre-

lysosomes. These pre-lysosomes were found to be formed de novo in the cytoplasm

from a cup-shaped membrane called a phagophore. The extremities of the

phagophore expand while becoming spherical until they seal, enclosing the engulfed

pieces of cytoplasm with whatever might lie inside, and giving rise to a double-

membrane vesicle. Farquhar, for the first time, observed these closed vesicles, which

are later on called autophagosomes. Autophagosomes take up damaged molecules or

organelles and carry this cargo to the lysosomes. When De Duve observed

autophagosomes, he realized that cells could degrade their own components and

named the process "autophagy" (Smith & Farquhar 1966; De Duve Christian 1965).

Few years later, it was shown that the degradation of the proteins can happen not

only inside lysosomes, but also outside these structures in an energy-dependent

manner through the proteasomes(Coux et al. 1996; Hilt & Wolf 1996; Hochstrasser

1995; Peters 1994; Rubin & Finley 1995).

15

Autophagy seems to be active in the degradation of long-lived proteins and was in

the beginning identified as a bulk degradation mechanism for protein turnover during

periods of nutrient starvation (Klionsky & Emr 2000). In contrast to autophagy,

degradation depending on proteasome removes selectively ubiquitinated, and

aberrant short-lived proteins(Heinemeyer et al. 1991), thus it is denoted as the

ubiquitin-proteasome system (UPS). Since many of these short-lived proteins have

essential cellular regulatory roles; UPS plays a major role in different cellular

activities such as in the regulation of the cell cycle, and gene expression, in response

to variant cellular stresses, and in apoptosis (Attaix et al. 2001; Hershko et al. 2000).

Nevertheless, both UPS and autophagy are involved in the removal of misfolded

protein that fails to be refolded by chaperones, as well as different cytoplasmic

cargoes under various conditions (Figure 1).

1.1.1 UPS pathway:

The UPS machinery is composed of the proteasome and the enzymatic cascade,

through which the targeted ubiquitinated substrates are catalysed for the degradation.

First of all, the unfolded protein is tagged for proteasomal degradation with a chain

of 4 or more ubiquitin units by ubiquitin ligases. Ubiquitin (Ub) is a highly

conserved protein of 76 amino acids that is covalently linked to lysine residue(s) of

the targeted protein. Further on, these units form polyubiquitin chains that are bound

to each other through seven internal lysines (K6, K11, K27, K29, K33, K48, and

K63). It was reported that degradation by the UPS depends mainly on K48-linked

polyubiquitination of the targeted misfolded substrate(Komander 2009). Although

K48 represent the canonical proteasomal degradation tag, yet some substrates tagged

with K63-linked polyubiquitin can be also degraded through the

proteasome(Komander 2009). Next, these tagged substrates are delivered to the

proteasome where they will be degraded to short peptides and reusable ubiquitin.

The proteasome that is exclusively used in mammals is the cytosolic 26S

proteasome. It is a barrel-shaped structure with a molecular weight of about 2000

kDa, containing one 20S proteolytic core particle and two 19S regulatory particles.

26S proteasome structure has two openings at the ends where the target protein is

allowed to enter. 19S regulatory particle contains at least 18 subunits with the

16

multiple ATPase active sites and ubiquitin binding sites. This structure is responsible

for the recognition of the polyubiquitinated proteins, their unfolding and

translocation into the catalytic core. Catalytic core is composed of four rings that

form a central pore. The two inner rings are made of seven β subunitsthat contain

three to seven protease active sites. The target protein must enter the central pore

through these sites to be degraded. The two other outer rings have seven α

subunitswhose function is to maintain the pore entry through which proteins enter the

barrel (Komander 2009).

Autophagy (discussed fully in the following chapter) and UPS are both critical in the

maintenance of cellular homeostasis suggesting that their activity should be highly

orchestrated. Nevertheless, it was reported a complex and often an unexpected

interplay between these two cellular waste conveyors.

The narrow size of the proteasomal catalytic pore suggests that protein substrates

need to be partially-unfolded prior to their entry into the proteasome. Thus, protein

complexes and aggregates that have bigger size are reported to be poor proteasome

substrates (Nandi et al. 2006).In contrast to the UPS, macroautophagy is capable of

degrading a much wider spectrum of substrates, which, on average, tend to be

longer-lived and bulkier. These substrates include misfolded soluble proteins, protein

complexes, and aggregates. Ubiquitination appears to be a universal tag targeting

substrates for destruction via both catabolic systems, yet the exact type of

modification recognised by each pathway appears to be different. While K48-linked

polyubiquitin chains are employed by the UPS, substrates recognised by

autophagosome-lysosome pathway are thought to be modified either by K63-linked

chains (adopting a more open conformation than K48 chains), or might be just

monoubiquitylated (Welchman et al. 2005).

Several studies have focused on how changes in the activity of one of the degradative

pathways affect the flux through the other system (Ding et al. 2007).

Impairment of the UPS leads to an increase of the autophagic function.

Autophagy here is considered as a compensatory mechanism, since it is

allowing cells to reduce the accumulated UPS substrates. Indeed, in cells and

17

mice, where proteasome activity was inhibited using lactacystin, it was

shown that treating these cells and mice with rapamycin (inducer of

autophagy) tends to protect them against cell death by reducing the

accumulated UPS substrates (Pan et al. 2008). It was also shown that

upregulation of autophagy has a protective role when the UPS is impaired in

Drosophila (Pandey et al. 2007). Moreover, this protective effect is reported

to be HDAC6 (histone deacetylase 6) dependent (Pandey et al. 2007; Iwata et

al. 2005). However, the role of HDAC6 in this process is to ensure an

efficient delivery of autophagic machinery substrates for degradation.

HDAC6 was earlier found to regulate the formation of ubiquitinated inclusion

bodies, called aggresomes (see further in aggrephagy session) (Kawaguchi et

al. 2003). These aggresomes has been hypothesised to be formed in order to

be degraded more efficiently by autophagy (Iwata et al. 2005). These

molecular mechanisms may not be mutually exclusive and may be of

different importance in different cell types after the proteasome inhibition.

Thus more studies have to be done to identify different players connecting

two cellular proteolytic pathways.

On the other hand, when autophagy is inactivated in vivo and in vitro by the

knockout of essential autophagic genes (Atg5 or Atg7), significant

accumulation and aggregation of ubiquitinated proteins was reported (Hara et

al. 2006; Komatsu et al. 2006).Autophagy pathway clearance impairment

leads to the dysfunctional UPS. Indeed, several studies support this claim, as

it was shown that impaired autophagy also leads to the impaired degradation

of specific UPS substrates (Korolchuk, Mansilla, et al. 2009)

Decreased UPS flux in autophagy-compromised cells was not due to

impaired catalytic activity of proteasomes isolated from these cells, but due to

the accumulation of p62 protein that acts as an autophagy receptor and

connects ubiquitinated protein aggregates to the autophagic machinery.In

fact, when p62 was knocked down, the levels of UPS substrates in these

autophagy-deficient cells was rescued(Korolchuk, Menzies, et al. 2009).

Thus, lack of compensation for autophagy dysfunction by the UPS is in

agreement with the fact that p62, when accumulates, oligomerizes and is

therefore too bulky to be a good substrate for the proteasome with its narrow

18

catalytic pore. Yet, further studies are recommended to be done in order to

understand more, the molecular cross talk between the two main degradative

pathways.

F igure 1. Intracellular protein degradation via two main proteolytic systems: the ubiquitin-proteasome system (UPS) and the lysosomal-autophagy pathway. Delivery of cytoplasmic material to the lysosomes by autophagy can occur by three different pathways: (1) macroautophagy, which involves the sequestration of cytoplasmic components by a membrane forming an autophagosome, which fuses with the lysosome (2) microautophagy, which involves engulfment of small volumes of cytoplasm by a direct invagination of the lysosomal membrane (3) Chaperone-Mediated Autophagy (C M A), a process by which soluble substrates associated with a specific chaperone complex are translocated into the lysosome through the LAMP-2A lysosomal receptor. Proteins tagged with the polyubiquitin chain can be targeted by both the UPS and autophagy. Adapted from (Nedelsky et al. 2008).

19

1.1.2 Autophagy:

The term autophagy (originated from Greek word: auto phagin) means self-eating or

self-cannibalism. The term “autophagy” was first introduced by Christian De Duve

in mid-sixties to define the digestion of endogenous cellular material. Autophagy is

an evolutionarily conserved process through which the intracellular cytoplasmic

material (cargo) is delivered to lysosomes for degradation. Three major types of

autophagy in eukaryotes are known: chaperone-mediated autophagy (CMA),

microautophagy and macroautophagy (Figure 1).

Chaperone-mediated autophagy (C M A)Two main properties differentiate

chaperone-mediated autophagy (CMA) from the other types of autophagy in

mammalian cells: (1) the selectivity towards a particular pool of cytosolic proteins

and (2) the mechanism of delivery of the substrate proteins to the lysosomes. Mostly,

proteins bearing a targeting motif in their amino acid sequence, biochemically related

to the pentapeptide KFERQ, are selectively recognized by the Heat shock cognate

protein of 70 kDa (Hsc70), the chaperone that mediates their delivery to lysosomes

for degradation via CMA. It is estimated that about 30% of soluble cytosolic proteins

contain this CMA-targeting motif. After this targeting step, the substrate protein–

chaperone complex enclosures at the lysosomal membrane through interaction with

the cytosolic end the lysosome associated membrane protein type 2A (LAMP-2A),

which acts as a receptor for this autophagic pathway. Translocation of the substrate

across the lysosomal membrane also requires the presence of a luminal form of

Hsc70 (lysHsc70). After translocation, substrate proteins are rapidly degraded by the

abundant array of lysosomal hydrolases(Chiang et al. 1989; Cuervo 2010; Cuervo &

Dice 1996).

Microautophagy was first described in 1966 by De Duve and Wattiaux (De Duve

& Wattiaux 1966). Microautophagy is the non-selective lysosomal degradative

process. This process translocates cytoplasmic materials into the lysosome or

endosomes for degradation by direct invagination, protrusion, or septation of the

lysosomal or vacuolar membrane (Ahlberg & Glaumann 1985).With its constitutive

characteristics, microautophagy of soluble substrates can be induced by amino-acids

starvation or rapamycin via regulatory signaling complex pathways (Dubouloz et al.

2005). The maintenance of the organelles size, membrane homeostasis, and cell

survival under nutrients restriction are the main functions of microautophagy. In

20

addition, microautophagy is coordinated and complements macroautophagy and

chaperone-mediated autophagy (Dubouloz et al. 2005; Mijaljica et al. 2011). The

exact mechanism and molecular core machinery involved in the microautophagy in

mammalian cells is still unknown. Reviewed in Ref.(Li et al. 2012).

Macroautophagy, referred hereafter as autophagy, is the most studied form of

autophagy. Autophagy is a process of self-degradation of cellular components in

which double-membrane autophagosomes sequester organelles or portions of cytosol

and fuse with lysosomes or vacuoles for degradation. Autophagy is induced in

response to extra- or intracellular stress and signals such as starvation, growth factor

deprivation, ER stress, and pathogen infection. Nevertheless, autophagy can also be

observed at basal level in the cell, since cells have to maintain its integrity by

selectively removing unfolded proteins, damaged organelles, or invasive pathogens.

Cargo-specific names have been given to describe these various forms of selective

autophagy (e.g. mitophagy and pexophagy for the selective degradation of damaged

mitochondria and peroxisomes, respectively) (Klionsky et al. 2007; Mizushima et al.

2008).

F igure 2:Schematic representation of macroautophagy. Autophagosomes form from a pre-autophagic structure (phagophore) of unknown origin. The phagophore extends to envelop cytoplasmic constituents and damaged organelles. This structure elongates generating autophagosomes that are directed to the endocytic/lysosomal compartments that form autolysosome where engulfed constituents are degraded.

21

The most typical inducer of autophagy is nutrient starvation; in this sense, lack of

any type of essential nutrient can induce autophagy. In yeast, nitrogen starvation is

the most potent stimulus, but withdrawal of other essential factors such as carbon,

auxotrophic amino acids and nucleic acids, and even sulphate can induce autophagy.

In mammals, it is well known that serum starvation can induce autophagy in many

types of cultured cell. Amino acid and insulin/growth factor signals are thought to

turn on mTOR (mammalian target of rapamycin), which is a master regulator of

nutrient signalling (Yorimitsu & Klionsky 2005). Indeed, treatment with inhibitors of

mTOR such as rapamycin induces autophagy. In addition to insulin and amino acid

signaling, the involvement of many other factors in autophagy regulation has

recently been reported. These include Bcl-2, reactive oxygen species (ROS), AMP-

activated protein kinase (AMPK), and myo-inositol-1,4,5-triphosphate (IP3).

More than 30 different genes regulating autophagy (Atgs) have been identified in

yeast, and many of these have mammalian orthologs. These genes are involved at

key stages of the autophagy pathway: initiation, elongation, and maturation/ fusion

with the lysosomes (Figure 2)(Harding 1995; Thumm et al. 1994; Tsukada &

Ohsumi 1993).

A . Initiation of autophagy

Autophagy is initiated by the formation of an essential vesicle that is typical for

autophagy: the autophagosome. Membrane dynamics during autophagy are highly

conserved from yeast to plants and animals. Autophagy initiation starts by the

autophagosome formation, where cytoplasmic constituents, including organelles are

sequestered by a unique membrane called the phagophore or isolation membrane.

Where and how autophagosomes emerge has been a major question. It has been

hypothesized that autophagosomes can either be generated de novo from pre-existing

intracellular precursor molecules, or could arise from other intracellular

membranestructures like the endoplasmic reticulum (ER). The latter hypothesis has

recently been supported by more evidence suggesting that ER could contribute to

autophagosome formation, as an accumulation of vesicles associated with the ER

22

outer surface was seen in cells that had a block in maturation of pre-autophagosomal

structures. Thus, ER may be one of the membrane sources contributing to the

formation of pre-autophagosomal structures, so-called phagophores.The formation of

new autophagosomes requires the activity of (1) the class III phosphatidylinositol 3-

kinase complex (PI3K) (also known as Vps34/Beclin1 complex), and (2) FIP200-

ULK1/Atg1 complex (Figure 3).

Vps34 is the only identified PI3K in yeast so far, and its essential role in vacuolar

protein delivery was initially described through yeast genetics studies (Herman &

Emr 1990). The essential role of Vps34 in autophagy has been established largely

through the use of the pharmacological inhibitors wortmannin and 3-methyladenine

(3-MA), which have been used to suppress autophagy in many studies.Moreover, it

was shown that inhibition of Vps34 activity with wortmannin leads to

autophagosome formation inhibition. Furthermore, Vps34 is activated upon its

interaction with Beclin 1 that is the orthologus gene of Atg6 in yeast and is one of the

main pro-autophagic genes. In fact, if Vps34/Beclin1 interaction is disrupted, the

autophagosome formation is impaired. Upon Vps34 and Beclin1 interaction, Vps34

phosphorylates a phosphatidylinositol. Phosphatidylinositol-3-phosphate (PI-3-P) is

the major product of this activity. PI-3-P plays an essential role in the early stages of

the autophagy pathway. Recent studies have identified strong colocalization of early

autophagosome markers in PI-3-P-enriched structures that were formed upon

starvation(Kim et al. 2013).

A second macromolecular complex associated with the initiation step of

autophagosome formation is the FIP200 -ULK1/Atg1 complex. Upon starvation,

Atg13 binds to the mammalian Atg1 homolog ULK1 and mediates their interaction

with FIP200, thereby forming a ULK1-Atg13-FIP200 stable complex that triggers

the autophagic response downstream of mTOR. Under nutrient-rich conditions,

mTORC1 suppresses autophagy through direct interaction with this complex and

mediates phosphorylation-dependent inhibition of the kinase activities of Atg13 and

ULK1. Under starvation conditions or rapamycin treatment, mTOR dissociates from

the complex, resulting in the inhibition of mTOR-mediated phosphorylation of Atg13

and ULK1. This leads to dephosphorylation-dependent activation of ULK1 and

23

ULK1-mediated phosphorylations of Atg13, FIP200, and ULK1 itself, which triggers

autophagy. Therefore, the ULK1-Atg13-FIP200 complex acts as an integrator of the

autophagy signals downstream of mTORC1 (Hosokawa et al. 2009). However, it is

not clear yet how phosphorylation of these proteins regulates their activities (Figure

3).

F igure 3. Molecular pathway of autophagy initiation:the first step of the

autophagy pathway is its initiation (shown inside the frame) the initiation step starts

when Vps34 interacts with Beclin 1, in one hand and with FIP200-ULK1 complex,

on the other hand initiating the phagophore formation, once sensing the presence of

autophagic cargos in the cytoplasm.

B . E longation

The elongation of the membrane sac called: isolation membrane involved many Atg

proteins. Atg9 is one of the proteins that plays an essential role in phagophore

elongation. Atg9 is a transmembrane protein that cycles between the trans-Golgi

network and endosomes, probably carrying membrane for expansion of phagophore

24

(Figure 4). Furthermore, it has been reported that this process requires different other

Atgs, that are involved and grouped in two ubiquitin-like reactions (Figure 5).

(1) In the first cascade of the reactions, the ubiquitin-like protein Atg12 is covalently

tagged to Atg5. Atg12 is first activated by Atg7 (E1 ubiquitin activating enzyme-

like) and then transferred to Atg10 (E2 ubiquitin conjugating enzyme-like). Atg12 is

then linked by its C-terminal glycine to an internal lysine residue of Atg5. The

Atg12-Atg5 later on forms a conjugate with Atg16L1 (Atg12-Atg5-Atg16L1),

resulting in an 800-kDa complex tetramers. This complex is essential for the

elongation of the pre-autophagosomal membrane, but it dissociates once the fully

autophagosomes are formed (Mizushima et al. 1998).

(2) The second cascade ubiquitin-like reaction involves the protein microtubule-

associated protein 1 light chain 3 (MAP1-LC3/LC3/Atg8). LC3 is synthesized as a

precursor form and is cleaved at its C- terminus by the protease Atg4B, resulting in

the cytosolic isoform LC3-I. LC3-I is conjugated to PhosphatidylEthanolamine (PE)

in a reaction involving Atg7 (E1-like) and Atg3 (E2-like) to form LC3-II. LC3-II is

specifically targeted to the elongating autophagosome membrane and, unlike the

Atg12-Atg5-Atg16L1 complex, remains on completed autophagosomes until fusion

with the lysosomes. After fusion LC3-II on the cytoplasmic face of autolysosomes

can be delipidated by Atg4 and recycled. In addition, LC3 is also found on the

internal surface of autophagosomes that is degraded in the autolysosomes. The

association of LC3-II with autophagosomes makes it a specific marker for studying

autophagy. Upon vesicle completion and engulfing the sequestered substrates,

autophagosomes are delivered to lysosomes in the maturation step.

25

F igure 5. A tg12 and Atg8/L C3 ubiquitin-like conjugation pathways

required for autophagosome formation. Atg4 encodes a cysteine protease that

cleaves Atg8/LC3. Atg7 is similar to an E1-like protein, and Atg10 and Atg3

encode E2-like proteins. Atg5, Atg12 and Atg16 are physically associated with

the isolation membrane, whereas Atg8/LC3 is directly conjugated to the lipid

PhosphatidylEthanolamine (PE) that is inserted in the isolation membrane.

Adapted from (Geng & Klionsky 2008).

F igure 4.Autophagy elongation:Atg9 exists already on the surface of circulating

vesicles inside cytoplasm, once the pre-autophagosomal structure is formed, Atg9

vesicles fuse with these structures, and subsequently Atg9 becomes integrated into

the outer autophagosomal membrane. The autophagosomal membrane elongation

requires also Atg8 (LC3) and the complex formed by Atg16/Atg5/Atg12.

26

C . Maturation and fusion

Nascent autophagosomes undergo a stepwise maturation, resulting in the creation of

amphisomes and autolysosomes by fusion with multiple endocytic compartments,

such as early endosomes, multivesicular bodies (MVBs), late endosomes, and

lysosome (Tooze et al. 1990). Amphisomes, an intermediate hybrid vesicular

compartment, contain both autophagosomal and endosomal contents, while

autolysosomes are formed either from amphisomes or directly from autophagosomes

by fusion with lysosomes.These degrading structures are generally called

“autolysosomes” or “autophagolysosomes.”

While in yeast the autophagosome fuses directly with the vacuole, in higher

eukaryotes the process seems to be more complicated. Endocytosis is a process

involving a simple invagination of plasma membrane when the extracellular material

is internalized, whereas, the fusion step of autophagy involves proteins such as

endosomal sorting complex ESCRT, Rab7 and UVRAG/VPS34 protein (Figure 5).

ESCRTs are necessary for formation of multivesicular bodies (MVBs) and they are

indeed involved in sorting ubiquitylated membrane proteins into multivesicular

bodies (Simonsen & Tooze 2009). UVRAG, a Beclin 1 interacting protein, is also

involved in the maturation step by regulating positively the maturation of both

autophagosomes and endosomes. This function is independent of its interaction with

Beclin 1 (an ortholog of Atg6 and is a positive regulator of autophagy). UVRAG

recruits the VPS34 and via this interaction activates Rab7. Rab7 is found to be

important in the maturation of the autophagosome (Gutierrez et al. 2004; Stein et al.

2005). It is involved in vesicle transport from early endosomes to late

endosomes/lysosomes as well as in the fusion between autophagosomes and

lysosomes. Moreover, Rab7 is associated with mature autophagosomes and

autolysosomes, as well as with LC3, which preferentially labels immature

autophagosomes. Indeed, Rab7 is not essential for the initial step of autophagosome

maturation, but is involved in the final step of the maturation of late autophagic

vacuoles, possibly in the fusion with lysosome.

In mammalian cells, autophagosomes/lysosomes fusion is facilitated by microtubules

and appears to necessitate dynein. While, in yeast autophagosome/vacuole fusion

seems not to require these structures (Aplin et al. 1992; Fass et al. 2006; Fengsrud et

27

al. 1995; Kirisako et al. 1999; Köchl et al. 2006; Punnonen & Reunanen 1990;

Ravikumar et al. 2005; Webb et al. 2004). In mammalian cells, the destabilization of

microtubules by either vinblastin or nocodazole blocks the maturation of

autophagosomes, whereas their stabilization by taxol increases the fusion between

autophagic vacuoles and lysosomes. Autophagosomes move bidirectionally along

microtubules. Their centripetal movement is dependent on the dynein motor.

Dyneins are the motors that move intracellular cargos (including organelles and

vesicles) along microtubule tracks. It was suggested that dyneins could simply act by

moving autophagosomes to perinuclear locations where the lysosomes are

concentrated. Yet, the exact mechanism by which dyneins facilitate

autophagosomes/lysosomes fusion require further studies.

Another molecule recently added to the core machinery of maturation step is histone

deacetylase-6 (HDAC6), an ubiquitin-binding deacetylase, which selectively targets

the ubiquitinated proteins to autophagosomes. It has been shown that knockdown of

HDAC6 leads to the accumulation of autophagosomes, suggesting that HDAC6

controls the autophagosomematuration rather than theautophagosome formation.

Moreover, it has been demonstrated that HDAC6 role in this step is to control the

actin cytoskeleton (Lee et al. 2010). Despite of such recent progress, molecular

mechanisms underlying coordinated regulation of multiple maturation steps by these

factors are still incompletely understood. The maturation step is ended by lysosomal

degradation of cargos. The autophagosome content is degraded inside autolysosome

structures that contain different hydrolytic enzymes, such as proteases, lipases, and

nucleases that are capable of breaking down all types of biological polymers.

Here, it is also essential to mention that lysosomal positioning is dynamically

regulated by nutritional conditions, hence by autophagy. Starvation induces

preferential re-localization of lysosomes from cell peripheries to juxtanuclear regions

close to the microtubule-organizing center (MTOC), thus regulating the autophagic

flux in cells (Maday et al. 2012). It is important to mention that in neurons,

bidirectional movements of lysosomes within axons are observed, while

autophagosomes and endosomes are also bidirectionally moved mainly in distal

axons. They are then transported exclusively in retrograde direction after fusion with

lysosomes to the cell body of the neuron for complete degradation of their cargos.

28

Taken together, lysosomal degradation of engulfed cargos in neurons could be

strictly dependent on retrograde transport and late autophagosome/lysosomal

trafficking (Katsumata et al. 2010).

F igure 6. Molecules involved in the maturation step of autophagy:

UVRAG interacts independently with Vps 34 and ESCRT complex to promote

autophagosome maturation that is later on fused either with endosome or lysosome to

generate autolysosome, UVRAG-Vps34 is also involved in the endosome–lysosome

transition by activation of Rab7. Adapted from (Liang et al. 2008).

1.1.2.1 Physiological role of Autophagy:

To understand the various roles of autophagy, it may be useful to subclassify

macroautophagy into “induced autophagy” and “basal autophagy” (Mizushima

2005). The former is used to produce amino acids following starvation, while the

latter is important for constitutive turnover of cytosolic components. However, this

distinction is too simplified and cannot be applied to more complicated issues.

Autophagy has a greater variety of physiological roles than expected, such as

starvation adaptation, intracellular protein and organelle clearance, development,

anti-aging, elimination of microorganisms, cell death, and the antigen presentation.

The physiological roles of autophagy are based on the respective processes involved

29

in, such as usage of degradation products, elimination of macromolecules and

organelles, and sequestration/packing.

A-U tilization of degradation products:

Under normal conditions and during very short periods of starvation, maintenance of

the amino acid pool seems to rely primarily on the ubiquitin–proteasome system

rather than autophagy (Vabulas & Hartl 2005). However, during starvation that

persists for several hours, necessary amino acids are produced by autophagy, which

is up-regulated as an adaptive response. It is important to emphasize that excess

production of amino acids by autophagy is an acute response or emergency action.

Therefore, induction of autophagy can support cell survival only for a short time. For

example, during cell growth, autophagy is activated at initial stages, but returns to

basal levels after a normal conditions are stablished (Degenhardt et al. 2006). In

contrast, little is known about how useful autophagy is in overcoming chronic

starvation.

B-Elimination of macromolecules and organelles:

The second purpose of autophagy is the elimination of cytoplasmic contents.

Although this role has been thought to be the specialty of the ubiquitin–proteasome

system, many recent studies have shown that autophagy also participates in

intracellular clearance or protein/organelle quality control. The most direct evidence

is the accumulation of abnormal proteins and organelles in autophagy-deficient

hepatocytes, neurons, and cardiomyocytes even in the absence of any

disease(Komatsu et al. 2005; Hara et al. 2006; Nakai et al. 2007), ubiquitin-positive

inclusion bodies, and deformed organelles accumulate in these cells. Since induced

autophagy is not observed in the brain during starvation, low levels of basal

autophagy are likely sufficient for quality control.Some types of induced autophagy

are aimed at the elimination of excess or unneeded organelles. For example,

peroxisomes induced by metabolic demand are selectively degraded primarily by

microautophagy(Sakai et al. 1998).The elimination of cytoplasmic contents by

autophagy is so important that defects cause various cellular malfunctions. One of

the possible outcome of autophagy defects is neurodegeneration. Indeed, The

accumulation of autophagic vacuoles has been observed in many human

30

neurodegenerative diseases, including Alzheimer’s disease for instance (Okamoto et

al. 1991). It remains largely unknown whether these represent up-regulation of

autophagy or blockage of autophagic flux.

C-Sequestration/packing:

In some cases, sequestration in autophagic membranes, even without degradation,

seems to be important to exert special functions. Autophagy can be induced by

several stresses, including ER stress. ER stress-induced autophagy is basically

protective against cell death in both yeast and mammals(Bernales et al. 2006; Ogata

et al. 2006). How autophagy protects cells during ER stress is not exactly known, but

it was suggested that sequestration of ER into autophagosomes might be sufficient to

relieve ER stress(Bernales et al. 2006).Moreover, autophagy has been suggested in

the context of neurodegenerative disease. Autophagy likely has a beneficial role in

the clearance of misfolded or other harmful proteins. However, if autophagic

degradation is not rapid enough, sequestration of cytoplasm might rather have an

adverse effect. It was proposed that autophagosome maturation into autolysosomes is

impaired in Alzheimer’s disease brains for instance(Yu et al. 2005).

In fact, defective physiological autophagic role plays a significant effect in human

pathologies, neurodegeneration diseases, for instance.

1.1.2.2. Selective autophagy

Although autophagy is historically described as a non-selective bulk protein

degradation system, recent findings confirm that it can be also a highly selective

process. Selective autophagy is mediated by both autophagy receptors and adaptor

proteins that link the cargo with the core autophagic machinery. The term selective

autophagy refers to the selective degradation of organelles, bacteria, ribosomes,

specific proteins, and protein aggregates by autophagy. When autophagy is

selectively recognizing a specific cargo, its name is attributed according to the

targeted cargo such as aggrephagy (aberrant protein aggregates), mitophagy

(damaged mitochondria), reticulophagy (ER) and xenophagy (invasive pathogenes)

(Klionsky et al. 2007). Selective autophagy functions as a quality control system, but

the signals involved in recognition of selective cargo for autophagy degradation are

31

still poorly understood. Autophagy receptors, such as p62 and NBR1 are the main

proteins involved in the selective autophagy of protein aggregates, so-called

aggrephagy.

Furthermore, autophagy receptors could also bindto other specific adaptors, which

function as scaffolding proteins that selectively bring the cargo-receptor complex in

contact with the core autophagic machinery to allow sequestration of the substrate. In

addition to the autophagy receptors and specific adaptors, such as Atg19, Atg34, p62,

NBR1, NDP52 (nuclear dot protein 52 kDa for bacteria recognition), OPTN

(optineurin), Nix (NIP3-like protein X), and ATG32, selective autophagy in general

relies on the same molecular core machinery of non-specific autophagy.

The protein aggregates represent intermediates in autophagic degradation of

aggregation-prone proteins. The assembly of autophagy substrates into larger

aggregates or clustered structures is a common feature of selective autophagy. Their

assembly may facilitate uptake into autophagosomes, and aggregates may work as

nucleation sites for the phagophore. Damaged proteins are recognized and sorted by

chaperone and co-chaperone complexes containing chaperone-assisted ubiquitin E3

ligases to three different degradation pathways: UPS, CMA, and/or aggrephagy. In

the following section, current knowledge about aggrephagy will be discussed.

1.1.2.3. Aggrephagy

To accomplish their normal cellular function, proteins have to be correctly folded. In

fact, the accumulation of misfolded or unfolded proteins into protein aggregates is a

pathological trait of many clinical disorders. The term aggrephagy was introduced

for the first time by Per O. Seglen to describe the selective sequestration of protein

aggregates by autophagy(Øverbye et al. 2007). Generally, protein aggregation is

caused by an abnormal protein conformation, leading to the formation of oligomeric

intermediates, and further on to the small protein aggregates(Merlini et al. 2001).

These small protein aggregates can again form variety of structures (Dobson 2003),

termed as intracellular inclusions bodies (Grune et al. 2004; Kopito 2000). Larger

cytoplasmic inclusions can develop further and fuse to form an aggresome, a

32

pericentriolar, membrane-free cytoplasmic inclusion formed specifically at

themicrotubule organizing center (MTOC) containing misfolded, and ubiquitinated

proteins (Johnston et al. 1998; Kopito 2000).

Aggresome is thought to be a protective cellular structure, since it is made by

sequestered misfolded proteins that cannot be degraded by the proteasome and that

could be toxic circulating in the cytoplasm (Johnston et al. 1998; Kopito 2000). On

this way, the misfolded proteins inside the aggresome are straightly transported

through autophagy for degradation. Protein aggregates are able to be sequestrated

inside the cell because of a variety of cellular stressors, such as mis/unfolded protein

formation, impaired proteasomes, oxidative stress, or aging (Kopito 2000).

Misfolded proteins generally become poly-ubiquitinated. Such proteins are normally

degraded by the UPS, while aggregate-prone proteins are poor substrates for

proteasomal degradation as they are highly insoluble and too big to pass through the

narrow proteasome pore (Stefanis et al. 2001; Verhoef et al. 2002). It has been also

shown that the ubiqutin linked chain K48 is a classical signal for UPS dependent

degradation and it has been suggested that autophagic substrates present a modified

K63-linked ubiquitin chains (Tan et al. 2008). Moreover, it has been demonstrated

that the autophagy receptors p62 and NBR1 recognize K63-linked ubiquitin

chains(Kirkin et al. 2009; Long et al. 2008; Wooten et al. 2008) and inclusions

containing K63-linked ubiquitin chains and target them for autophagic degradation

(Tan et al. 2008).

p62/A170/ SQSTM1 (hereafter referred to as ‘p62’) and NBR1 (neighbor of BRCA1

gene 1) are the components of the ubiquitin-positive inclusion bodies found in some

neurodegenerative and liver diseases. These molecules have unique features, N-

terminal Phox and Bem1p domain (PB1), which has the ability in self-

oligomerization, C-terminal ubiquitin associated domain (UBA) capable of

interaction with ubiquitinated proteins and LC3-interacting region (LIR) responsible

for the interaction with ATG8 family proteins. Both p62 and NBR1 mediate

aggregate formation, also called p62 bodies, sequestosomes or aggresome-like

inducible structures (ALIS) (Bjørkøy et al. 2005; Szeto et al. 2006; Clausen et al.

2010)and thus connect ubiquitinated protein aggregates to the autophagic machinery

players(Bjørkøy et al. 2005; Lelouard et al. 2002; Szeto et al. 2006) (Figure 7).

33

It was shown through two main studies that p62 is critical for protein aggregates

formation and their clearance by autophagy. In the first study, in p62 deficient

models, it was observed that the aggresome-like inclusion bodies formation is

notably impaired (Pankiv et al. 2007; Clausen et al. 2010). In accordance with this,

using Atg7 knock-out (KO) mice or Atg8 mutant flies, it was confirmed that large

ubiquitin-positive protein aggregates accumulate in these mutated models, and no

longer persist when P62 is impaired(Komatsu et al. 2007; Nezis et al. 2008).

Accordingly, over-expression of p62 leads to accumulation of ubiquitinated protein

aggregates(Bjørkøy et al. 2005; Seibenhener et al. 2004). In the second study, using

electronic microscopy p62 was localized inside double membrane vesicles (Bjørkøy

et al. 2005). It is important to point out that p62 was also identified as one of the

specific substrates that are degraded through the autophagy–lysosomal pathway.

Recently, the ubiquitin-binding histone deacetylase 6 (HDAC6) is reported to be a

key player in recruiting ubiquitinated, misfolded proteins to the aggresome (Iwata et

al. 2005; Kawaguchi et al. 2003; Olzmann et al. 2007). While most members of the

histone deacetylase (HDAC) family are localized in the nucleus, HDAC6 is found

exclusively inside cytoplasm and it has been shown to contain an ubiquitin-binding

domain (BUZ finger). As it has been demonstrated that HDAC6 may bind to the

microtubule motor protein dynein as well, it was proposed that HDAC6 can facilitate

transport of aggregates to the MTOC in order to form the aggresome (Kawaguchi et

al. 2003). In HDAC6 deficient cells dispersed micro-aggregates were formed inside

the cytoplasm, while formation of aggresomes was impaired suggesting a failure in

aggregates transport to the MTOC region (Kawaguchi et al. 2003).

34

F igure 7. Aggrephagy main players. P62 and NBR1 mediate aggresome/inclusion

body formation and are substrates prone to autophagosome engulfement and

degradation inside the autolysosome. In addition, aggresome formation is

coordinated by HDAC6 and is directly carried to lysosome via microtubules (e.g:

dynein), adopted from(Kraft et al. 2010).

1.2. Autophagy and proteinopathies Proteinopathies belong to conformational diseases that are caused by the failure of a

protein to attain or maintain correct conformation which is usually due to genetic

mutations. With aging, an increase in intracellular oxidative molecules leads to the

accumulation of unfolded proteins in cells, too. Misfolded proteins are engaged in

inappropriate interactions with other cellular components and can accumulate in

potentially toxic protein inclusions(Lashuel & Lansbury 2006). A number of studies

have identified small oligomeric aggregates as species linked to toxicity(Kayed et al.

2003). For example, in mammalian cell culture and yeast, accumulation of small

oligomeric Huntingtin aggregates has shown to correlate with toxicity(Kitamura et

al. 2006; Finkbeiner et al. 2006).

On the other hand, some studies have shown that soluble oligomers coming out from

proteasomal degradation of misfolded proteins in neurons could be more toxic than

their sequestration into aggregates. So, formation of aggregates might prevent their

toxicity awaiting for their transport to lysosomes for a complete degradation via

autophagy(Arrasate et al. 2004; Szeto et al. 2006; Tanaka et al. 2004). In the same

context, it has been suggested that formation of insoluble amyloid inclusions can also

promote cell survival and may serve as a protective mechanism by sequestering

potentially harmful aggregates from the cytosol. Several studies suggest that the

formation of tightly packed Huntingtin deposits is beneficial for cell survival. When

Greenberg and colleagues have transfected with mutant Huntingtin primary striatal

neuron, they could induce the formation of inclusions. These inclusions resembled

the protein deposits found in the brains of Huntington patients, as they were

intranuclear and ubiquitinylated. But these inclusions were not sufficient to induce

apoptosis(Saudou et al. 1998). Incorporation of toxic oligomers into protective

amyloid-like protein inclusions has been observed to reduce toxicity of Huntingtin

35

expressed in mammalian cells and mouse models(Cohen et al. 2009; Cheng et al.

2007). These studies support the hypothesis that sequestration into an

aggresomeremoves toxic misfolded species from the cellular environment.

Under normal conditions the cell can efficiently degrade misfolded/unfolded proteins

through targeted proteolysis. The UPS and autophagy are the two major intracellular

protein degradation pathways. While UPS is mediating degradation of most normal

proteins after performing their normal functions as well as the removal of the

abnormal soluble proteins, autophagy is mainly responsible for the degradation of

defective organelles and the bulk of aggregated proteins. UPS proteolytic function

often becomes inadequate in proteinopathies, which leads to activation of autophagy

for the removal of abnormal proteins especially the aggregated forms. Enhancement

of autophagy via pharmacological intervention has shown great promises in relieving

proteinopathies in the cell. Even though few different strategies are employed to

increase proteasome function, it seems that UPS is not as efficient as autophagy in

alleviating proteinopathies (Díaz-Hernández et al. 2003).

Proteinopathies are exemplified mainly by human neurodegenerative diseases such

as: Alzheimer’s disease (AD), Parkinson’s disease (PD), Huntington’s disease (HD)

and Amyotrophic Lateral Sclerosis (ALS) (Ravikumar 2002; Berger et al. 2006).

Mentioned disorders characterize intracellular protein aggregates that contain a bulk

of misfolded or abnormal aggregate-prone proteins, such as: β-amyloid protein(Aβ),

polyglutamine-rich repeats of Huntingtin protein, alpha synuclein, and SOD1 or

recently reported TDP-43 protein.

A direct link between autophagy and neurodegeneration has been established by loss

of basal autophagy in mouse brains through conditional knockout of key autophagy

genes, Atg5 or Atg7, which results in a neurodegeneration phenotype with the

accumulation of ubiquitinated protein aggregates (Hara et al. 2008; Komatsu et al.

2006). This suggests that autophagy is vital for cellular quality control in neurons for

the turnover of proteins and organelles (Rubinsztein 2006). Thus, cellular quality

control through the autophagy is essential for neuronal survival. Growing evidences

propose that autophagy has a protective role in neurodegenerative diseases, yet its

detailed mechanism is still unclear. Moreover, the basal level of autophagy is usually

36

low in healthy neurons, because of the very rapid turnover of autophagosomes

(Boland et al. 2008). Neurons might indeed be susceptible to the alterations in any of

autophagy steps. Indeed, abnormal number of autophagosomes and/or amphisomes

(fusion product of autophagosomes and late endosomes) indicating an impaired

autophagy, are observed in neurodegenerative diseases. The defect in autophagy

could be thus due to the deficiency of the endocytosis at the late endosome-lysosome

stage, the inhibition of autophagosome maturation, and/or lysosomal degradative

function impairment (Wong & Cuervo 2010). In fact, a decrease in autophagosome

maturation and autophagic flux turnover impairment explains the overdue

phenotypes often characteristic for many of the neurodegenerative proteinopathies.

On the other hand, it was shown that abundant autophagy vacuoles sequestration and

the increase of autophagy proteins such as p62were found in different experimental

models of motor neuron degeneration and in ALS (Laird et al. 2008; Sasaki 2011). In

fact, the increase in autophagy proteins is due to the impairment of autophagy

progression as a sort of non-effective compensatory mechanism. In the presence of a

defective autophagy progression, LC3 II levels increase and an excess of misfolded

proteins and/or altered organelles is observed.

Growing evidence are proposing that strengthening autophagy could be a beneficial

approach in neurodegenerative diseases treatment (Xilouri & Stefanis 2011). It is

widely suggested that induction of autophagosome formation would be a solution to

clear the aggregates specially at early stages of the disease (Xilouri & Stefanis 2011).

In contrast, other studies suggested that boosting autophagosome formation is not a

wise solution, especially if the maturation of autophagosomes to lysosomes is

interrupted. Therefore, an increase in the autophagic flux progression might be a

better way to delay the onset of neurodegenerative diseases.

Autophagy inducer, rapamycin (mTOR inhibitor), has shown neuroprotective effect

in several neurodegenerative disease models, including Huntington disease.

Rapamycin protects against mutant Huntingtin-induced neurodegeneration in cell, fly

and mouse models of Huntington's disease (Ravikumar et al. 2004). Spilman et al.,

have shown that long-term inhibition of mTOR by rapamycin prevented AD-like

cognitive deficits and lowered levels of Aβ42, a major toxic species in mouse

37

Alzheimer’s disease model (Spilman et al. 2010). In line with this, inParkinson

disease PC12 cell-based α-synuclein aggregation model, treatment with rapamycin

has shown increased clearance of these aggregates (Webb et al. 2003).

Despite its beneficial role in the aggregation clearance, the use of mTOR inhibitors

are considered not to be a prominent long term treatment in neurodegenerative

disease patients because mTOR has many vital cellular functions, like translation and

cell growth. Therefore, induction of autophagy independent of mTOR may provide a

more rational treatment approach. Further studies reported by Rubinzstein’s group

demonstrated the first mTOR-independent autophagy pathway in which autophagy

could be induced by other agents like lithium and carbamazepine (Sarkar et al. 2005).

The same group, via chemical screening, identified a number of small-molecule

enhancers of rapamycin effect (SMERs) that have been shown to induce autophagy

independent of mTOR and rapamycin. Sarkar et al., have shown using Huntingtin

aggregation model that these small molecules have beneficial effect by enhancing the

clearance of aggregate-prone proteins, and may therefore counteract the autophagy

dysfunction in proteinopathies (Sarkar et al. 2007).

1.3 Amyotrophic Lateral Sclerosis

Amyotrophic lateral sclerosis (ALS), known as Lou Gehrig‘s disease, is a

neurodegenerative motor neuron disease. Since 1869, the French neurologist Jean-

Marin Charcot (Rowland 2001), has diagnosed the first clinical trait of ALS and has

defined it as a progressive motor function disability due to motor neuron

degeneration in the primary motor cortex, brainstem and spinal cord (Shaw et al.

2001). The earliest symptoms of ALS are obvious muscle weakness and/or atrophy.

The parts of the body affected by these early symptoms depend on which specific

motor neuron populations in the body are damaged first. About 75% of ALS cases

have “limb onset ALS” which affects one of the legs or arms. Patients might feel

awkwardness while walking or running. The other 25% of cases show “bulbar onset

ALS”. These patients will have problems in speaking clearly, and consequently a

dramatic loss of tongue mobility. As the disease progresses, muscle weakness and

atrophy will quickly extend the rest of the body (Brooks 1994; H Mitsumoto 1997;

H. Mitsumoto 1997) with eventual paralysis and death within 2-5 years of clinical

onset (Rowland 1998). ALS disease progression leads ultimately to the death mainly

due to the enervation of respiratory muscles and subsequent respiratory failure. ALS

38

disease is settled in patients that have ages between 55-65 years (Haverkamp et al.

1995),(Leigh 2007) and its incidence is between 1 to 2 per 100 000 each year in

Europe and North America, for instance (Worms 2001).

Although ALS disease has been discovered over a hundred of years ago, up to now

there is no unambiguous test to clearly diagnose ALS since clinical symptoms are

frequent to several other neuromuscular and skeletal disorders. In order to assist the

diagnosis of ALS patients for research and drug trials, the World Federation of

Neurology (WFN), Research Group on Motor Neuron Disease developed diagnostic

criteria since early 1990s (Brooks 1994; Brooks et al. 2000). Through these criteria,

patients can be classified into the several categories: “Definite ALS”, “Probable

ALS”, “Probable ALS- Laboratory supported”, “Possible ALS” or “Suspected ALS”,

based on which regions of the bodyhave been degenerated. The development of these

diagnostic guidelines has been defined in order to facilitate the diagnosis and

therefore encouraging the advancement of our understanding of ALS and the

implementation of efficient treatment.

It is still not known the cause of ALS and its exact molecular core involved in motor

neuron degeneration. Although, the majority of ALS cases happen sporadically, it is

proposed that genetic background and environmental factors are both involved in

ALS development. Clear evidence to define the exact factors in sporadic cases has

still to be elucidated, yet the evidence for the influence of genetic contribution has

been suggested to be highly involved.

Since 1980s, our understanding of ALS physiopathology has significantly increased.

Because the progressive paralysis in ALS is due to motor neurons degeneration, early

thought about the pathogenesis suggested that motor neurons are dying

autonomously. Many mechanisms have been proposed to lead directly to motor

neurons degeneration such as: oxidative stress (Yim et al. 1996), abnormal protein

accumulation (Johnston et al. 2000; Watanabe et al. 2001), glutamate excitotoxicity

(Rothstein et al. 1992), dysfunctional mitochondria (Cozzolino et al. 2008); (Shaw

2005) and recently described, deficient RNA processing(Lagier-Tourenne et al.

39

2010). In addition, many mutations in multiple genes are associated with motor

neurons death (Valdmanis & Rouleau 2008).

1.3.1 G enetics of Amyotrophic Lateral Sclerosis

Even though ALS disease has common clinical feature due to motor neuron

degeneration, it remains a heterogeneous disease. More than 90% of ALS cases

occur sporadically (sALS), and the 5-10% of cases have a familial background

(fALS) caused by mutations in more than ten different genes (Valdmanis & Rouleau

2008). Interestingly, fALS cases are identical to sALS from a neuropathological

point of view, hence proposing a common pathogenesis feature (Bruijn et al. 2004).

Eleven disease-associated mutations in the superoxide dismutase (SOD1) gene were

discovered for the first time by Rosenet al in 1993 (Rosen et al. 1993). Since then,

many other genes has been shown to be affected in ALS patients, such as FUS/TLS,

OPTN (optineurin), and TDP-43 (Table 1). Despite an outstanding heterogeneity of

ALS, the presence of ubiquitin-positive inclusion bodies in the cytoplasm of defected

neurons of ALS patients seems to represent a common pathological feature

(Giordana et al. 2010). In 2006, a staggering study conducted by Neumann et

al.(Neumann et al. 2006), identified the TAR DNA binding protein 43 (TDP-43) as

the main component in ALS patient’s aggregated protein inclusions. Later studies

have shown that about 4% of fALS cases and up to 2% of sALS carry dominant

missense mutations in the TARDBP gene further suggesting that TDP-43 plays an

important role in neurodegeneration (Corrado et al. 2009; Daoud et al. 2009; Van

Deerlin et al. 2008). Therewith, other studies have shown mutations in another

RNA/DNA binding protein, known mainly as Fused in Sarcoma (FUS) or

Translocated in Liposarcoma (TLS). Alltogether, it is estimated that in around 1-2%

of fALS cases RNA processing and metabolism is affected (Kwiatkowski et al. 2009;

Vance et al. 2009).

Adapted from Habib and Mitsumoto(Habib & Mitsumoto 2011)

A LS Subtype Gene Onset Inheritance Locus Protein

A LS1 SOD1 Adult Dominant 21q22.1

Cu/Zn superoxide dismutase

A LS10 TDP-43 Adult Dominant 1p36.22

40

TAR DNA-binding protein A LS-F T D*1

Unknown Adult Dominant 9q21-22 -

A LS-F T D2 Unknown Adult Dominant 9p13.3-21.3 -

Undesignated OPTN Adult Dominant 10p15-15 Optineurin

Table 1. Some genes and loci for familial A LS*Frontotemporal Dementia

SO D1 gene

The Cu/Zn superoxide dismutase 1 (SOD1) is a gene encoding a 32kDa cytoplasmic

enzyme that is involved in the catabolism of toxic superoxide radicals (O2-) to

innocuous molecular oxygen (O2) and hydrogen peroxide (H2O2) (McCord &

Fridovich 1969). Moreover, SOD1 is ubiquitously expressed and highly conserved

among species. Since 1993, when the first SOD1 missense mutations were described

(Rosen et al. 1993), the number of mutations linked with fALS in this gene has been

amplified to more than 140. Mutations in SOD1 are described in about 20% of

familial cases and infrequently among sporadic ALS cases (Andersen et al. 2003).

Some reports based on mutant SOD1 mouse models have demonstrated that mutation

in SOD1 results in disease due to a toxic gain of function more than the loss of SOD1

activity (Bruijn et al. 2004). Mutant SOD1 proteins have been known to accumulate

abnormally in the form of insoluble inclusions within affected spinal motor neurons

of SOD1- related ALS patients (Bruijn et al. 1998).

F US

FUS/TLS or fused in sarcoma/translocated in liposarcoma is a 526 amino acid

protein encoded by 15 exons and characterized by an N-terminal domain enriched in

glutamine, glycine, serine and tyrosine residues (QGSY region), a glycine-rich

region, an RRM, multiple arginine/glycine/glycine (RGG) repeats in an arginine- and

glycine-rich region and a C-terminal zinc finger motif. FUS/TLS is predominantly

localized in nucleus (Andersson et al. 2008). Mutations in this RNA/DNA-binding

protein were mostly described in familial ALS (Kwiatkowski et al. 2009). Most of

these mutations are clustered in the glycine-rich region and in the extreme C-terminal

part of the protein. It was also reported, that the patient’s cells contain an abnormal,

41

high levels of ubiquitinated full-length FUS/TLS protein, as well as C-terminal

fragments in cytoplasmic insoluble fractions.

T DP-43

Human TDP-43 protein is 43-kDa protein composed of 414 amino acids and encoded

by the Tar DNA-binding protein gene (TARDBP) located on chromosome 1. TDP-

43 is a highly conserved, ubiquitously expressed, mainly a nuclear protein capable of

shuttling between the nucleus and cytoplasm (Ayala et al. 2008; Banks et al. 2008;

Winton et al. 2008). TDP-43 protein is composed of five functional domains: two

conserved RNA recognition motifs (RRM1 and RRM2), a nuclear localization signal

and a nuclear export sequence that mediate nuclear shuttling, in addition to a

carboxy-terminal glycine-rich domain implicated in TDP-43 protein interactions and

aggregation (Lagier-Tourenne et al. 2010) (Figure. 8). This structure places TDP-43

in the heterogeneous nuclear ribonucleoprotein protein family (hnRNP) (Dreyfuss et

al. 1993).

TDP-43 is able to bind single stranded DNA and RNA and has been shown to have a

role in repression of gene transcription (Ou et al. 1995), alternative splicing (Buratti

& Baralle 2001),(Lagier-Tourenne et al. 2010), and mRNA stability (Strong et al.

2007). Despite TDP-43 is an ubiquitously expressed protein, Northern analyses have

shown that mRNA expression levels of this protein are heterogeneous among

different tissues (Ou et al. 1995), which suggest that TDP-43 may have different

functions in different tissues. Increased expression levels of TDP-43 have been found

in the brain and spinal cord tissue of wild-type mice (Sephton et al. 2010), that can

explain why motor neurons located in these regions are specifically susceptible to

degeneration in ALS.

In 2006, Neumann et al. identified TDP-43 as a major component of ubiquitinated

inclusions in FTLD (Frontotemporal lateral dementia) and ALS patients. Wild-type

TDP-43 is found in aggregates in spinal cord motor neurons, hippocampal and

frontal cortex neurons and glial cells in the vast majority of ALS patients. In FTLD,

TDP-43 aggregates are present in the most common subtype of the disease with

ubiquitinated inclusions, referred to as FTLD-TDP.

42

In early 2008, successful identification of dominant mutations as a primary cause of

ALS provided evidence that aberrant TDP-43 can directly trigger neurodegeneration.

Therefore, understanding how the mutants cause neurodegeneration offers a

convenient entry point for exploring how TDP-43 plays this role. The first question

is whether a gain, or a loss of function mediates neurotoxicity. A resolution to this

question is of critical importance because it sets the direction of further research on

the disease mechanism and on the design of therapeutic strategies. To answer this

question, model systems of both gain or loss of function must be employed. Gain-of-

function models are usually achieved by gene overexpression, and loss-of-function

models by gene knockout or knockdown. The precise role and the exact mechanism

by which TDP-43 participates in ALS pathogenesis are still to be elucidated.

F igure 8.T DP-43 protein structure. TDP-43 has five functional regions including

two RNA recognition motifs (RRM1 and RRM2), a nuclear localization signal

(NLS) and nuclear export signal (NES) that mediates nucleo-cytoplasmic shuttling,

and a glycine rich region (GRR) that mediates protein-protein interactions. Below,

the prion-like Q/N rich region has been highlighted, in addition to the location of the

majority of ALS-associated mutations. A structural region mediating TDP-43

association with stress granules (SGs), as identified in two independent studies, also

spans part of the prion-like Q/N rich sequence adopted from(Dewey et al. 2012).

43

1.3.2 T DP-43 aggregation

Overview:

In the physiological conditions, TDP-43 is located mainly in the nucleus and it has a

role in the regulation of gene expression. In contrast, in the pathophysiological

conditions, TDP-43 is found located in cytoplasmic inclusions, where is hardly

soluble, hyperphosphorylated, ubiquitinated, and later on cleaved into small

fragments (Arai et al. 2006; Mackenzie 2007) demonstrating that the aberrant

subcellular localization is important for disease pathogenesis progress. In the

beginning, it was debated whether TDP-43 and its abnormal subcellular localization

and accumulation in inclusion bodies are directly involved in the development of

ALS. An increasing number of studies based on mutations in the TARDBP gene from

patients with fALS and sALS are suggesting that TDP-43 is directly involved in ALS

pathology. Up to now, more than 35 various mutations in the TARDBP gene have

been reported (Corrado et al. 2009; Daoud et al. 2009; Van Deerlin et al. 2008).

Interestingly, most of these mutations have been cited to be found in the C-terminal

glycine-rich domain of TDP-43 known to be involved in the interaction with other

hnRNP family proteins (Warraich et al. 2010; Lagier-Tourenne et al. 2010). These

findings propose that the mechanism of pathogenicity in ALS could be tightly related

to the interruption of the protein-protein interactions. However, this domain is also

the least conserved region of the protein and this low homology is thought to reflect

the differences in species-specific functions (Ayala et al. 2008).

TDP-43 is an intrinsically aggregation-prone protein capable of forming homodimers

in vitro(Shiina et al. 2010);(Arai et al. 2006). Several ALS-linked TDP-43 mutations

have been shown to increase its tendency to form aggregates(Johnson et al. 2009;

Kabashi et al. 2010). Protein extracts from affected brain andspinal cord has defined

a biochemical signature of disease that includes hyperphosphorylation and

ubiquitination of TDP-43, and the production of several C-terminal fragments

(CTFs) around 25 kDa and 35 kDa. These observations have leadinvestigators to

propose a toxic gain-of-function mechanism in the pathogenesis of ALS. The "gain

of function” model is based on potential aberrant action of TDP-43 when localized in

the cytoplasm, as an increased degradation can yield to toxic fragments and thus to

the potential toxic properties of the aggregates themselves. In fact, overexpression of

44

wild type TDP-43 has a toxic effect in yeast and cultured primary neurons derived

from mouse embryos, as well as in Drosophila (Li et al. 2010).

On the other hand, when TDP-43 is sequestrated in the cytoplasmic aggregates of

degenerative neurons, a dramatic depletion of TDP-43 from the nucleus is evident,

suggesting a redistribution/mislocalization of the protein (Winton et al. 2008).

Hence, functional TDP-43 is mandatory for the regulation of gene expression the

pathogenesis of the disease is likely to happen, at least partly, due to the normal

nuclear TDP-43 function(s) lost. This “loss of function” model is based on the

hypothesis that the cytoplasmic aggregates act as a “TDP-43 sink” and determine the

clearance of this protein from the nucleus. Indeed, this leads to the disruption of the

processes controlled by TDP-43 in the nuclear compartment.

Whether the TDP-43 in patients lead to disease due to the “gain of function” or “loss

of function” is still widely debated. Both models could either be mutually associated

or are exclusively acting in different time of the onset and progression of the disease.

Models of A LS:

Up to date, many different in vivo and in vitro models to study ALS were made based

on SOD1, FUS, or TDP-43 mutations, from yeast toCaenorhabditis elegans, flies,

zebrafish, and mouse. How accurately these models replicate ALS clinical symptoms

remains a troublesome question. Each model has its own advantages and

disadvantages. Choosing an appropriate model depends on the question being asked.

In vivo models of A LS SO D1 models: A number of C . elegans models have been developed that recapitulate many aspects

of ALS pathogenesis. These transgenic models mainly expressed either the SOD1

protein under the control of various gene promoters. The first C . elegans ALS model

was generated in 2001 by introducing human wild type and various human fALS

SOD1-linked mutations (A4V, G37R and G93A) under muscle-specific promoters

(Oeda et al. 2001). Some of the SOD1 mutants (G85R, G93A, 127X) in C . elegans

muscle cells resulted in mild cellular dysfunction(Gidalevitz et al. 2009).

45

Initial ALS studies in Drosophila Sod null flies showed either reduced longevity

and/or fertility, increased susceptibility to oxidative stress, motor deficits and/or

necrotic cell death in the fly eye(Phillips 1989).Several studies in this model showed

that selective expression of WT or human SOD1 disease linked mutants (A4V,

G85R) in motor neurons induced progressive motor dysfunctions, coupled with

electrophysiological defects and abnormal accumulation of the protein and a stress

response in surrounding glial cells (Watson et al. 2008).

Transgenic mice ubiquitously overexpressing various SOD1 gene mutations with

different biochemical properties, even in the presence of endogenous mouse Sod1

gene, develop a neurodegenerative disease that is quite similar to the human ALS.

Interestingly, transgenic mice overexpressing WT human SOD1 or mutant SOD1

only in neurons or only in glial cells do not develop disease (Bruijn et al. 1998; Gong

et al. 2000). A toxic gain-of-function rather than a loss-of-function of mutant SOD1

gene is therefore believed to be involved in ALS-linked SOD1 patients.

F US models

Overexpression of mutant FUS/TLS caused an accumulation of ubiquitinated

proteins, a pathological hallmark feature of ALS. A pathogenic role of human ALS-

associated FUS/TLS mutations (R524S and P525L) using Drosiphila has been

described (Chen et al. 2011). In this model overexpression of either wild type or

ALS-mutant in different neuronal subpopulations, including photoreceptors and

motor neurons led to an age-dependent progressive neuronal degeneration, including

axonal loss, morphological changes and functional impairment in motor neurons.

The drosophila ortholog of human FUS/TLS called Cabeza (Caz) has been also

disrupted in order to study FUS function. The Caz deficient fly exhibited reduced life

span and locomotion deficiency as compared with controls.

T DP-43 models:

Models of TDP-43-linked ALS have also been developed using different organisms,

such as Danio rerio and Caenorhabditis elegans.

46

Two groups have expressed human wild-type and mutant TDP-43 in zebrafish

embryos. Laird et al. haveobserved that expression of wild-type and mutant TARDBP

results in decreased axonal length and increased aberrant branching, with mutant

TDP-43 (TARDBPA315T) presenting more severe phenotypes than wild-type (Laird

et al. 2010).Interestingly, they noticed that both mutant and wild-type TDP-43

localized to the nucleus, suggesting mislocalization of the protein is not required for

axonal defects. Similarly, Kabashi et al. created transgenic zebrafish models

expressing wild-type or mutant (A315T, G348C or A382T) TARDBP. They found

that overexpression of mutant, but less so of wild-type TDP-43, causes motor defects

such as shorter motor neuron axons, aberrant branching and swimming deficits

(Kabashi et al. 2010). Altogether, these studies consistently show that overexpression

of wild-type TDP-43 is toxic in these systems and ALS-associated mutations in the

TARDBP gene seem to exacerbate the toxic phenotypes caused by overexpression of

this protein. Overall, these studies in zebrafish and C . elegans focuses on the

neuronal specific toxicity caused by expression of wild-type and mutant TDP-43

without providing insight into the way in which mutation in the TDP-43 gene can be

pathogenic in ALS. Despite their valuable contribution to the study of ALS, these

models have yet to address some of the main questions that still remain unclear in the

field, such as the mechanisms of neurodegeneration that lead to ALS pathology, or

the role that TDP-43 aggregation in the progression of the disease.

The fruit fly, Drosophila melanogaster, has long been recognized as one of the most

powerful genetic systems for studying neuromuscular development and function and

as such, it is no surprise that many current studies of ALS are using this organism to

advance our knowledge of this devastating disease. Recently, a number of studies

have demonstrated the importance of TDP-43 function in various aspects of neuronal

development and function in Drosophila (Feiguin et al. 2009; Hanson et al. 2010; Li

et al. 2010; Lu et al. 2009; Voigt et al. 2010). Loss of function analyses showed that

flies lacking the Drosophila TARDBP gene, TBPH, display anatomical abnormalities

at the neuromuscular junctions, reduced lifespan and motility defects (Feiguin et al.

2009). Interestingly, overexpression studies in Drosophila have demonstrated that

excess of human TDP-43 also results in a similar phenotypic outcome suggesting

that the cellular functions of this protein are complex and likely to be tightly

regulated since both loss and overexpression of the protein leads to pathology

47

(Hanson et al. 2010; Li et al. 2010). Therefore, the molecular mechanisms that

underlie TDP-43-linked ALS pathology are likely to be highly complex and both

loss- and gain- of function mechanisms of TDP-43 could have convergent

pathological outcomes. Furthermore, the expression of the ALS/FTLD-linked TDP-

43 mutations A315T and M337V in flies resulted in an increased neurodegeneration

(Estes et al. 2011);(Ritson et al. 2010).

New in vitro model of T DP-43 aggregation

TDP-43 became the one of the main players in ALS since was identified as the major

component of protein aggregates found in ALS/FLTD patients (Neumann et al.

2006). TDP-43 aggregation has been also found in the brain of patients affected by

other neurodegenerative disease such as Alzheimer disease (Geser et al. 2009). TDP-

43 is structurally similar to heterogeneous ribonucleoprotein (hnRNP) A∕B family

members. It contains two RNA recognition motifs (RRM) and a C-terminal domain

rich in glycine residues. It was first identified as a protein capable of binding the

TAR DNA and RNA sequences. Later it was identified as a factor that binds to an

intronic (UG) repeat element in the cystic fibrosis transmembrane receptor mRNA

and modulates its splicing. Glycine-rich C-terminal part of the protein was reported

to mediate protein-protein interactions with other hnRNPs (Buratti et al. 2005). This

region spans the part where the majority of human genetic mutations have been

identified(Del Bo et al. 2009; Kabashi et al. 2008; Kühnlein et al. 2008; Rutherford

et al. 2008; Van Deerlin et al. 2008). Currently, over 40 TARDBP mutations have

been found in ALS patients, while 3 have been found in FTLD patients (Borroni et

al. 2009; Gitcho et al. 2009).In addition, C-terminal region is cleaved through

caspase-3 generating C-terminal fragments (CTFs) that are associated with cellular

toxicity and/or increased TDP-43 mislocalization (Zhang et al. 2007).

Within the C-terminal region there is a domain rich in glutamine (Q) and asparagine

(N) residues, also known as aggregation-prone domain and is thought to confer

prion-like properties Therefore, wild-type form of TDP-43 is already highly

aggregation-prone and potentially toxic (Fuentealba et al. 2010). In fact, inclusion

bodies found in the central nervous system of a subset of ALS and FTLD patients

without TDP-43 mutations are immunopositive for TDP-43.The neurotoxicity of

48

TDP-43 may be related to its high propensity to aggregate and its cytoplasmic

mislocalization. Thus other factors that contribute to the severe toxicity of TDP-43

could be related to its function in transcription regulation and RNA processing.

Interestingly, Gln/Asn-rich region was found to be responsible for sequestration of

TDP-43 in polyglutamine aggregates causing the loss-of-function effects (Fuentealba

et al. 2010; Udan & Baloh 2011). Other studies focused on the TDP-43 mutations in

the region of 342-366 aminoacids leading to increased aggregation and toxicity in

vivo and in vitro (Rutherford et al. 2008; Del Bo et al. 2009).

Better understanding of the biochemical features of the C-terminal region of TDP-43

had open the doors to establish a novel aggregation model able to mimic the majority

of patient’s inclusion features.

For several members of the hnRNP-A/B protein family was demonstrated to be able

to interact with TDP-43 region spaning 321–366 aminoacids. In addition, for these

interactions was shown to be very important to maintain the splicing functionality of

TDP-43 (Buratti et al. 2005; D’Ambrogio et al. 2009). D’Ambrogio et al., have

finely mapped the region of the TDP-43 between aminoacid residues 321 and 366 to

be responsible for the interaction with hnRNP A2. When using 321–366 synthetic

peptide they were able to disrupt the TDP-43-hnRNP-A2 complex, suggesting the

stronger interaction of TDP-43 with its own sequence was stronger than the

interaction with hnRNP-A2(D’Ambrogio et al. 2009). Further on, Budini et al.,

investigating in details the minimal sequence needed for these interactions came out

with the residues 331–369. Rich composition in Gln/Asn residues of this sequence

and the observation that all the competing peptides caused the production of pre-

aggregates, they have decided to make a cellular model that could better mimic the

patient’s cytoplasmic aggregates. As an aggregation focus, they have choose to

engineere several plasmids carring 1, 4, or 12 tandem repeats of the 331–369 region

fused to EGFP(Budini et al. 2012).

Twelve tandem repeats of 331-369 Q/N-rich region of TDP-43 (12xQ/N) was able to

trigger the formation of the phosphorylated and ubiquitinated aggregates, both in

non-neuronal and neuronal cells, that recapitulates many features of the aggregates

observed in patients (Budini et al. 2012). Observed aggregates are indeedround,

large, ubiquitinated and phosphorylated formationspredominantly in cytoplasm. It is

49

interesting to note that these cytoplasmic aggregations caused nuclear TDP-43

depletion in this cell-based model of ALS, as already seen in the patient’s cells.

This novel research tool demonstarted an aggregation process that involves mainly

TDP-43 self-interactions that leads to phosphorylation, ubiquitination and possibly

nuclear TDP-43 depletion. Such a model is valuable to investigate the impact of the

aggregation on the cellular metabolism in the final stage of the aggregates observed

in patient’s brain, as well as a powerful tool to test new therapeutic strategies aimed

at reducing this phenomenon.

1.4 T reatment of A LS In addition to protein misfolding and aggregation, many other cellular pathologies

have been described in ALS patients, such as: glutamate excitotoxity, an excessive

stimulation of neurons by neurotransmitters such as glutamate, mitochondrial

dysfunction, inflammation and oxidative stress (Dunkel et al. 2012). Thus,

therapeutic approaches were mainly based around targeting one or more of these

cellular mechanisms.

Since its approval in the 1990‘s, riluzole, a presynaptic glutamate release inhibitor,

remains the only prescribed treatment for ALS patients offering a very modest

survival benefit of 2 to 3 months (Miller et al. 2007). Riluzole is shown to act solely

as an inhibitor of glutamate release. In addition it was reported that it is blocking

muscle sodium channels and the acetylcholine receptor involved in muscle

strengthning. Yet, it seems that Riluzole does not significantly affect muscle

strength or functional outcome in the patients. There are increasing number of the

drugs that are under investigation as a good candidate for ALS pathology treatment

(Table 2). Neuroprotective compounds from steroid family such as

Dihydrotestosterone (DHT) or the neurosteroid dehydroepiandrosterone (DHEA). It

was reported using ALS mouse models that these compounds are acting on

strengthening muscle through their neuroprotective effect, since ALS patients display

a severe muscle wasting and atrophy. Yet the exact mechanism underlying their

effect is still unknown(Hayes-Punzo et al. 2012). Arimoclomol, is a Heat-shock

protein (Hsp) co-inducer drug actually in Phase III clinical trial (Kalmar et al. 2008),

50

it is reported to reduce the level of the aggregates by boosting expression of

chaperones.

51

Compound Chemical Name structure Effect/Mechanism

Refrences/notes

Riluzole 6-(trifluoromethoxy)benzo[d]thiazol-2-amine

Glutamate signalling reduction

(McDonnell et al. 2012) F D A-approved riluzole

Dexpramipexole (R)-N6-propyl-4,5,6,7-tetrahydrobenzo[d]thiazole-2,6-diamine

Stabilizing mitochondria

in phase I I I clini cal trial (Glicksman 2011),(Dunkel et al. 2012)

Arimoclomol (R,Z)-3-(chloro((2-hydroxy-3-(piperidin-1-yl)propoxy)imino)methyl)pyridine 1-oxide

Co-inducer of heat-shock protein

in phase I I I clinical trial (Glicksman 2011),(Dunkel et al. 2012)

Olesoxime (8S,9S,10R,13R,14S,17R,E)-10,13-dimethyl-17-((R)-6-methylheptan-2-yl)-6,7,8,9,10,11,12,13,14,15,16,17-dodecahydro-1H-cyclopenta[a]phenanthren-3(2H)-one oxime

Modulation of Mitochondrial pore

in phase I I I clinical trial (Dunkel et al. 2012), (Glicksman 2011)

Tamoxifen (Z)-2-[4-(1,2-diphenylbut-1-enyl)phenoxy]-N,N-dimethylethanamine

Autophagy inducer (Wang et al. 2012)

Spermidine N1-(3-aminopropyl)butane-1,4-diamine

Autophagy inducer (Wang et al. 2012)

52

Carbamazepine 5H-dibenzo[b,f]azepine-5-carboxamide

Autophagy inducer (Wang et al. 2012)

Rapamycin (3S,6R,7E,9R,10R,12R,14S,15E,17E,19E,21S,23S,26R,27R,34aS)-9,10,12,13,14,21,22,23,24,25,26,27,32,33,34,34a-hexadecahydro-9,27-dihydroxy-3-[(1R)-2-[(1S,3R,4R)-4-hydroxy-3-methoxycyclohexyl]-1-methylethyl]-10,21-dimethoxy-6,8,12,14,20,26-hexamethyl-23,27-epoxy-3H-pyrido[2,1-c][1,4]-oxaazacyclohentriacontine-1,5,11,28,29 (4H,6H,31H)-pentone

Autophagy inducer (Staats et al. 2013)

DHT dihydrotestosterone

Neuroprotective (Yoo & Ko 2012)

DHEA dehydroepiandrosterone

neuroprotective (Hayes-Punzo et al. 2012)

53

Table 2: Chemical structure and mechanism of action of compounds reported to treat A LS disease. Adopted and reviewed in details in (Limpert et al. 2013)

Several other drugs are targeting dysfunctional mitochondria such as: olesoxime and

dexpramipexole. Both are involved in the prevention of mitochondrial dysfunction

and maintenance of energy production in stressed mitochondria within motor

neurons. Eventhough, dexpramipexole gave some promising results in the preclinical

trials, yet it failed in phase III of the clinical trial (Limpert et al. 2013; Glicksman

2011; Dunkel et al. 2012).

1.5 A LS treatment through autophagy induction

As many neurodegenerative diseases are described as proteinopathies and face a

problem with the intracytoplasmic protein inclusions, a viable therapeutic strategy is

PBA sodium phenylbutyrate

HDAC inhibitor (Ryu et al. 2005)

Trichostatin A 7-[4-(dimethylamino)phenyl]-N-hydroxy-4,6-dimethyl-7-oxohepta-2,4-dienamide

HDAC inhibitor (Yoo & Ko 2011)

DCA dichloroacetate

inhibits the pyruvate dehydrogenase enzyme, and modulates mitochondrial activity

(Miquel et al. 2012)

54

needed to overcome the aggregation problem. Indeed, promoting the clearance of

aggregate-prone proteins via pharmacological induction of autophagy-endolysosomal

pathway has proved to be a useful mechanism protecting against neuronal loss.

Autophagy is shown to be critical for motor neuron survival in many

neurodegenerative disorders such as spinal and bulbar muscular atrophy (SBMA,

Kennedy’s disease). Clinical symptoms and the histopathology of SBMA show many

similarities with ALS.SBMA is a progressive neuromuscular disorder in which

degeneration of lower motor neurons results in proximal muscle weakness, and

muscle atrophy. SBMA is among polyglutamine (polyQ) diseases that are caused by

the expansion in specific genes of a trinucleotides repeat, cytosine–adenine–guanine

(CAG), which encodes glutamine. CAG repeat expansion in the androgen receptor

(polyQ AR) accumulates in ubiquitin-positive aggregatesinside affected neurons of

SBMA patients(La Spada et al. 1991). In 2009, it has been demonstrated that

autophagy was able to degrade the aberrant proteins (polyQ AR) retained inside the

cytoplasm in mouse model of SBMA, thus pointing out the importance of the

autophagy activation in prevention of the motor neuron death (Montie et al. 2009).

In different ALS and TDP-43 models, similar studies have shown that autophagy

seems to have a major protective role in spinal motor neurons against

neurodegeneration. Full-length TDP43 has an exceptionally long half-life of 12–34 h

in most cell types, whereas TDP43 C-terminal fragments (CTFs) have a half-life of

only ~4 h. It is believed that both UPS and autophagic degradation pathways are

involved in maintaining the normal levels of TDP-43 inside cell and also the removal

of CTFs. Disruption of the UPS and autophagy might contribute to increased levels

of ubiquitylated TDP43 in ALS and FTLD-TDP. Indeed, inhibition of the UPS and

autophagy leads to increased levels of phosphorylated TDP43 aggregates in cultured

cells(Winton et al. 2008). Investigating the proteolytic pathways especially

autophagy, could be a promising strategy to treat TDP-43 aggregated disorders.

In fact, in primary motor neuron cultures, obtained from SOD1-G93A transgenic

mice (model of ALS), several autophagy substrates such as SOD1, ubiquitin, and

alpha-synuclein were significantly cleared when autophagy was induced using

lithium or rapamycin (Fornai, Longone, Cafaro, et al. 2008). Moreover, Kabuta and

55

colleagues reported that using wild-type and mutant SOD1 in neuronal and non-

neuronal cells, autophagy decreases mutant SOD1-mediated toxicity and its mutant

SOD1 protein aggregates levels (Kabuta et al. 2006).In line with this, Caccamo and

co-workers have overexpressed 25-kDa C-terminal fragment of TDP-43, that tends to

accumulate together with endogenous full-length TDP-43 in the cell, and were able

to modulate this accumulation and TDP-43 localization with the autophagy

activation(Caccamo et al. 2010). Wang et al., overexpressed TDP-43 and its

pathogenic form TDP-25 in HEK 293 cells and have demonstrated that the protein

levels of TDP-43 and TDP-25 were increased in cells treated with an autophagy

inhibitor, 3-MA, whereas, they were decreased in cells treated with an enhancer of

autophagy, trehalose(Wang et al. 2012).

Another study also confirmed the effect of rapamycin, as well as other autophagy

inducers, such as: spermidine, carbamazepine, and tamoxifen, in rescuing the motor

dysfunction of TDP-43 affected mice through autophagy pathway.Using FTLD-U

mouse model with TDP-43 proteinopathy, Wang et al., have shown that rapamycin

treatment effectively rescues the learning/memory impairment of these mice at 3

months of age, and it significantly slows down the age-dependent loss of their motor

function. These behavioral improvements upon rapamycin treatment are

accompanied by decreased level of caspase-3 and a reduction of neuron loss in the

brain of these mice. Furthermore, the number of cells with cytosolic TDP-43 positive

inclusions and the amounts of full-length TDP-43 as well as its cleavage products (35

kDa and 25 kDa) are significantly decreased upon rapamycin treatment. All these

studies suggest that autophagy induction may be a valid therapeutic target for TDP-

43 proteinopathies (Wang et al. 2012).

1.6 Dextran sulfate as inducer of autophagy

Dextran sulfate (DS) is a synthetic branched polyanionic sulfated compound derived

from purified dextran, that is originally synthesized by bacteria Leuconostoc

mesenterioides and then synthetically sulphated. Dextran sulfate can reach the level

of sulfur between 17 and 20 % of the total molecule. Once after being sulfated, every

glucose residue will present between 2 and 3 sulfate groups per glucosyl residue,

normally localized on carbons 2 and 4. 95% of the linkages between monomeres are

α-D-(16), and the other 5% are bindings of the type α-D-(13), directly related to

56

the chain branching of DS (Figure 9). DS is available commercially, with different

molecular weights, depending on the total amount of type α-D-(13) bindings

present and they can range from 5000 Da to 500.000 Da.

DS is a well-known compound used routinely in the biological field for decades. It

has been shown that low concentrations of DS can be used to selectively precipitate

lipoproteins such as VLDL and LDL, and higher concentrations of DS can also be

used to precipitate HDL. DS is in use as an agent able to separate DNA from histone

complexes, inhibit the binding of DNA with ribosomes, inhibit ribonucleases and it

has also been proven to inhibit initial binding of several viruses with cells, such as

the herpes virus and the HIV virus. Sulfated glycosaminoglycans, such as DS, have

been widely used in biotechnology as cell culture media additives, as they have

shown to promote cell growth, disperse cell clumps, increase viable cell yield, and

overall recombinant protein productivity.

Furthermore, for DS it has been shown to be safe for human consumption and indeed

DS has been used as an alimentary supplement for more than 20 years(Gräßler et al.

2013). It was originally discovered that DS possesses an anticoagulant capacity. Oral

F igure 9.Chemical structure of dextran sulfate

57

intake of DS is responsible for the reduced levels of LDL-cholesterol, HDL-

cholesterol, and triglyceride, confirming its action in reducing lipid metabolites in

humans. In this manner, DS has been used to prevent the progression of

atherosclerosis through its action as a competitor for the binding of modified LDL

with scavenger receptors in macrophages (Tsubamoto et al. 1994). Moreover, it has

also been reported that DS suppresses the expression of adhesion factors involved in

metastases, conferring anti-metastatic capacity to this polyanion (Takagi et al. 2005).

Furthermore, DS was able to work as a cytoprotectant of myocardial ischaemic and

damaged endothelium (Banz et al. 2005).

It has been proven that DS 5.000 Da, due to the presence of the sulfated groups, has

the capacity of mimicking endogenous glycosaminoglycans, thus showing many

effects on cells.Recently, DS was shown to act as an anti-apoptotic agent on CHO

cells, under staurosporine-induced apoptosis (Jing et al. 2011).They have shown that

dextran sulfate action involves mitochondrial pathway, since they were able to

measure a significant decrease of pro-caspase-9 activation, cleaved caspase-3 levels,

cytochrome c release into the cytosol and mitochondrial transmembrane potential

collapse.

In this note, our group have shown that DS5000 Da (DS5000) is able to prolong the

life of cells interfering with the apoptotic pathway (Menvielle et al. 2013). At a

molecular level, we show that DS inhibits apoptosis by DNA fragmentation delay

and decrease of annexinV-labeled cells, causes a G0/G1 cell-cycle arrest, decreases

p53 expression and increases the pro-survival factor Hsc70 expression. Moreover,

we were able to demonstrate that DS treatment also resulted in an enhanced LC3-I to

LC3-II conversion and increased autophagosomes formation employing tagged-LC3.

The experimental evidence provided in this work indicates that DS 5000 Da

treatment in low doses is able to induce autophagosome formation most likely in a

cell type independent manner. Nonetheless, our findings strongly suggest that

sulfated glycosaminoglycans should be re-considered, in this new context, as

autophagy inducers.

Aim of the study:

58

Previously established cell-based TDP-43 aggregation model, where tandem repeats

carrying the C-terminal Q/N-rich region of TDP-43 were able to trigger the

formation of phosphorylated and ubiquitinated aggregates, will be used in this study

to investigate the autophagy-endolysosomal cell pathway response.

As there is continuous need for the new and more effective agents in the treatment of

proteinopathies, it would be interesting to investigate if TDP-43 12xQ/N cell-based

model of aggregation could benefit from the autophagy induction and if DS

treatment is able to enhance inclusions clearance and modulate subcellular

distribution of TDP-43.

With such approach we hope to gain more knowledge about new therapeutic

strategies/effectors, which might have a broad therapeutic potential in human TDP-

43 proteinopathies.

59

2. M A T E RI A LS A ND M E T H O DS 2.1. Chemical reagents:

General chemicals were purchased from Sigma Chemical Co., Merck, Gibco

BRL, Boehringer Mannheim, Carlo Erba and Riedel-de Haen.

Dextran sulfate sodium salt 5000 Da (DS, Sigma) was dissolved in water and

sterilized through 0.22 µm filter.

Trichostatin A (TSA, from Sigma) was dissolved in DMSO, Four hours before

harvesting, cells were treated with 5 µM final concentration of TSA, and the

same concentration of TSA was kept in the cell lysis buffer, DMSO (Riedel-de

Haen) was used as vector control.

2.2. Standard solutions:

All solutions are identified in the text except for the following:

TE: 10 mM Tris-HCl (pH 7.4), 1 mM EDTA (pH 7.4)

PBS: 137 mM NaCl, 2.7 mM KCl, 10 mM Na2HPO4, 1.8 mM KH2PO4, pH 7.4

10X TBE: 108 g/l Tris, 55 g/l Boric Acid, 9.5 g/l EDTA

5X DNA sample buffer: 0.25 % w/v bromophenol blue, 0.25 % w/v xylene

cyanol FF, 30 % v/v glycerol in H2O.

Transfer buffer (proteins transfer from the SDS-PAGE 10% gel to the

nitrocellulose membrane):

Electro-transfer buffer 10X: 1 liter of this buffer should be prepared as follows:

Reagent Quantity

Tris (Invitrogen, cat. Nº 15504-020, or similar) 30,3 gr

Glycine(Sigma, cat. Nº 33226, or similar). 144 gr

Running buffer 10X: 1 liter of this buffer should be prepared as follows:

Reagent Quantity

60

Tris (Invitrogen, cat. Nº 15504-020, or similar) 30,3 gr

Glycine(Sigma, cat. Nº 33226, or similar). 144 gr

SDS (BDH, cat. Nº 301754L, or similar) 5 ml

2.3. Cells:

Bacter ia (for cloning)

E .coli D H5α competent cells

Genotype: F- 80dlacZ M15 (lacZY A-argF) U169 recA1 endA1hsdR17 (rk-, mk+)

phoAsupE44-thi-1 gyrA96 relA1.

Mammalian cells

H E K 293 F lp-In™ 293 T-Rex (Invitrogen™): Human Embryonic Kidney 293 cell

lines is designed for rapid generation of stable cell lines that ensure homogenous

expression of your protein of interest from a Flp-In™ expression vector. These cells

contain a single stably integrated FRT site at a transcriptionally active genomic

locus. Targeted integration of a Flp-In™ expression vector ensures high-level

inducible expression of the gene of interest (EGFP-12xQ/N) upon the addition of

tetracycline (see further).

U2OS: U2OS cell line, originally known as the 2T line, was cultivated from the bone

tissue of a fifteen-year-old human female suffering from osteosarcoma.

NSC-34 cell line: The motor neuron-like cell line NSC-34 is a hybrid cell line

produced by fusion of neuroblastoma with mouse motor neuron-enriched primary

spinal cord cells.Cultures contain two populations of cells: small, undifferentiated

cells that have the capacity to undergo cell division and larger, multi-nucleate cells

that express many properties of motor neurons.

2.4. Cell growth media:

Bacteria

61

Luria-Bertani (L B) medium: 10 g bacto-trypton, 5 g bacto-yeast extract, 10 g NaCl

per 1 liter medium. Ampicillin was added at a concentration of 50-100 µg/ml. For

hardening 1.5% agar-agar was added to the liquid medium.

Mammalian cells: D M E M complete media D M E M/Glutamax-I: Dulbecco's modified minimal

essential medium (Gibco-Invitrogen) supplemented with antibiotic-antimycotic

solution (Sigma) and decomplemented 10 % fetal calf serum (FCS, Gibco-

Invitrogen).

C ryo medium: For long-term storage, cells were frozen in liquid nitrogen in 90 %

FBS + 10 % DMSO.

2.5. Preparation of bacter ial competent cells:

Bacterial competent cells were prepared following the method described by Chung

and Niemela (Chung et al. 1989). Effectively, E.coli strains were grown overnight in

10 ml of LB at 37° C. The following day, 140 ml of fresh LB were added and the

bacteria were grown in the shaker at room temperature for 30-40 min until the

Optical Density (OD)600 nm was 0.3-0.38 at λ=600 nm. The cells were transferred

into a 50 ml falcon, then placed in ice, and then centrifuged at 4° C and 1000 rpm for

10 minutes. The pellet was resuspended in 1/10 of the initial volume of cold TSS

solution (10% w/v PEG molecular weight 4000, 5% v/v DMSO, 35 mM MgCl2, pH

6.5 in LB medium). The cells were aliquoted, rapidly frozen in liquid nitrogen and

stored in -80° C. Competence was determined by transformation with 0.1 ng of

PUC19 and was deemed satisfactory if this procedure resulted in more than 100

colonies.

2.6. Amplification of selected DN A fragments (PC R):

The polymerase chain reaction (PCR) was performed using plasmid DNA as

template, and following the basic protocols of Roche Diagnostic TaqDNA

polymerase. The volume of the reaction used was 50µl and comprised: 1x Taq

buffer, dNTPs mix (250 µM each), oligonucleotide primers (100 nM each), Taq

DNApolymerase (between 2.5 and 5 U), and 100 ng of the DNA template. All

62

synthetic DNA oligonucleotides used for PCR amplification were purchased from

integrated DNA technologies (http://www.idtdna.com). The standard amplification

conditions were the following: 94°C for 5 minutes for the initial denaturation, 94°C

for 45 seconds, 55° C for other 45 seconds, 72°C for 1 minute. This was repeated for

35 cycles, and after there was a final elongation phase at 72°C for 10 minutes.

Products of the PCR were routinely fractionated in 0.8-1% agarose gels. All

amplifications were performed on Cetus DNA thermal cycle (Perkin Elmer).

2.7. Enzymatic modifications of DN A :

Restriction enzymes

Restriction enzymes used for analytical digests were used according to

manufacturer's instructions. For these analysis, 100 to 500 ng of DNA were digested

in a volume of 20 µl containing the appropriate buffer for each particular enzyme.

The digested was incubated for 2 to 4 hours at the optimal temperature required by

the enzyme that is used. Preparative digestions of vectors and inserts were made of 5

to 10 µg DNA using the appropriate condition needed by the restriction enzyme in

100 to 200 µl reaction volume. Enzymatic activity was then removed either by heat

inactivation or by phenol-chloroform extraction.

DN A polymerase I , Large (K lenow) fragment

Klenow enzyme was used to treat PCR products from blunt-end creation by 3'

recessed fill-in. Briefly, the product obtained was supplemented with 1 µg of Klenow

fragment for every µg of DNA, and 4 mM MgCl2. The samples were afterwards

incubated for 30 minutes at 37° C. Enzyme was later inactivated by incubation at 65°

C for 20 minutes.

Calf intestinal phosphatase

Calf intestinal phosphatase CIP, provided from New England Biolabs Inc., catalyzes

the removal of 5' phosphate groups from DNA. When cloning a DNA fragment onto

a particular vector, CIP-treated fragments lack the 5' phosphoryl termini required by

ligases, for which vectors would be unable to self-ligate. This strategy was therefore

used in order to reduce the vector background in cloning strategies. The standard

63

reaction was carried out in a final volume of 50-100 µl using 1 U of enzyme for

every 0.5 µg of DNA, at 37° C for one hour.

Agarose gel electrophoresis of nucleic acid

DNA samples were size fractionated by electrophoresis in agarose gels ranging in

concentrations from 0.8% w/v (large fragments) to 1.5% w/v (small fragments). The

gels contained Ethidium Bromide (0.5 µg/ml) and 1X TBE solution. Horizontal gels

were used for fast analysis of DNA restriction enzyme digests, estimation of DNA

concentration, or DNA fragment separation prior to elution from the gel. DNA was

visualized by UV transillumination and the result recorded by digital photography.

E lution and purification of DN A fragments from agarose gels

This protocol was used to purify both vectors and inserts of DNA for cloning

purposes. The DNA samples were electrophoresed onto an agarose gel as described

previously. The DNA was visualized with UV light and the required DNA fragment

band was excised from the gel. This slab was cut into pieces, and the QIAquick gel

extraction kit was used, according to the manufacturer's instructions. Briefly, 1

volume of binding buffer QX1 was added to 1 volume of gel for DNA fragments and

incubated in 65° C until the gel was completely dissolved. Afterwards, the mixture of

binding buffer with the DNA was put into a prepared column, which was later

centrifuged at 10000g for 1 minute. Flow-through was discarded. Two rounds of

washing steps were later performed with 750 µl of wash PE buffer. After, the DNA

bound to the column was eluted with 30-50 µl of sterile water.

2.8. T ransformation of bacteria:

Transformation of bacterial clones was carried out using 1 ng of the plasmid DNA.

The DNA was incubated with 60 µl of competent cells for 20 minutes on ice and

then at 42° C for 2 minutes. After the step of the heat shock, 60 µl of LB were added

and the bacteria allowed to recover for 10 minutes at 37° C. the cells were then

spread onto agar plates containing the appropriate antibiotic. The plates were then

incubated for 12-15 hours at 37° C. When DNA inserts were cloned into β-

64

galactosidase-based virgin plasmids, 30 µl of IPTG 100 mM and 20 µl of X-Gal (4%

w/v in dimethylformamide) were spread onto the surface of the agar before plating to

facilitate screening of positive clones (white colonies) through identification of β-

galactosidase activity (blue colonies) which indicates the negative clones.

2.9. RN A preparation from cultured cells:

Cultured cells were washed and then RNA Trizol (Invitrogen SRL, Milano, Italia)

was added. Then, chloroform extraction was performed. Supernatant was

precipitated with isopropanol. The pellet was resuspended in 50 µl of dH2O and

digested with 1U of DNase RNase free. The mix was incubated at RT for 30 minutes,

and then the RNA was purified by acid phenol extraction. The final pellet was

resuspended in 33 µl of dH2O and frozen at -80° C. the RNA quality was checked by

electrophoresis on 0.8% agarose gels.

2.10. Estimation of nucleic acid concentration:

The concentration of a DNA sample can be checked by the use of UV

spectrophotometry. DNA absorbs UV lights, thus making it possible to detect and

quantify on very low concentrations. The nitrogenous bases in nucleotides have an

absorption maximum at about 260 nm. An optical density of 1.0 at 260 nm is usually

taken to be equivalent to a concentration of 50 µg/ml for double stranded DNA, and

approximately 40 µg/ml for single-stranded DNA and RNA. The ratio of the

absorbance at 260 nm vs 280 nm is a major of the purity of a sample. A sample in

which this ratio equals of tops the value of 1.8 is considered a pure sample. If this

value is lower than 1.8, then it is considered that samples present protein

contamination.

2.11. cDN A preparation and R T-PC R:

In order to synthesize cDNA, 3 µg of total RNA extracted from cells were mixed

with random primers (Finnzymes Oy., Vantaa, Finland) in a final volume of 20 µl.

After denaturation at 65° C for 5 minutes, the RNA and the primer were incubated

65

for 1 hour at 37° C in the following solution: 1X First standard Buffer (Gibco), 0.1

mM DTT, 10 mM dNTPs, RNase inhibitor 20U (Ambion) and Moloney murine

leukemia virus reverse transcriptase 100 U (Gibco) and retrotranscribed with

poly(dT) primer. 3µl of the cDNA reaction mix was used for the RT-PCR analysis.

The conditions used for the RT-PCRs were the following: 94° C for 5 minutes for the

initial denaturation, 94° C for 45 seconds, 56° C for 45 seconds, 72° C for 45

seconds, for 35 cycles, and 72° C for 10 minutes for the final extension. The RT-

PCRs were optimized to be in the exponential phase of amplification. PCR products

were resolved by 1% agarose gel electrophoresis (it was used to quantify mRNA of

Atg7 in the siAtg7 experiment before buying the Atg7 antibody).

2.12. Mammalian cell line and cell culture conditions:

U2OS, HEK 293 and HeLa cell lines were grown in DMEM/Glutamax-I (Gibco)

supplemented with 10% fetal bovine serum (Gibco) and antibiotic/antimycotic

suspension (Sigma).

NSC34 cell line was grown in DMEM/Glutamax-I (Gibco) supplemented with 5%

fetal bovine serum (Sigma) and antibiotic/antimycotic suspension (Sigma) and

differentiation was induced decreasing the serum concentration to 1% for at least 48

hours.

All cells were maintained at 37°C in a humidified atmosphere of 5% CO2. Cells

were grown in flasks and passaged once a week. Flp-In T-Rex HEK 293 were

maintained under Hygromycin and 5 µg/l of Blasticidin.

Passage of cell lines:

Plates containing a confluent monolayer of cells (U2OS, andFlp-In T-RexHEK 293

cell lines) were washed with PBS solution, then incubated at 37° C with 1-2 ml of

PBS/EDTA/Trypsin solution (PBS containing 0.04% w/v EDTA and 0.1% Trypsin)

for two minutes or until cells were dislodged. After adding 10 ml of media, cells

were pelleted by gentle centrifugation and resuspended in 5 ml pre-warmed medium.

1-2 ml of this cell suspension was added to 10 ml medium in a fresh plate and was

gently mixed before incubation.

66

NSC34 cells were plated in P100 plates, and only "Corning"® tubes were used, these

cells should not be treated with trypsin, instead NSC-34 cells were harvested by

scrapping them gently using cell scraper.

The cells suspension was split into flasks, dishes, plates or coverslips as required.

2.13. Plasmids:

pE G FP-12xQ /N construct was generated as mentioned in (Budini et al. 2012).

Construction of pF L A G-C M V4-T DP-43wt plasmid was described before (Ayala et

al. 2005). HcRed-hLC3 plasmid was a gift from Dr. Isei Tanida.

Plasmids ptfL C3 (Addgene 21074),

pcDN A-H D A C6-flag (Addgene 30482) and pcDN A-H D A C6.D C-flag (Addgene

30483) were purchased from Addgene.

Generation of the F L A G-tagged H D A C6 wild-type and catalytically-inactive

mutant

To make siRNA-resistant constructs pcDNA-HDAC6-flag and pcDNA-HDAC6.DC-

flag were mutated introducing the silent mutations with the use of the following

forward primer HDAC216G219T: gggactgcagggtatggatctgaac and corresponding

reverse primer.

2.14. T ransfection:

For transient transfection in U2OS cells, pEGFP-12xQ/N andHcRed-hLC3 we used

Lipofectamine 2000 (Invitrogen) according to the manufacturer's instructions.

The plasmid pFLAG-CMV4-TDP-43wt, in the adding back experiment of pcDNA-

HDAC216G219T.WT-flag and the catalytically inactive mutant(pcDNA-

HDAC216G219T.DC-flag) were introduced with calcium phosphate transfection

method.

Calcium phosphate transfection method:

The calcium phosphate transfection method is based on the formation of a calcium

phosphate-DNA precipitate. The calcium phosphate is thought tofacilitate the

67

binding of the DNA to the cell surface; subsequently, the DNA enters the cell by

endocytosis.

The transfection was done in this study according to the method established in the

lab. 4 hours before transfection, we change the medium for the cells. In one

Eppendorf tube, 25 µl of 500 mM CaCl2 and 5 µg of plasmid DNA were added, and

the final volume was adjusted to 50 µl with ddH2O. The formed DNA-CaCl2

complex was added drop by drop to 50 µl of 1X HBS prepared in another tube.

Aeration was cared to form during this step in order to obtain the maximum contact

between the DNA-CaCl2 complex and the HBS buffer, therefore very fine DNA-

CaCl2 precipitate. The DNA precipitate was drop wise added to the cells after

incubation for 25-30 minutes at RT. The cells were further incubated overnight at

37°C in a CO2 5% incubator.

2.15. G eneration of F lp-In™ T-Rex™ HEK 293 cells that

inducibly expressing E G FP-12xQ/N:

For these studies, we chose to use Human Embryonic Kidney 293 (HEK-293) cells

transfected with the Flp-In™ T-Rex™ Expression System (Flp-In™ T-REx™ 293

Cell Line, Invitrogen). This expression system uses a tetracycline-based repressor to

control transgene expression. In the Flp-In system, the integration of the expression

construct or thetransgene into the host cell genome occurs via Flp recombinase-

mediated intermolecular DNA recombination.We took EGFP-12xQ/N fragment from

the homonymous plasmid as XhoI and BamHI (Budini et al. 2012) and blunt-ended it

using Klenow fragment. Next, the expression vector pcDNA5/FRT/TO (Invitrogen)

was digested with the same restriction enzymes and EGFP-12xQ/N insert was cloned

inside. The pcDNA5/FRT/TO vector has a tetracycline-inducible promoter upstream

of the multiple cloning sites allowing the regulation of EGFP-12xQ/N expression.

Flp recombinase is expressed by the pOG44 plasmid. The recombination occurs

between the specific FRT sites on theinteracting DNA molecules.

The Flp-In™ TRex™ 293 cells were maintained in Dulbecco’s Modified Eagle

Medium (DMEM, Gibco), containing 10% fetal bovine serum (FBS) and

penicillin/streptomycin.EGFP-12xQ/N repetitions were digested from pEGFP-

68

12×Q/N plasmid and sub-cloned in the pCDNA5/FRT/TO vector. Flp-In-293 T-Rex

stable cell line was constructed upon co-transfection with the pCDNA5/FRT/TO-

EGFP-12xQ/N and pOG44 vector encoding Flp recombinase in a 1:4 ratio with

Lipofectamine The Flp-In T-Rex 293 cells were maintained in a humidified

incubator at 37ºC with 5% CO2under selective pressure (100 μg/ml hygromycin,

Invivogen). When necessary, the expression of pEGFP-12×Q/N was induced upon

tetracycline in the concentration 1.0 μg/ml.

2.16. Antibodies:

Antibody anti-Atg7 was obtained from Sigma (A2856), anti-LC3 (Sigma, L7543),

anti-p62 (Abcam, ab96134 for immunofluorescence), anti-p62 (Progen, GP62-C for

immunoblotting), anti-lamp1 (Abcam, ab25630), anti-flag (Sigma, F1804), anti-

TDP-43 (anti-TDP-43deltaQ/N, Protein Tech, 10782-2-AP), anti-actin (Sigma,

A2066), anti-tubulin (Sigma, T5168), anti-p84 (Abcam, ab487), anti-GFP (Santa

Cruz, sc-8334), anti-HDAC6 (Santa Cruz, sc-11420), anti-acetylated tubulin (Sigma,

611B-1), goat anti-mouse-AlexaFluor (Molecuar probe, A-11005), and goat anti-

rabbit-Texas red (Molecular probe, T-2767).

2.17. RN A interference (siH D A C6) and adding-back

experiment (H D A C6-W T ,H D A C6-D C):

Flp-In-293 T-Rex stable cell line expressing EGFP-12xQ/N was transfected twice

with oligofectamine (Invitrogen), distanced 24 hours, with 20 µM of siRNA for

Atg7. Sense strand of RNAi oligo (Sigma) for Atg7 was

GGUCAAAGGAUGAAGAUAA and siCONTROL against the firefly luciferase

gene was used as a non-targeting siRNA control. Cells were collected 24 hours after

the second transfection for total protein and total RNA extractions. Trizol reagent

(Invitrogen) was used for total RNA isolation. The level of Atg7 expression was

checked both by RT-PCR and immunoblotting using anti-Atg7 antibody.

Flp-In-293 T-Rex stable cell line expressing EGFP-12xQ/N were transfected three

times with 160nM of siRNA for HDAC6 using Hiperfectamine HiperFect (Qiagen)

in three consecutive days. Sense strand of RNAi oligo (Sigma) for HDAC6 was

69

CUGCAAGGGAUGGAUCUGAAC and siCONTROL against the firefly luciferase

gene was used as a non-targeting siRNA control. The first 24 hours the cells were

treated with tetracycline for EGFP-12xQ/N expression. The third day the cells were

transfected with siRNA-resistant pcDNA-HDAC6-flag and pcDNA-HDAC6.DC-

flag constructs and six hours post-transfection 100mg/l of dextran sulfate was added

to the cells culture media. The cells were harvested on the day four and whole cell

lysates were prepared for further immunoblotting.

2.18. SDS-PA G E :

For assessing the protein expression level, we ran protein samples on NuPAGE®

Novex® 4-12% Bis-Tris Precast Gel (Life technologies). Electrophoresis was

performed using a XCell SureLock® Mini-Cell, according to the manufacturer

instructions. In brief, protein samples were added with 2.5 µl of NuPAGE® LDS

sample buffer and 1 µl of NuPAGE® reducing agent, and water was added until a

volume of 10 µl was reached. Samples were then heated for 10 minutes at 70°C. for

preparing the 1X running buffer, 50 ml of 20X NuPAGE® MES running buffer were

added to 950 ml of deionized water. Sample was loaded onto the gel. After, both

upper and lower chambers were filled with 1X running buffer. Samples were ran for

40 minutes at constant 200V.

2.19. W estern blot analysis:

Cell pellets were lysed on ice in 300 μl of lysis buffer (15 mM HEPES, pH 7.5, 250

mM NaCl, 0.5% Nonidet P-40, and 10% glycerol) supplemented with proteases

inhibitors (Roche Applied Science mixture) and phosphatase inhibitors (10 mM NaF,

1 mM β-glycerol phosphate, and 1 mM Na3VO4). Cells were sonicated with 10

pulses at 50% of power (Bioruptor UCD-200, Diagenode) and centrifuged at 700 × g

for 10 min at 4 °C. Pellet was discarded and the total protein concentration in the cell

lysate was determined with Bradford reagent. Samples were resolved on a 4-12%

NuPAGE gel (Invitrogen), and then transferred to nitrocellulose membrane. Blots

were probed with different primary and secondary antibodies. The primary antibody

was diluted in 2% milk-PBS-Tween 0.1% v/v and left incubating overnight.

70

Secondary antibody was diluted 1:2000 in 2% milk-PBS-Tween 0.1% v/v and left

incubating for 1 hour. Reactive species were visualized by ECL chemiluminescence

kit (Thermo scientific).

Cell lysate fractionation and filter trap assay:

For the cell lysate fractionation, 50 μg of each total cell lysate was centrifuged at

16,000 × g for 30 min at 4 °C. The supernatant was transferred into a new tube as a

soluble fractions (S), and the pellet (insoluble fractions, P) was washed again with

lysis buffer and resuspended in 8 M Urea, and sonicated again. To analyze each

fraction by immunoblotting, 50 μg of total lysates (IN), the entire soluble fraction

(S), and all pellet volume (P) were loaded in a 10% SDS-PAGE under reducing

conditions. The membrane was tested with anti-GFP, anti-flag, anti-actin and anti-

tubulin antibodies. For filter trap assay 50 µg of total lysates were loaded on the pre-

wetted cellulose acetate 0.2 µm membrane (Sterlitech) using a slot-blot device. Each

well was washed twice with PBS and immunoblotting with anti-GFP antibody was

performed.

2.20. Immunofluorescence microscopy:

U2OS cells (1×105) and Flp-In-293 T-Rex stable cell line (5x104) were plated in 6-

well plates containing coverslips. The next day U2OS cells were transfected with

pEGFP-12×Q/N plasmid. Flp-In-293 T-Rex stable cells were treated with

tetracycline for 24 hours in order to induce EFGP-12xQ/N expression. When

immunofluorescence required, the cells were washed with PBS and fixed in 4%

paraformaldehyde in PBS (15 min at room temperature) and permeabilized by using

0.3% Triton X-100 in PBS for 5 min on ice. After blocking with 2% BSA/PBS for 20

min at room temperature, an immunolabeling was carried out at room temperature

for 1 h in 2% BSA/PBS. Cells were washed 3 times with PBS and incubated with

conjugated secondary antibody at room temperature for 1 h. Nuclei were stained with

4, 6-diamidino-2-phenylindole (DAPI). The analysis was done on a Zeiss LSM 510

confocal microscope (63x1.4NA plan-apochromat oil immersion lens) using Zeiss

LSM510 software. Image J software was used for subsequent image processing.

71

Colocalization was determined by calculating the Pearson’s coefficient from the

individual cells with Volocity software (Perkin Elmer).

2.21. Clearance assay in stable Hek293 cells expressing

E G FP-12xQ/N

Expression of EGFP-12xQ/N in Flp-In-293 T-Rex stable cell line was induced with

1ug/ml tetracycline for 24 hours and then the transgene expression was switched off

by removing tetracycline from the medium. Further on, cells were treated with

dextran sulfate (100 μg/ml) and the amount of transgene product decay was detected

after 48 hours later by immunoblotting with antibody against GFP. In addition,

approximately 800 cells were counted by fluorescence microscope for the proportion

of cells with EGFP-12xQ/N aggregates after DS treatment in comparison to the

control. Error bars represent standard deviations from three independent experiments.

2.22. Viable cell count determination:

Tetrazolium dye assay (MTT) was used to measure cytotoxicity (loss of viable cells)

after dextran sulfate 5000 Da additions into the media. NSC34 cells were plated (10

000 cells per well) before or after EGFP-12xQ/N transfection using lipofectamine

2000. After 6 hours dextran sulfate was added into the media and the number of

viable cells was assessed after 48 hours. 20 µl of 5mg/ml of MTT solution (Sigma)

was added to each well and incubated for 4 hours at 37°C in a humidified, 5% CO2

atmosphere. Later, the micro-plate was centrifuged at 1500 rpm, 15 min, 5ºC and the

culture media was removed from the wells by gentle vacuum suction. The micro-

plate was frozen for at least one hour and 100 µl of isopropanol was added in each

well, mixed and the absorbance was read at a spectrophotometer (570 nm).

72

3. R ESU L TS

3.1. Cellular model of T DP-43 aggregation and autophagy-lysosome

pathwayinvolvement

TDP-43 became one of the main players in Amyotrophic Lateral Sclerosis (ALS) and

Fronto-Temporal Dementia (FTD) since it was identified as the major component of

the pathological protein aggregates in patients. In patient’s affected neurons TDP-43

is mislocalized to the cytoplasm, forming inclusions that contain ubiquitinated,

hyperphosphorylated and cleaved TDP-43 (Neumann et al. 2006; Arai et al.

2006).Over 40 TDP-43 mutations have been associated with pathological conditions

collectively termed as TDP-43 proteinopathies (Barmada & Finkbeiner 2010). Most

of these mutations were localized in C-terminal region of TDP-43 (Pesiridis et al.

2009), generally found to be responsible to mediate protein-protein interactions

(D’Ambrogio et al. 2009). In addition, C-terminal region once cleaved, it generates

C-terminal fragments (CTFs) that were found associated with cellular toxicity and

TDP-43 mislocalization in ALS patients (Winton et al. 2008; Zhang et al. 2009). C-

terminal region of TDP-43 was mapped and has been shown to contain Gln/Asn

(Q/N) rich domain able to enhance inclusion formation(Igaz et al. 2009), thus acting

as a potential prion-like domain(Udan & Baloh 2011). In previous works, it has been

demonstrated that the C-terminal Q/N–rich region of TDP-43 is involved in the

73

interaction with the other members of the hnRNP family and with itself

(D’Ambrogio et al. 2009; Fuentealba et al. 2010). On the base of these findings,

Budini et al., have made twelve tandem repeats of 331-369 Q/N-rich region

(12xQ/N) fused to EGFP that was able to trigger the formation of cytoplasmic

aggregates in non-neuronal and neuronal cell lines. Indeed, these EGFP-12xQ/N

aggregates are predominantly cytoplasmic, phosphorylated and ubiquitinated and are

able to recapitulate many features observed in the ALS patients (Budini et al. 2012).

Therefore, since one of the major problems of ALS patients are intra-cytoplasmic

protein inclusions, this cellular model is considered as a useful in vitro model of

TDP-43 aggregation that can be used to investigate differential roles of cellular

protein degradation pathways.

Ubiquitin proteasome system (UPS) and autophagy-lysosome pathway are the two

main cellular mechanisms for abnormal proteins removal. The UPS usually

successfully degrades misfolded and/or damaged proteins. If the amount of proteins

targeted to the UPS surpasses the degradation efficiency rate of the proteasome,

subsequent oligomerization and accumulation of proteins activate autophagy

(Korolchuk et al. 2010). On this way, aggregated proteins that are too large to enter

the UPS pore become substrates for autophagy (Verhoef et al. 2002). Autophagy is

the catabolic process by which the cell degrades cytoplasmic components within

lysosomes (Kroemer et al. 2010). Autophagy has been long thought to be an essential

non-selective bulk degradation pathway. However, recently accumulating evidence

has highlighted the selective elimination of unwanted components such as aberrant

protein aggregates, dysfunctional organelles and invading pathogens by autophagy

(Hubbard et al. 2012). The selective degradation of protein aggregates by autophagy

is called aggrephagy (Lamark & Johansen 2012). Some of the molecular components

involved in selective autophagy have been identified, such as specific autophagy

receptors able to guide the cargo to the site of autophagosomal engulfment. For

example, protein ubiquitination can be a sensor that targets proteins to the

autophagosomes (McEwan & Dikic 2011). The recognition of ubiquitinated

substrates is mediated by molecular adaptors including p62, NBR1, and optineurin,

which bind on one side to ubiquitin and on the other to the autophagosome-specific

proteins, such as LC3/GABARAP/Gate16 family members (Johansen & Lamark

2011).

74

We decided to use EGFP-12xQ/N TDP-43 aggregation model in transiently

transfected U2OS cells as described by Budini et al. to test the autophagic response

of the cell. In order to test this, we performed immunocytochemistry using

autophagy-lysosome pathway markers. A commonly used method to measure

autophagic activity is the direct determination of LC3 localization employing tagged-

LC3 (Mizushima 2004).Tagged LC3 can be seen diffused throughout the cell, while

inside autophagosomes can be distinguished as clear fluorescent dots. To

characterize aggregates labeling by an autophagosome marker LC3 (Kabeya et al.

2000), we cotransfected U2OS cells with pEGFP-12xQ/N and pHcRed-hLC3

constructs and followed their expression and localization under fluorescence

microscope. We were able to see certain cells where exogenous LC3 (in red)

colocalize with EGFP-12xQ/N aggregates (in green) (Figure 10A). This result

suggested an autophagic fate of the EGFP-12xQ/N aggregates.

As amyotrophic lateral sclerosis disease is a late-onset progressive neurodegenerative

disease affecting motor neurons, we wanted to see if the aggregation pattern of

EGFP-12xQ/N would be maintained in the motor neuronal cells in vitro. NSC-34 is a

hybrid cell line produced by fusion of motor neuron enriched, embryonic mouse

spinal cord cells with mouse neuroblastoma, therefore suitable for this study. We

decided to assess the expression of EGFP-12xQ/N in NSC-34 cells and to check

whether it is possible to see GFP-12xQ/N aggregates sequestered inside

autophagosomes. Therefore, 48 hours post differentiation (cell culture medium

conditions change from 5% to 1% of serum), we cotransfected NSC-34 cells with

EGFP-12xQ/N and HcRed-hLC3 plasmids. Cells were fixed and visualized under

confocal microscope. We were able to see some cells where EGFP-12xQ/N

aggregates were sequestered inside the autophagosomes (Figure 10B) in agreement

with the results obtained in U2OS.

We further examined the localization of degradation adaptor protein, p62/SQSTM1

(p62/sequestosome) and EGFP-12xQ/N aggregates. Protein p62 is involved in

targeting ubiquitinated proteins to the autophagy pathway and is responsible for

autophagy elimination of misfolded proteins(Johansen & Lamark 2011). In addition,

it has been reported that aggregated TDP-43 in ALS and FTD tissue were labeled by

p62. We have found some large EGFP-12xQ/N aggregates (in green) decorated with

75

an endogenous p62 (in red) (Figure 10C). This evidence indicated adaptor protein

p62 recruitment in response to the EGFP-12xQ/N aggregation.

Furthermore in order to confirm autophagy-lysosome pathway respond to the

aggregation we sought to investigate if EGFP-12xQ/N aggregates are labeled with

the late endosome/lysosome marker Lamp 1 (lysosomal associated membrane protein

1). After transfecting U2OS cells with EGFP-12xQ/N plasmid, the cells were stained

for the endogenous Lamp1. Under the fluorescence microscope some cells show an

overlapping of EGFP-12xQ/N (in green) and Lamp1 (in red) signals (Figure 10D).

Our findings to date suggested that autophagy-lysosome markers LC3 and Lamp 1

and adaptor protein p62 label some EGFP-12xQ/N aggregates in our cellular TDP-43

aggregation model, probably targeting them for degradation through the autophagy.

76

F igure 10. E G FP-12x Q /N aggregates are labeled with L C3, p62 and Lamp1.

77

(A) U2OS cells cotransfected with pEGFP-12xQ/N (in green) and HcRed-hLC3

(in red) plasmids show clear colocalization of an autophagosome marker

LC3 with aggregates (merge in yellow). To visualize the cells, nuclei were

stained with 4, 6-diamidino-2-phenylindole (DAPI). Cells were analyzed on

Leica DMIRE2 inverted fluorescence microscope. Images were acquired

using 63X oil immersion objective. (B) Mouse motor neuronal-like cells

NSC-34 cells were cotransfected with pEGFP-12xQ/N (in green) and

HcRed-hLC3 (in red). Merge channel shows a clear colocalization of an

autophagosome marker LC3 with aggregates (merge in yellow). Nuclei are

shown with dotted lines. Images were acquired using Zeiss LSM 510

confocal microscope under 63X oil immersion objective. (C) EGFP-12xQ/N

expressing U2OS cells were labeled with p62 antibody (in red) in order to

visualize this degradation adaptor protein within cell. Some cells with

aggregates (in green) show clear colocalization with p62 protein (in red).

Cells were analyzed on Zeiss LSM 510 confocal microscope. Images

represent a single confocal section of 0.1 µm acquired using 63X oil

immersion objective and 2X zoom. (D) Antibody against lysosome marker

Lamp1 (in red) revealed colocalization of this late endosome/lysosome

marker with EGFP-12xQ/N aggregates (in green) in some U2OS transfected

cells. Dotted lines are showing the position of the nuclei. Cells were

analyzed on Zeiss LSM 510 confocal microscope. Images represent a single

confocal section of 0.1 µm acquired using 63X oil immersion objective and

2X zoom.

78

3.2. Dextran sulfate 5000 Da decreases the number of cells containing E G FP-

12xQ /N aggregates

Viable therapeutic strategy is needed to overcome the aggregation problem in TDP-

43 proteinopathies and one of the strategies would be to promote the clearance of

aggregates via pharmacological induction of the intracellular protein-degradation

pathways such as autophagy-lysosome pathway. Indeed, inducing autophagy-

lysosomal pathway has been proved to be a useful mechanism against aggregation in

ALS (Otomo et al. 2012). One of the well-known inducers of autophagy and widely

tested in different protein-aggregates prone models is rapamycin. Even in ALS

model, rapamycin has been tested and proved to have a beneficial effect in correcting

TDP-43 mislocalization through the enhancement of autophagy (Caccamo et al.

2009).

In this context, it was rational to use our EGFP-12xQ/N cellular model of TDP-43

aggregation as a useful tool to test possible drug candidate able to reduce

aggregation. In the previous work we have shown that sulfated polyanion dextran

sulfate 5000 Da (DS) was able to enhance survival of the different stressed cell

cultures, not just through the apoptosis inhibition, but also through the autophagy

induction (Menvielle et al. 2013). As this finding suggested that DS in low

concentration favors autophagy, we decided to investigate if our TDP-43 aggregation

model could benefit from it.

To investigate the effect of DS on protein aggregation, we transfected U2OS cells

with pEGFP-12xQ/N construct and six hours after transfection, cells were treated

with different concentrations of dextran sulfate 5000 Da (50, 100 and 150 mg/l).

After 24 hours of DS treatment, the cells with aggregates were counted within the

population of the transfected cells under the microscope. Figure 11A is showing the

profile of cells expressing EGFP-12xQ/N in control and DS treated sample. In the

control, most of the transfected cells are showing dot like structures in the cytoplasm

(green dots), whereas among transfected cells that are treated with different

concentrations of DS, cells were containing either the aggregates or diffused GFP

signal in the cytoplasm. We have counted approximately 100 EGFP-12xQ/N

expressing cells for each sample and scored as positive any EGFP-expressing cell

that has at least one dot (big or small sized aggregates) and any cell showing a

diffused GFP profile was scored as negative. Figure 11B shows statistical analysis

79

from three independent experiments for control and each DS concentration. DS

treatment decreased the percentage of cells with aggregates to about 50% in all three

tested concentrations (t-test, **p<0.01) (Figure 11B). Since all three tested DS

concentrations have been shown to be effective in aggregation reduction, we have

decided to use 100 mg/l concentration for further experiments.

F igure 11. Dextran sulfate 5000 Da decreases the number of cells containing

E G FP-12xQ /N aggregates. (A)U2OS cells were transfected with EGFP-12xQ/N

constructand analyzedunder fluorescence microscope. EGFP-12xQ/N aggregates are

visible as green dot-like structures in the cytoplasm.Treatment with dextran sulfate

decreased significantly number of aggregates-containing cells. Upon DS treatment

green signal was diffused in the 50% of transfected cells after 24 hours. To visualize

the cells, nuclei were stained with 4, 6-diamidino-2-phenylindole (DAPI). Cells were

analyzed on Leica DMIRE2 inverted fluorescence microscope. Images were acquired

using 63X oil immersion objective. (B) Percentages of transfected cells with EGFP-

12xQ/N aggregates were determined for control and dextran sulfate treated samples

from 100 transfected cells. DS treatment decreased the percentage of cells with

aggregates in all three tested concentrations. Three independent experiments were

included for three different DS concentrations (50, 100, and 150 mg/l). Statistical

significance was evaluated using t-test (n=3, independent experiments, t-test,

**p<0.01). Error bars represent SD.

80

3.3 Positive trend in autophagic response to the aggregation upon DS treatment

In order to quantify the aggregates labeled with LC3 and thus marked for autophagy,

the cells were co-transfected with EGFP-12xQ/N and HcRed-hLC3 plasmids and

treated with 100 mg/l of DS. Percentage of cells with pEGFP-12xQ/N aggregates

that are positive for autophagy (cells with more than 8 LC3 red dots, thus exceeding

the basal autophagy) was determined by fluorescence microscopy. We noted that

almost all DS-treated cells with aggregates are LC3 positive (containing more than 8

dots) unlike the non-treated cells (n=3, independent experiments, t-test, ***p<0.001)

(Figure 12A). This result is suggesting that the cells with aggregates after DS

treatment show more autophagic activity compared to the non-treated cells.

Additionally, we have used Pearson’s correlation coefficient to quantify

colocalization between EGFP-12xQ/N and EGFP-LC3 by tracing individual cells

with Volocity software (Perkin Elmer). DS treatment increased colocalization of

EGFP-12xQ/N and red-LC3 when compared with control (Figure 12B and 12C).

This effect is an indicator of more efficient LC3 recruitment (autophagosomes) on

the EGFP-12xQ/N aggregates upon DS treatment in the cell. Altogether, these

findings led us to the idea that DS treatment is able to enhance an autophagic activity

in the cells with EGFP-12xQ/N inclusions.

81

F igure 12.Positive trend in autophagic response to the aggregation upon DS

treatment.(A) U2OS cells were seeded on cover slips and cotransfected with both

pEGFP-12xQ/N and pHcRed-LC3 plasmids. Six hours post transfection dextran

sulfate 5000 Da was added in the media in 100 mg/l concentration. 24 hours later,

cells were examined under fluorescence microscope for the number of cells with

EGFP-12xQ/N aggregates containing more than 8 LC3 puncta. Positive cells for LC3

(>8 red-dots per cell) within the population of cells with EGFP-12xQ/N aggregates

were counted from 340 cells in control and dextran sulfate treated samples in three

independent experiments (n=3, independent experiments, t-test, ***p<0.001). Error

bars represent SD (B) Colocalization between EGFP-12xQ/N and red-LC3 is

expressed in terms of the Pearson’s coefficient. Pearson’s correlation coefficient was

calculated using Volocity software (Perkin Elmer) in 38 control and 30 DS treated

cells from three independent experiments (n=3, independent experiments, t-test,

*p<0.05). Error bars represent SEM. (C) Cells were analyzed on Zeiss LSM 510

82

confocal microscope. Images displayed are a single confocal section of 0.1 m.

Images were acquired using 63X oil immersion objective and 2x zoomed.

3.4. Establishing stable cell model of E G FP-12xQ /N aggregation

We have demonstrated that U2OS cells transiently transfected with EGFP-12xQ/N

and treated with DS show significant reduction of the EGFP-12xQ/N aggregation

compared with the control (Figure 11). In order to examine in details the observed

DS enhanced aggregates clearance, to exclude the possibility that DS inhibit

aggregates formation and to obtain more consistent and regulated expression system,

we have developed stable HEK293 cell line expressing EGFP-12xQ/N under

tetracycline inducible promoter. We used Human Embryonic Kidney 293 cells (HEK

293) transfected with the Flp-In™ T-Rex™ Expression System (Flp-In™ T-REx™

293 Cell Line, Invitrogen) as an inducible cell line model. This system, enable us to

have one copy of EGFP-12xQ/N integrated in the genome. Moreover, this expression

system uses a tetracycline-based repressor to control transgene expression, which

allows us to switch on/off the system within any desired time.

To make HEK293 stable cell line, we took EGFP-12xQ/N fragment from the

homonymous plasmid as XhoI and BamHI (Budini et al. 2012) and blunt-ended it

using Klenow fragment. Next, the expression vector pcDNA5/FRT/TO (Invitrogen)

was digested with the same restriction enzymes and EGFP-12xQ/N insert was cloned

inside. The pcDNA5/FRT/TO vector has a tetracycline-inducible promoter upstream

of the multiple cloning sites allowing the regulation of EGFP-12xQ/N expression.

The advantage of the Flp-In™ T-Rex™ expression system is the use of the FRT

recombination sites in the pcDNA5/FRT/TO expression vector to mediate the site-

specific recombination of the target gene into the genome. We took the advantage of

the Flp recombinase protein, encoded by the pOG44 vector (Invitrogen) that

mediates DNA recombination, to introduce the gene of interest into the genome. On

this way, pcDNA5/FRT/TO-GFP-12xQ/N vector was cotransfected with pOG44

vector into the HEK 293 cells following the manufacturer’s protocol. Following a

single positive clone selection with hygromycin (100 µg/ml), the expression of

EGFP-12xQ/N construct was confirmed under the fluorescence microscope upon 24

hours of tetracycline induction. This system was capable to reproduce the feature of

thetransiently transfected U2OS cells where dot-like structures are extensively

formed in the cytoplasm (Figure 13).

83

F igure 13. E G FP-12xQ /N aggregates formed inH E K293 F lp-In T-Rex stable cell

lineafter tetracycline induction. HEK293 cells were seeded on cover slip and

treated with tetracycline for 24 hours. Aggregates are visible as green fluorescent

dots in the cytoplasm of all cells. Cell nuclei were visualized with DAPI (blue). Cells

were analyzed on Leica DMIRE2 inverted fluorescence microscope. Images were

acquired using 63X oil immersion objective. Right panel is the zoomed image.

3.5. Dextran sulfate 5000 Da enhances the clearance of E G FP-12xQ /N

aggregates in stable H E K 293 cell line

Next, we examined if DS 5000 Da is able to enhance the clearance of EGFP-12xQ/N

in stable HEK 293 cell line using the assay conditions previously described by

Sarkar et al. This method named clearance assay is used when targeted transgene

expression is under inducible promoter and it allows one to assess if specific agent is

able to alter the clearance of the transgene product, following the amount of

transgene product decay when its synthesis is stopped (Sarkar et al. 2007).

Effectively, we used the same principle to test if DS was able to enhance the

clearance of EGFP-12xQ/N aggregates. The transgene expression was first induced

with tetracycline and 24 hours later EGFP-12xQ/N expression was switched off by

removing tetracycline from the media and DS was added at 100 mg/l concentration.

48 hours later, approximately 800 cells were counted under the fluorescence

microscope for the proportion of cells with EGFP-12xQ/N aggregates. As positive

were counted cells with small and/or big green fluorescent aggregates, while cells

with green diffused signal were considered as negative. DS increased the clearance

of the EGFP-12xQ/N inclusions, as the number of cells containing aggregates is

84

reduced by 65% in comparison to the control (n=3, independent experiments, t-test,

***p<0.001) (Figure 14). Error bars represent standard deviations from three

independent experiments.

F igure 14. Dextran sulfate 5000 Da enhances the clearance of E G FP-12xQ /N

aggregates. (A) A microscopic view of EGFP-12xQ/N aggregates (in green) in both

control and DS samples. HEK293 Flp-In T-Rex EGFP-12xQ/N were seeded on

cover slips and treated with tetracycline for 24 hours, then expression of EGFP-

12xQ/N is switch off by taking out tetracycline from the medium and dextran sulfate

(100mg/l) was added for additional 48 hours. Cells were then fixed and visualized

under fluorescence microscope. Cells expressing EGFP-12xQ/N with green big

and/or small sized dots were considered as positive, while cells with green diffused

signal are scoredas negative. DS is enhancing the clearance of EGFP-12xQ/N

aggregates from the cell. Nuclei were stained in blue with DAPI and upper figures

are merged views from green and blue channels. Cells were analyzed on Leica

DMIRE2 inverted fluorescence microscope. Images were acquired using 63X oil

immersion objective. (B) Statistical analysis of aggregates in control and DS treated

cells in the clearance assay.

85

Next, and in order to confirm the DS clearance activity in our TDP-43 aggregation

model, we performed a western blot of whole cell lysates with anti-GFP antibody.

DS treated cells' extracts showed a significant EGFP-12xQ/N product decay

compared to the control (Figure 15A) giving an evidence of the enhanced clearance

of the transgene product when its synthesis is stopped.It should be noted that the

efficiency of the aggregate clearance induced by Dextran Sulfate showed variations

with different Dextran Sulfate batches, we are currentlytrying to understand the role

of possible impurities in the DS preparations.We next measured the effect of DS on

EGFP-12xQ/N insoluble fraction using filter binding assay, where whole cell lysates

from control and DS treated cells were applied on a pre-wetted cellulose acetate 0.2

µm filter using a slot blot chamber. After washing with PBS the membrane was

usedto perform a Western blot with anti-GFP antibody. Insoluble aggregates are not

able to pass through the filter and are detected with anti-GFP antibody, while soluble

proteins are passing through the membrane. HEK293 cells expressing GFP were

used as a protein solubility control as all GFP is soluble and should not be leaving

the track on the filter (Figure 15B, GFP line). In the clearance assay, 48 hours DS

treated samples showed aggregates content reduction to 0.6 fold of control as

quantified with Image J software (Figure 15B, DS vs ctrl). These results allowed us

to propose DS 5000 Da as an effective EGFP-12xQ/N aggregates clearance agent.

86

F igure 15. DS enhances the clearance of E G FP-12xQ /N aggregates. (A)Western

blot analysis of whole cell lysates of control and DS treated cells in the clearance

assay. DS is causing significant decay in the EGFP-12xQ/N product amount after 48

hours in the clearance assay quantified with Image J software.Please note that some

problems of reproducibility and variability in the aggregateclearance efficiency was

observed if the DS batches were changed (See text).(B)Filter trap assay was used to

quantify the aggregates in DS treated samples in comparison to the control.

Wholecell lysates were prepared and loaded on the cellulose acetate membrane (0.2

µm), where insoluble aggregates cannot pass the membrane pores unlike soluble

proteins GFP which its expressing cell'extract was used as a protein solubility control

(line GFP). Membrane was used for western blot with anti-GFP antibody. DS is

reducing EGFP-12xQ/N aggregates content to 0.6-fold of the control as quantified

with Image J software (graph on the right).

87

3.6. DS induces E G FP-12xQ /N aggregates clearance through the autophagy

pathway

Macroautophagy (here referred to as autophagy) is an intracellular process

responsible for the degradation of protein aggregates. During autophagy, a double-

membrane vesicle (the autophagosome) engulfs the cargo and subsequently fuses

with the lysosome, which degrades the cargo from the autophagosome. Over the last

few years it has been shown that autophagic membranes can selectively target cargo

via [MAP1(LC3)] and gamma-aminobutyrate receptor associated protein

(GABARAP), which act as receptor molecules for further degradation. Important

players acting as autophagy adaptor proteins, such as p62 and NBR1 that bind

directly to LC3 were identified. The autophagy receptors are themselves degraded by

autophagy, and they mediate selective autophagy via interactions with substrates that

are simultaneously degraded(Bjørkøy et al. 2005; Ichimura et al. 2008; Pankiv et al.

2007).

In order to check if autophagy was involved in EGFP-12xQ/N aggregates clearance

upon DS treatment, we determine the level of LC3-I to LC3-II conversion in both

control and DS treated cells. Autophagy can be measured by changes in LC3

localization as the level of conversion of LC3-I to LC3-II provides an indicator of

autophagic activity. In particular the level of LC3-II correlates with autophagosome

formation, due to its association with the autophagosome membrane. We performed

a clearance assay, as described before, and after 48 hours of DS addition into the

media whole cell lysates were prepared. Here again, we checked for a level of EGFP-

12xQ/N in the DS treated cells versus control and confirmed the decline of the

product (Figure 16, EGFP-12xQ/N). Next, we observed an increase of LC3-II

formation versus LC3-I after DS treatment (Figure 16, LC3I/LC3II) suggesting an

autophagy promotion upon DS treatment.

The other important player involved in the autophagy pathway isp62, an aggregate-

bound protein that recruits the autophagy machinery to the aggregates(Myeku &

Figueiredo-Pereira 2011). Moreover, the protein itself is degraded by autophagy and

accumulates when autophagy is inhibited, and thus may be used as a marker to study

autophagic flux. Therefore, we decided to evaluate, by immunobloting, cellular p62

level after DS treatment versus control in the clearance assay. We found that p62

level was significantly decreased in DS treated samples when compared to the

88

control, suggesting its more efficient degradation within autophagy pathway (Figure

16, p62). This result supports the idea that DS is enhancing the clearance of the

EGFP-12xQ/N aggregates through the autophagy.

F igure 16.DS enhances the clearance of E G FP-12xQ /N aggregates through the

autophagy. Stable HEK 293 cells were induced for EGFP-12xQ/N expression for 24

hours and then switched to the media without tetracycline inducer and treated with

DS 100mg/l for next 48 hours. Western blot analysis of whole cell lysates of control

and DS treated cells in the clearance assay. Antibody anti-GFP detects EGFP-

12xQ/N amount, anti-LC3 antibody reacts with two forms of LC3, upper band of 18

kDa corresponds to LC3-I and the lower band of 16 kDa to LC3-II, antibody anti-p62

reveals the cellular level of p62 protein and anti-actin antibody was used to check the

total amount of the proteins loaded on the gel. DS is causing the significant decay in

the EGFP-12xQ/N product amount after 48 hours in the clearance assay quantified

with ImageJ software. An increase of LC3-II/LC3-I and decrease of p62 level in the

cell could be related to an increase of autophagy in DS-treated cells.

89

A number of studies have identified autophagy as a crucial cellular process to avoid

accumulation of abnormal proteins in different degenerative diseases (reviewed in

(Nixon 2006)). Inhibiting autophagy using known chemical inhibitors such as

wortmanin, 3-MA, chloroquine and bafilomycin, or via silencing genes involved in

the initiation of autophagic activity was commonly used in several works linked to

degenerative diseases, in order to check autophagy direct involvement. For instance,

conditional liver-specific knock-out of Atg7,an essential autophagy gene that

functions as an E1 enzyme essential for multi-substrates such as ATG8/LC3 to

initiate the autophagy, demonstrated liver dysfunction accompanied by intracellular

accumulation of ubiquitinated protein aggregates(Komatsu et al. 2005). Indeed,

many reports agreed that autophagy is necessary in the cell in order to prevent

accumulation of ubiquitinated protein aggregates (Hara et al. 2006; Komatsu et al.

2006).

After demonstrating that DS is enhancing EGFP-12xQ/N aggregates clearance most

probably through the autophagy activation in the cell we wanted to investigate this

link in the autophagy deficient cells. We decided to inhibit the autophagy initiation

through the silencing of the one of the essential proteins for mammalian autophagy,

Atg7 (autophagy-related gene 7) using Atg7 siRNA. In the Atg7 silenced cells, the

effect of DS on EGFP-12xQ/N aggregates clearance was abolished (Figure 17, lane 1

and 2). Control cells that were treated with siRNA against luciferase maintain the DS

clearance potential (Figure 17, lane 3 and 4). With this experiment, conclusively, we

have demonstrated that DS-induced EGFP-12xQ/N clearance was abolished in an

autophagy-deficient cells confirming DS clearing potential through the autophagy

pathway.

90

F igure 17.Autophagy inhibition with siRN A against A tg7 counteracts the

clearance action of DS. Stable HEK 293 cells were induced for EGFP-12xQ/N

expression for 24 hours and then switched off by removing tetracycline. Cells were

then silenced twice within 48 hours for Atg7, an essential gene for autophagy

initiation, and treated with DS 100 mg/l for next 48 hours. Control cells were

silenced for luciferase to keep the same silencing conditions. Western blot was done

using anti-GFP antibody to detect the presence of EGFP-12xQ/N, anti-ATG7

antibody to confirm the silencing, and anti-tubulin to normalize the protein loading.

DS is enhancing EGFP-12xQ/N clearance in the cells treated with the control siRNA

against luciferase while the clearance was abolished in Atg7 silenced cells.

91

3.7. E G FP-12xQ /N aggregates are able to sequester T DP-43 wild type in the

cytoplasm

We were wondering if stable cell line HEK293 expressing EGFP-12xQ/N is able to

reproduce the same morphological and physiological response shown before for

transiently transfected cells (Budini et al. 2012).Therefore, we have decided to check

colocalization of EGFP-12xQ/N aggregates with autophagosome marker LC3, as

shown before for transiently transfected U2OS cells in the Figure 10. Here we

performed an immunofluorescence with anti-LC3 antibody in order to detect

endogenous LC3. LC3 can be seen as diffused signal throughout the cell or as clear

fluorescent dots inside the autophagosomes. We confirmed a certain degree of

colocalization between different sized aggregates (in green) and LC3 red fluorescent

dots (Figure 18), suggesting once again an autophagy response of the cell on the

EGFP-12xQ/N aggregation.

F igure 18.E G FP-12xQ /N aggregates colocalize with autophagy marker L C3 in

stable cell line.Immunofluorescence of HEK 293 T-Rex Flp-In EGFP-12xQ/N cells

upon tetracycline induction and labeling with anti-LC3 antibody, where aggregates

are present as green dots and LC3, inside autophagosomes, as red fluorescent dots.

Merged channels are showing that some EGFP-12xQ/N aggregates are inside the

autophagosomes (in yellow). Cells were analyzed on Zeiss LSM 510

92

confocalmicroscope. Images displayed are a single confocal section of 0.1 m.

Images were acquired using 100X oil immersion objective.

Another feature to investigate was the ability of EGFP-12xQ/N to recruit the wild

type TDP-43 protein in the aggregates as demonstrated before in the transfected

U2OS cells. Budini et al., have shown that 331-369 TDP-43 amino acid sequences

are able to interact with an endogenous TDP-43. In fact, they were able to see the

sequestration of endogenous and exogenous TDP-43 wild type within the EGFP-

12xQ/N aggregates with occasional, depletion of an endogenous TDP-43 from the

nucleus. Thus the pathological significance of the aggregates mainly resides in their

ability to function as a "TDP-43 sink" in the cytoplasm, where an increasing capture

of TDP-43 by growing cytoplasmic aggregates may lower the amount of TDP-43

returning to the nucleus, resulting in various loss-of-function effects(Budini et al.

2012).

Therefore, to investigate the mentioned interactions in our stable cell line, we

induced EGFP-12xQ/N expression with tetracycline and 24 hours later we performed

immunofluorescence with anti-TDP-43deltaQ/N antibody that is reacting with

endogenous TDP-43, without recognizing its C-terminal part, thus not reacting with

EGFP-12xQ/N product. On this way we were able to see EGFP-12xQ/N aggregates

as clear green dots and endogenous TDP-43 as red fluorescent signal (Figure 19A).

We were able to detect some cells with aggregates that had the ability to recruit the

wild-type TDP-43 protein (shown with an arrow in Figure 19A). However, we have

to underline that this phenomenon was not observed in all cells.

In parallel, we transfected stable cells with flag-TDP-43 wild type, induced EGFP-

12xQ/N expression with tetracycline and 24 hours later we performed

immunofluorescence with anti-flag antibody that can recognize an exogenous flag-

TDP-43. In the figure 19B, EGFP-12xQ/N aggregates are shown as green fluorescent

dots and flag-TDP-43 wild type as red fluorescent signal. Consistently with the

previously published results, in some cells the aggregates may colocalize with

overexpressed flag-TDP43 wild type (shown with an arrow in the figure 19B).

From the data presented, we confirmed the ability of EGFP-12xQ/N to sequester the

wild type TDP-43 protein to form the aggregates.

93

F igure 19.E G FP-12xQ /N aggregates may entrap endogenous T DP-43 and flag-

T DP-43 wild type.(A)EGFP-12xQ/N aggregates (green) may entrap endogenous

TDP-43 (in red) in our inducible cell line model. Immunofluorescence of HEK 293

T-Rex Flp-In EGFP-12xQ/N cells was done upon tetracycline induction and labeling

with anti-TDP-43deltaQ/N antibody. (B)HEK293 cells expressing EGFP-12xQ/N (in

green) and flag-TDP-43 wild type (in red) labeled with anti-flag antibody. Flag-TDP-

43 wild type is colocalizing with EGFP-12xQ/N aggregates in some cells (shown in

yellow in merged channels). Cells were analyzed on Zeiss LSM 510 confocal

microscope. Images displayed are a single confocal section of 0.1 µm. Images were

acquired using 100X (A) and 63X (B) oil immersion objective.

To further follow TDP-43 wild type fate in EGFP-12xQ/N aggregation model, we

performed western blots of whole cell extracts from our stable cell line with and

without tetracycline induction, and with and without flag-TDP-43 wild type co-

expression. We were able to see, for the first time, TDP-43 cleaved products (35 kDa

TDP-43 fragment, and in several instances 25 kDa truncation product) in our

aggregation model. TDP-43 truncation species have been found just in the

conditionswhere EGFP-12xQ/N expression is induced (+Tet) together with

constitutive overexpression of flag-TDP-43 wild type (Figure 20, line 4). It is

94

important to mention that the antibody used in this experiment was anti-TDP-

43deltaQ/N which does not react with C-terminal part of TDP-43 protein, thus not

recognizing EGFP-12xQ/N product. On this way we are sure that detected products

are truncation fragments solely produced from TDP-43 wild type.

C-terminal fragments of TDP-43 (35 kDa and 25 kDa) are commonly observed in the

neurons of affected patients(Neumann et al. 2006; Arai et al. 2006). It is also known

that two C-terminal fragments of TDP-43 are products of caspase-3 cleavage

action(Zhang et al. 2007). To understand if the observed TDP-43 fragments are

caspase cleaved products, we used pan caspase inhibitor Z-VAD-FMK (N-

benzyloxycarbonyl-Val-Ala-Asp-fluoromethyl ketone). The inhibitor was added to

the cell culture immediately after transfection in the concentration of 100 µM(Zhang

et al. 2007). Indeed, TDP-43 fragments were not present any more when caspase

action was inhibited in the cell (Figure 20, line 5), clearly demonstrating their

caspase cleavage nature.

On this way, we have shown that our aggregation model of TDP-43 is able to

produce C-terminal truncation products of caspase-cleaved TDP-43 in described

conditions, mimicking TDP-43 inclusions found in patients. Described observations

are giving an extra value to the EGFP-12xQ/N cell-based aggregation model of TDP-

43.

95

F igure 20.E G FP-12xQ /N aggregation model is able to produce truncation

products of T DP-43.Total cell extracts from stable HEK293 cell line without (-Tet)

or with (+Tet) tetracycline induction of EGFP-12xQ/N expression and without or

with coexpression of flag-TDP 43. In lane 4 we can appreciate 35 kDa TDP-43

fragment beside full length TDP-43 upon EGFP-12xQ/N and flag-TDP-43 wt

coexpression. Observed cleavage product was absent if the cells were treated with

pan caspase inhibitor Z-VAD-FMK (line 5). Western blot was performed with anti-

TDP-43deltaQ/N antibody reacting with TDP-43, without recognizing C-terminal

part of TDP-43 including 331-369 Q/N-rich region. Antibody anti-actin was used as

protein loading control.

3.8. DS decreases the capture of flag-T DP43 wild type in the insoluble fraction

The ability of EGFP-12xQ/N to recruit the wild type TDP-43 in the aggregates

demonstrated before (Budini et al. 2012) and here (Figure 21) was exploit to examine

if DS is able to reduce this sequestration in the clearance assay in stable EGFP-

12xQ/N expressing cell line after transfection with flag-TDP-43 wt. Clearance assay

was performed as mentioned in the previous section, and 48 hours after DS

supplementation whole cell lysates were prepared. The soluble (S) and insoluble (P)

fractions were separated as described in the section material and methods

andimmunobloted with antibodies against GFP, FLAG, Actin and Tubulin (Figure

21). First three lines correspond to the situation where EGFP-12xQ/N expression is

switched off (-Tet), while flag-TDP-43 wt is constitutively expressed. It is known

that TDP-43 has a natural tendency to aggregate and accordingly small fraction

(around 10%) of the protein flag-TDP 43 wt was found to be insoluble when EGFP-

12xQ/N expression is switched off (Flag-TDP-43, ctrl-Tet, fraction P). When EGFP-

12xQ/N expression was turned on, (ctrl+Tet) aggregates formation occurs (see

insoluble fraction P in EGFP-12xQ/N line) leading to the 2-fold change of the

insoluble flag-TDP-43 wild type portion in the cell. This result is suggesting an

increased capturing of the flag-TDP-43 wt protein in the aggregated insoluble

fraction. DS is decreasingthe amount of flag-TDP-43 wt in the P fraction (0.4-fold

change) with respect to the control, in line with the EGFP-12xQ/N aggregates

behavior (0.5-fold change). As the entrapment of TDP-43 wild type inside the

96

insoluble fraction was diminished due to the presence of DS, we hypothesize that the

enhanced clearance of TDP-43 aggregates by the DS could be a potential therapeutic

agent able to lower the capture of the TDP-43 wild type protein in the cytoplasm,

thus compensating its loss from the nucleus, as the main compartment of its normal

cellular function. On this way, we are proposing the scenario where by increasing the

efficacy of the main degradation pathway for TDP-43 aggregates with low doses of

DS 5000 Da, we may decrease TDP-43 wild type entrapment in the cytoplasm and

thus restore the nuclear loss of TPD-43 level and function in TDP-43

proteinopathies.

F igure 21. DS decreases the capture of flag-T DP43 wild type in the insoluble

fraction.Stable HEK 293 cells were transfected with flag-TDP-43 wt, induced for

EGFP-12xQ/N expression for 24 hours (+Tet) and then switched to the media

without tetracycline inducer and treated with DS 100 mg/l for next 48 hours. Whole

cell lysates were prepared (inputs IN) and fractioned in soluble (S) and insoluble (P)

fractions and the presence of EGFP-12xQ/N was detected with antibody anti-GFP

and flag-TDP-43 with anti-FLAG antibody in the Western blot. Actin and tubulin

levels were used as loading controls. In the cells with no EGFP-12xQ/N aggregates

(first three lines) flag-TDP-43 wild type is present in the insoluble form (P) in around

10%. Under EGFP-12xQ/N expression, aggregates are formed (insoluble fraction P)

97

and 2-fold change of insoluble flag-TDP-43 wt is observed. DS decreases the

insoluble portion of EGFP-12xQ/N (0.5-fold change), as well as flag-TDP-43 wt

(0.4-fold change).

Additionally, the distribution of EGFP-12xQ/N, flag-TDP-43 wt and endogenous

TDP-43 wt was investigated in the clearance assay using biochemical fractionation

of the cytoplasm and nucleus. It was observed that EGFP-12xQ/N product decays in

the cytoplasmic fraction (C) in DS treated cells versus control consistent with the

former result (Figure 22, anti GFP). As TDP-43 is predominantly nuclear, we were

able to detect flag-TDP-43 wt just in the nuclear fractions, using antibody anti-FLAG

under the described conditions. Despite no flag-TDP-43 was detected in the

cytoplasm wewere able to detect the higher level of flag-TDP-43 in the nuclear

fraction of DS treated cells comparing to the control, probably indicating the higher

rescue from the aggregate sequestration in the cytoplasm (Figure 22, anti FLAG).

The same pattern was observed with antibody anti TDP-43deltaQ/N that can

recognize an endogenous TDP-43 wt and flag-TDP-43 wt together with cleavage

fragments (TDP 35 kDa and 25 kDa), but not EGFP-12xQ/N product. Once again,

we succeeded to detect cleaved TDP fragments of 35 kDa and 25 kDa in our

aggregation model as demonstrated before in the figure 20 (Figure 22, anti

TDP43deltaQ/N). The amount of cytoplasmic TDP cleaved fragments is slightly

decreased (after normalization with tubulin level) while the amount of nuclear TDP-

43 is slightly increased upon DS treatment suggestinglower TDP wt cytoplasmic

retention, strengthening the action of DS as the clearance agent. To exclude an

eventual nuclear or cytoplasmic contamination during fractionation, we included

cytoplasm-specific (tubulin) and nucleus-specific controls (p84) in the immunoblots.

98

F igure 22. DS enhances E G FP-12xQ /N aggregates clearance and decreases

T DP-43 wild type retention in the cytoplasm. Stable HEK 293 cells were

transfected with flag-TDP-43 wt, induced for EGFP-12xQ/N expression for 24 hours

and then switched to the media without tetracycline and treated with DS 100 mg/l for

next 48 hours. Biochemical fractionation of the cytoplasm and nucleus was done as

described in the material and methods and the presence of EGFP-12xQ/N was

detected with antibody anti-GFP, flag-TDP-43 with anti-FLAG antibody and anti-

TDP-43deltaQ/N antibody was used to detect TDP-43 wild type in the Western blot.

Tubulin and p84 were used as cytoplasm and nucleus-specific proteins. EGFP-

12xQ/N product decays upon DS treatment in the cytoplasm (C), while flag-TDP-43

wt amount increases in the nucleus. After normalization with tubulin level, slight

decrease of TDP-43 and its cleavage products of 35 and 25 kDa, in the cytoplasmic

fraction (C), was observed in DS treated cells in comparison with the control.

99

3.9. DS-induced E G FP-12xQ /N clearance requires H D A C6

HDAC6 (Histone deacetylase 6) is an unusual microtubule-associated deacetylase

characterized by cytoplasmic localization, ubiquitin‑binding capacity, and impact on

tubulin and actin cytoskeleton(Boyault et al. 2007). HDAC6 has been associated with

autophagic degradation of aggregated proteins (Kawaguchi et al. 2003) and was shown

to rescue neurodegeneration in a fly model of spinal muscular atrophy where

accumulation of ubiquitin-positive protein aggregates in neurons occurs(Pandey et al.

2007). Since the discovery of an ubiquitin-binding domain in HDAC6(Verdel &

Khochbin 1999), it has been shown to participate in cellular functions through protein

ubiquitination and decision of the fate of polyubiquitinated proteins. Furthermore,

HDAC6 has been associated with maturation of autophagosomes and knockdown of

HDAC6 results in autophagosomes and protein-prone aggregates accumulation(Lee et

al. 2010). Therefore, HDAC6 putatively has an important role in proteinopathies.

Besides, it was also shown that TDP-43 binds to HDAC6 mRNA and the knockdown

of TDP-43 destabilizes HDAC6 mRNA and leads to downregulation of HDAC6

expression (Fiesel et al. 2010). Taking into the consideration HDAC6 involvement in

the turnover of aggregation‑prone proteins associated with neurodegenerative diseases

we decided to investigate whether DS is promoting autophagy through the HDAC6.

First of all, we checked HDAC6 protein level in the cells where EGFP-12xQ/N

expression was induced for 24 hours and then switched off by removing the

tetracycline from the media. Then the cells were treated with 100mg/l of DS for next

48 hours. We could not observe any difference in HDAC6 protein expression level

upon dextran sulfate treatment compared to the control (Figure 23, lane 1 and lane 3,

anti-HDAC6). It is known that HDAC6 is able to deacetylate several substrates and

the best characterized one is alpha-tubulin (Zhang et al. 2003). This catalytic property

of HDAC6 enabled us to evaluate indirectly HDAC6 activity under DS treatment and

we decided to check the acetylation level of tubulin in DS treated cells versus control.

In fact, figure 23 shows that upon DS treatment acetylated tubulin was significantly

decreased in comparison to the control (Lane 1 and lane 3, anti ac-TUB). These

findings support the idea of the DS action through the enhanced HDAC6 deacetylase

activity and decreased tubulin acetylation. The field had benefitted enormously from

the early identification of chemical compounds, such as Trichostatin A (TSA), which

100

inhibit the histone deacetylases and thus led to hyperacetylation of histones and

tubulin. We decided to treat the cells that form EGFP-12xQ/N aggregates with TSA

together with DS and to look at the acetylation level of tubulin in the Western blot. We

have demonstrated that TSA treatment reversed the effect of DS on tubulin acetylation

level in the cell (Figure 23, compare lane 4 with lane 3, antiac-tubulin). On this way,

we gave an evidence for the link between the DS action and an enhanced HDAC 6

activity in the cells with TDP-43 aggregates.

F igure 23.DS is decreasing acetylated tubulin level in H E K293 cells forming

E G FP-12xQ /N aggregates. HEK 293 cells were induced for EGFP-12xQ/N

expression for 24 hours with tetracycline and then treated with 100mg/l of DS in the

absence of presence of the HDAC6 inhibitor Trichostatin A (TSA). Biochemical

analysis of the whole cell extracts by Western blot using anti-HDAC6 and anti-

acetylated tubulin antibodies is demonstrating significantly lower level of acetylated

tubulin in DS treated cells (DS) in comparison with the control (ctrl). The effect was

abolished when HDAC6 activity was inhibited with TSA. Tubulin was used as a

control of equal protein loading.

Furthermore, and in order to confirm the hypothesis that DS action on EGFP-12xQ/N

enhanced clearance is HDAC6 dependent, we knockdown HDAC6 from the cells

with specific siRNA. In the Figure 24, we have demonstrated that the DS induced

EGFP-12xQ/N product decay in clearance assay (Figure 24, lane 3 versus lane 1,

anti-GFP) was blocked in the HDAC6-deficient cells (lane 4 DS+siHDAC6 vs lane 3

DS-siHDAC6, anti-GFP). Control cells were mock silenced with siRNA against

Luciferase in order to maintain the same conditions of silencing. An efficient

101

knockdown of HDAC6 was confirmed with antibody anti-HDAC6 in the Western

blot (Line 2 and line 4, anti-HDAC6) and the concomitant accumulation of the

HDAC6 substrate acetyl‑tubulin (Line 2 and line 4, anti-ac TUB) due to HDAC6

silencing was comparable to the effects achieved with the inhibitor TSA in the

previous experiment.

Next, to confirm that the DS clearance action depends on the deacetylase activity of

HDAC6, we decided to perform an adding back experiments. We reintroduced wild-

type HDAC6 (HDAC6-WT) and catalytically inactive mutant HDAC6-DC

(deacetylase-deficient mutant) into the silenced cells with expression plasmids

containing the silent mutations in the sequences targeted by the HDAC6 siRNA in

order to protect them from degradation As shown in the Figure 24, reintroduction of

HDAC6 wild type in the HDAC6 silenced cells led to a significant EGFP-12xQ/N

aggregates clearance upon DS treatment compared to its respective control (line 6 vs

line 5, anti-GFP), while catalytically inactive HDAC6 mutant had no effect (line 8 vs

line 7, anti-GFP). We have to mention that antibody anti-HDAC6 was not able to

recognize HDAC6-DC mutant protein in the Western blot, so the expression level of

the same was confirmed with the antibody anti-FLAG (line 7 and line 8, anti-FLAG).

Here we have identified the histone deacetylase-6 (HDAC6), as an effector of the DS

clearance action in our TDP-43 aggregation model. This finding supports the

conclusion that the observed DS induced EGFP-12xQ/N product decay is mediated

by HDAC6. Indeed, it seems that DS-induced autophagic degradation of EGFP-

12xQ/N aggregates requires catalytically active deacetylase HDAC6.

102

F igure 24.DS action on E G FP-12xQ /N clearance is H D A C6 dependent. HEK 293

cells were induced for EGFP-12xQ/N expression for 24 hours with tetracycline,

while also silenced with siRNA against HDAC6. The next day we removed

tetracycline from the medium and silenced HDAC6 again. On the third day, we

silenced HDAC6 and transfected cells with siRNA resistant HDAC6-wt or HDAC6-

DC constructs and six hours post transfection we added 100 mg/l of dextran sulfate

into the media. We harvested the cells on day four and prepared whole cell lysates.

Mock cells were silenced with siRNA against luciferase to maintain the same

silencing conditions. Cell lysates were Western blotted and probed with antibodies

against GFP, HDAC6, Flag, acetyl-tubulin and total tubulin. DS induced EGFP-

12xQ/N clearance was abolished with HDAC6 silencing and restored again with

overexpression of HDAC6 wild type, but not with catalytically inactive mutant

HDAC6-DC.

103

3.10. Mouse motor neuronal-like cell line NSC-34 model of E G FP-12xQ /N

aggregation

Amyotrophic lateral sclerosis disease is a late-onset progressive neurodegenerative

disease affecting motor neurons. We wanted to see if the aggregation pattern of EGFP-

12xQ/N would be maintained in the motor neuronal cells in vitro. NSC-34 is a hybrid

cell line produced by fusion of motor neuron enriched, embryonic mouse spinal cord

cells with mouse neuroblastoma, therefore suitable for this study. We decided to assess

the expression of EGFP-12xQ/N in NSC-34 cells and to check whether it is possible to

see GFP-12xQ/N aggregates sequestered inside autophagosomes. Therefore, we co-

transfected NSC-34 cells, after 48 hours of differentiation (cell culture medium

conditions change from 5% to 1% of serum), with EGFP-12xQ/N and HcRed-hLC3

plasmids. Cells were fixed and visualized under confocal microscope. We were able to

see some cells where EGFP-12/Q/N aggregates were sequestered inside the

autophagosomes (Figure 25) in agreement with the results obtained before in U2OS

and HEK 293 cells. On this way we tested the motor neuron-like cell model of EGFP-

12xQ/N aggregation that will be useful for further evaluation of DS suitability as a

potential anti-aggregation agent in neurons.

F igure 25. E G FP-12xQ /N aggregates are co-localizing with autophagosome

marker L C3. NSC-34 cells were co-transfected with pEGFP-12xQ/N (in green) and

HcRed-hLC3 (in red). Merge channel shows a clear co-localization of the

autophagosomal marker LC3 with aggregates (merge in yellow). Nuclei are shown

with dotted lines. Images were acquired using Zeiss LSM 510 confocal microscope

under 63X oil immersion objective.

104

3.12. Dextran sulfate 5000 Da is not toxic for the motor neuronal-like cell line

NSC-34

In vitro models are valuable in assessing of chemicals toxicity to the neuronal cells.

In the literature, NSC-34 cells were proposed as a good model for neurotoxicity

testing(Durham et al. 1993). To assess the toxicity of different compounds many

endpoints can be used, such as cell viability, cell morphology, cell proliferation and

membrane damage. The most standardized and validated are tests that assess cell

viability as enzyme release, neutral red uptake and MTT reduction assay(Balls &

Horner 1985). We have decided, by using the MTT reduction assay as an indicator of

cell redox activity, to measure cytotoxicity or viable cells loss on NSC-34 cells after

dextran sulfate 5000 Da addition to the medium. On this way we wanted to test if DS

usage as a potential anti aggregation drug on neurons is not causing the cellular

death.

NSC-34 cells were plated without or with EGFP-12xQ/N expression plasmid.

Dextran sulfate 100 mg/l was added into the cell culture media and the number of

viable cells was assessed after 48 hours of dextran sulfate supplementation. MTT

solution was added into the cell media and four hours later the reaction was stopped

as described in the material and methods. MTT reduction was determined in

spectrophotometer and presented as a percentage of the control (100%). In three

independent experiments, DS in concentration of 100 mg/l did not show any toxicity

in NSC-34 cells in either the absence (Figure 26, panel A) or presence of EGFP-

12xQ/N aggregates (Figure 26, panel B). The data reported here are supporting the

DS safe-use for the motor neuronal-like cell line and might have a broad therapeutic

potential in human TDP-43 proteinopathies, as a potential non-toxic

pharmacologically active compound able to reduce the aggregation of TDP-43.

105

F igure 26.Dextran sulfate 5000 Da is not toxic for the neuronal-like cell line

NSC-34.(A) NSC-34 cells were seeded in the 96-well plate and six hours later

dextran sulfate 100 mg/l was added to the cell culture media (DS). 48 hours later

MTT assay was performed and the reduction of MTT was determined by multiwall

scanning spectrophotometer. No toxicity in NSC-34 cells was detected by DS

treatment in comparison with the control (ctrl). Results are meanSD of three

independent experiments expressed as a percentage of the control (100%). (B) NSC-

34 cells were transfected with EGFP-12xQ/N construct in 6 well-plate, 24 hours later

the cells were transferred to 96-well plate and six hours later 100 mg/l of DS was

added (DS). No toxicity in EGFP-12xQ/N transfected NSC-34 cells was detected by

DS treatment in comparison with the control (ctrl). Results are meanSD of three

independent experiments expressed as a percentage of the control (100%).

106

4. DISC USSI O N

The presence of abnormal protein aggregates or inclusions containing specific

misfolded proteins is a commun feature in the nervous system of most

neurodegenerative diseased patients such as frontotemporal lobar degeneration with

ubiquitinated inclusions (FTLD-U), amyotrophic lateral sclerosis (ALS), Alzheimer

disease (AD), Parkinson disease (PD), and Huntington disease (HD). The underlying

pathogenic mechanism involves the misfolding of normally soluble proteins into β-

pleated conformation and the deposition of the misfolded proteins into insoluble

fibrillar aggregates with the physical and chemical properties of amyloid. For

example, most of these neurodegenerative disease-specific aggregates (as

exemplified by tangles and plaques in AD and Lewy bodies in PD) are filamentous

and display the ultrastructural and dye-binding properties of amyloids. However,

inclusions in FTLD-U/ALS are not detected by amyloid-binding dyes, suggesting

that FTLD-U/ALS might be a unique neurodegenerative proteinopathy characterized

by protein misfolding in the absence of brain amyloidosis,reviewed by (Kwong et al.

2007). In FTLD-U/ALS patients inclusions, transactive response DNA-binding

protein 43 (TDP-43) was shown to be the major component of these ubiquitinated

inclusions(Neumann et al. 2006; Arai et al. 2006). The pathological samples of these

diseases, which have been termed TDP-43 proteinopathies, are characterized by

cytoplasmic and, to a much lesser extent, nuclear TDP-43-positive (+) and

ubiquitinated inclusions (UBIs) containing full-length TDP-43, polyubiquinated

TDP-43, phosphorylated TDP-43, as well as 35- and 25-kDa carboxyl fragments of

TDP-43, for reviews, see refs. (Chen-Plotkin et al. 2010; Arai et al. 2006; Neumann

et al. 2006; Neumann et al. 2009; Da Cruz & Cleveland 2011). It has been estimated

that ∼50% of FTLD-U and 80–90% of ALS are TDP-43 (+) (Logroscino et al.

2010).

TDP-43 is a 43-kDa, ubiquitously expressed protein, well conserved among

eukaryotes(Wang et al. 2004). This DNA/RNA-binding factor is predominantly

located in the nucleus(Kuo et al. 2009), and it has been implicated in multiple

cellular functions, e.g., transcriptional repression, splicing, and translation(Buratti &

107

Baralle 2008), (Wang et al. 2008), (Lagier-Tourenne et al. 2010), (Polymenidou et al.

2011).

To date, numerous experimental data have suggested that misregulation of the

metabolism of TDP-43 leading to the accumulation of TDP-43 in the cytoplasm

including the formation of TDP-43 (+) UBIs, plays a causative role in the

pathogenesis (for reviews, see refs.(Wang et al. 2008; Chen-Plotkin et al. 2010;

Kwong et al. 2007). Evidence that the dysregulation of TDP-43 can cause disease

comes from many animal models. A neurodegenerative ALS-like phenotype can be

generated by either overexpression or knockdown of TDP-43 in flies, fish and

mice(Kabashi et al. 2010), (Li et al. 2010);(Schmid et al. 2013). These publications

are pointing out that overexpression of both mutant and wild type TDP-43 can cause

a neurodegenerative phenotype thus supporting a gain-of-function mechanism (Da

Cruz & Cleveland 2011). Loss-of-function models generated in different species

showed neurological and neurodegenerative phenotypes as well (Feiguin et al. 2009;

Lu et al. 2009; Li et al. 2010; Estes et al. 2011; Kabashi et al. 2010; Zhang et al.

2012). However, the degenerative phenotypes in the loss-of-function models appear

less overwhelming than the overexpression models and are often difficult to separate

the developmental effectsfrom a lack of TDP-43 function. Importantly, there is a lack

of evidence in mammalian models that a loss of TDP-43 function causes

neurodegeneration. This is largely due to the failure in generating such a model using

a gene knockout approach. In ALS patients carrying TDP-43 mutations, TDP-43 was

found in inclusions of their spinal cords and brains. It is unclear whether TDP-43

mutations lead to motor neuron loss through a gain of toxic properties or a loss of

TDP-43 normal function. This loss of function is proposed to be due to its

sequestration in nuclear or cytoplasmic inclusions and the corresponding disruption

of its interactions with protein or RNA targets.

The accumulation of abnormally modified TDP-43 in FTLD-U and ALS patients

suggests that dysfunction in the intracellular quality control systems in clearing the

abnormal aggregated proteins may be involved in the disease pathogenesis.Previous

studies provide evidences, finding that wild-type TDP-43 either is degraded by the

UPS (Wang et al. 2010; Zhang et al. 2010) or by a synergistic effect between both

UPS and autophagy (Urushitani et al. 2010). Recently, Scotter and colleagues have

shown that inhibiting UPS and autophagy have a strong effect on degradation,

108

localization and mobility of soluble and insoluble TDP-43. In their study, it was

suggested that soluble TDP-43 is degraded primarily by the UPS, whereas the

clearance of aggregated TDP-43 requires autophagy(Scotter et al. 2014).

Consistently, a very recent work of Huang et al.(Huang et al. 2014), have shown that

the endogenous and exogenously over-expressed TDP-43 are degraded not only by

ubiquitin proteasome system (UPS) and macroautophagy (MA), but also by the

chaperone-mediated autophagy (CMA) mediated through interaction between (Heat

shock conjugate 70) Hsc70 and ubiquitinated TDP-43(Huang et al. 2014).Although,

it is still unclear whether aggregation of TDP-43 is a primary event in ALS

pathogenesis or whether it is by-product of the disease process.

The misregulation of TDP-43 in both FTLD-U and ALS is age-dependent. This

indicates that one or more age-related changes in brain function may act as a co-

trigger and induce TDP-43 accumulation. Age is a major risk factor for the prevalent

of many neurodegenerative diseases. In fact, ageing affects many cellular

processes that predispose to neurodegeneration, and age-related changes in

cellular function predispose to the pathogenesis of ALS/FTLD in one hand,

and on the other hand a number of studies suggest a role of autophagy in

aging. Interestingly, protein degradation and specifically via autophagy (both

macroautophagy and chaperone mediated autophagy CMA) decline with age,

in the central nervous system (Martinez-Vicente et al. 2008; Ward 2002; Cuervo et

al. 2005). Age related decline of Atg genes has been shown in D . melanogaster,

while Atg8 overexpression increases the fly's lifespan (Simonsen et al. 2008),

whereas silencing autophagy genes in C . elegans leads to decreased lifespan (Hars et

al. 2007). In ALS, it was reported that lithium, at least in part by inducing autophagy,

delayed the progression of amyotrophic lateral sclerosis (ALS) in 44 human patients

affected by ALS (Fornai, Longone, Cafaro, et al. 2008), all lithium treated patients

were alive until the end of the clinical trial follow up, while 30% of the Riluzole

treated patients died before. It is tempting to speculate that the age-dependent decline

in autophagy function may contribute to TDP-43 accumulation, as it may alter the

balance between production and degradation of TDP-43 and its C-terminal

fragments.

It was demonstrated that autophagy degrade both TDP-43 and its pathological

fragment TDP-25. In particular, it was found that TDP-25 more than TDP-43 , is

109

greatly increased in cells treated with autophagy inhibitors, and it decreases in cells

treated with an autophagy enhancer(Wang et al. 2010), suggesting that autophagy is

primarily involved in clearing the pathological forms of TDP-43. Moreover, TDP-25

was found to colocalize with punctuated LC3 in transfected cells(Wang et al. 2010),

thus suggesting that TDP-25 is more specifically regulated by autophagy, and Huang

et al, very recently have also demonstrated that TDP-25/TDP-35 fragments are the

main substrates for autophagy degradation(Huang et al. 2014).

The macro-autophagy pathway (referred to as autophagy) is the main route to the

degradation of long-lived proteins, damaged organelles, and large protein aggregates.

Although autophagy was initially thought to operate mainly as a central mechanism

for survival under conditions of nutrient starvation, studies using brain-specific

knock out mice for autophagy specific genes (i.e., Atg7or Atg5) show a key role of

autophagy in maintaining and balancing protein homeostasis in neurons (Hara et al.

2006; Komatsu et al. 2006) In neurons, macroautophagy is the only cellular

mechanism capable of degrading expired organelles (reviewed by (Jaeger & Wyss-

Coray 2009)). Moreover, autophagy is a potential mechanism of clearance for protein

aggregates that occur frequently in aging neurons. Moreover, in the normal brain

autophagosome numbers (Nixon et al. 2005) and the levels of LC3-II protein (Yu et

al. 2005; Pickford et al. 2008) are low when compared with other tissues.

Nevertheless, recent findings show that autophagy in neurons is indeed constitutively

active (Hara et al. 2006; Komatsu et al. 2006) and autophagosomes accumulate

rapidly when their clearance is blocked (Boland et al. 2008), indicating the fast basal

turnover in normal neurons.

Distinct steps of the autophagic pathway can be altered in variety of

neurodegenerative diseases and possible linked to neuronal death. The different

alterations linked to neurodegeneration affecting autophagic flux including reduced

autophagy induction, altered cargo recognition, inefficient autophagosome/lysosome

fusion, and/or ineffective autophagosome clearance. In Alzheimer disease, it was

reported that autophagosome maturation was impaired (Yu et al. 2005). In

Huntington disorder, macroautophagy was shown to be impaired in early stage of the

disease (Pasquali et al. 2009). Autophagy progression seems to be impaired in ALS

patients, as sized autophagosomes and autolysosomes were found inside diseased

neurons of the ALS patients(Sasaki 2011), in agreement with the similar findings in

110

the animal model of ALS(Li, Liang; Zhang, Xiaojie; Le 2008). There is a solid

evidence that links directly a defective autophagy to ALS. This was for the first time

demonstrated by Fornai et al. (Fornai, Longone, Cafaro, et al. 2008; Fornai,

Longone, Ferrucci, et al. 2008), in SOD1 mice model aggregates and firmly

confirmed by other subsequent work (Laird et al. 2008). In fact, autophagy inhibitors

worsen the viability of motor neurons in a variety of ALS models while the

stimulation of autophagy alleviates motor neuron degeneration(Wang et al. 2012).

Autophagy inducers such as trehalose (Gomes et al. 2010), rapamycin (Berger et al.

2006),SMERs(Sarkar et al. 2007), lithium(Fornai, Longone, Cafaro, et al. 2008), and

resveratrol (Kim et al. 2007) have shown beneficial effects by decreasing protein

aggregation through the autophagy pathway activation and promoting neuronal

survival in different neurodegenerative diseases. Therefore, it is reasonable to expect

that tuning autophagy to enhance autophagy mediated clearance may have an actual

therapeutic impact to alleviate the central cause of degeneration: accumulation of

protein aggregates.

Yet, one has to take into consideration that induction of autophagy might also show

to be non beneficial. Indeed, Zhang et al., have shown that in the SOD1G93A mice,

treatment with autophagy enhancer rapamycin accelerates the MNs degeneration,

shortens the life span of the ALS mice, and did not have any positive effects on the

reduction of SOD1 aggregates (Zhang et al. 2011). Rapamycin treatment augments

motor neuron degeneration in SOD1G93A mouse model of amyotrophic lateral

sclerosis.). Since this negative effect was reported up to now just on rapamycin,

another work could explain why this effect was negatively affecting SOD1

aggregates clearance. In fact, rapamycin could suppress protective immune responses

while enhancing protective autophagy reactions during the ALS disease process.

Their results indicate that maximal therapeutic benefit may be achieved through the

use of compounds that enhance autophagy without causing immune

suppression(Staats et al. 2013).

Interestingly, it was shown that autophagy itself is regulated by TDP-43. Indeed,

Bose et al, have shown that TDP-43 functions as maintenance factor of the

autophagy system. Moreover, they have shown that the depletion of TDP-43 with the

consequent loss of the Atg7 mRNA/ATG7 protein causes impairment of the

111

autophagy and facilitates the accumulation of TDP-43 inclusions (Bose et al.

2011)(Tdp- et al. 2011).

Taking all this facts together, we have decided to investigate the autophagy role in

TDP-43 aggregation model previously established in the lab (Budini et al. 2012),

where 12 repetitions of TDP-43 amino-acid sequence 331-369 are able to induce

cytoplasmic aggregates formation. It has been previously shown by Budini et al that

the EGFP-12xQ/N aggregates are able to sequester full-length TDP-43 wild type and

are showing ubiquitinated and phosphorylated features. Thus the first thing to

investigate was if these TDP-43 inclusions may colocalize with some markers of

autophagosome-lysosome pathway. We have shown that EGFP-12xQ/N inclusions

were able to colocalize with some players of the autophagy-lysosome pathway in

transfected non-neuronal cells U2OS and mouse motor neuron-like cells NSC-34.

This observation of colocalization of EGFP-12xQ/N inclusions with autophagic

markers LC3 and p62 (sequestosome-1 (SQSTM1) (Figure 10 Panel A, B, and C)

and with lysosomal marker Lamp1 (Figure 10 Panel D) is in aggreement with the

data published by Sasaki et al, where the round bodies and skein-like inclusions from

degenerated motor neurons of ALS patients were immunostained for LC3 and for

p62(Sasaki 2011). P62 is one of the first proteins recognized as an adaptor for

delivering cargo marked by ubiquitination to the autophagic organelles(Bjørkøy et al.

2005). P62 binds ubiquitin and polyubiquitinated proteins via its C-terminal UBA

domain and it bridges the cargo and autophagic machinery by binding directly to the

autophagic effector protein LC3 Several other studies have shown an evidence that

neuronal inclusions were p62-positive and that p62 binds directly to TDP-43 in

brains of FTLD patients with TDP-43 inclusions, and furtheremore the aggregation

of TDP-43 C-terminal fragments was reported to be regulated by p62

phosphorylation through autophagy and proteasome-mediated degradation pathways

(King et al. 2011; Tanji et al. 2012; Al-Sarraj et al. 2011; Brady et al. 2011). Hence,

these results suggest that autophagy-endolysosomal pathway markers LC3, Lamp 1

and the adaptor protein p62 label EGFP-12xQ/N aggregates, probably targeting them

for degradation through the autophagy-lysosome pathway.

Inducing autophagy-lysosome pathway has proved to be an useful mechanism

against protein aggregation in different neurodegenerative diseases such as

huntington disorder (Qin et al. 2003). One of the well-known inducer of autophagy,

112

rapamycin has been used to induce TDP-43 clearance and the correction of the TDP-

43 mislocalization through the autophagy in ALS and FTD models(Wang et al. 2012;

Caccamo et al. 2009). Therefore, there is still a great interest and need to investigate

possible drug candidates able to reduce TDP-43 aggregates. We have previously

shown in our lab that the sulfated polyanion dextran sulfate 5000 Da (DS) in low

concentration favors autophagy (Menvielle et al. 2013).Therefore, we have tested if

DS could help the cell to clear the aggregates in our TDP-43 aggregation model. We

have shown that upon DS treatment, less cells were possessing dot-like structure

inside the cytoplasm and more diffused GFP signal in comparison with the control

(Figure 11A and B). This result encouraged us to investigate DS-related aggregates

LC3-labelling. When U2OS cells were co-transfected with both EGFP-12xQ/N and

HcRed-hLC3 plasmids we noted that most of DS-treated cells with EGFP-12xQ/N

aggregates were LC3 positive (cells containing more than 8 LC3 red-dots) (Figure

12A). Moreover, DS treatment increased colocalization of EGFP-12xQ/N and red-

LC3 compared with control (Figure12B and 12C). This result is suggesting that there

was more LC3 recruitment (autophagosomes) to the EGFP-12xQ/N aggregates in DS

treated cells when compared to the control.

In order to obtain more consistent and regulated expression system we then

developed a stable HEK293 cell line expressing EGFP-12xQ/N under tetracycline

inducible promoter. We have demonstrated that our inducible stable cell system is

able to reproduce the dot-like EGFP-12xQ/N inclusions in the cytoplasm (Figure 13).

Next, we confirmed that DS increased the EGFP-12xQ/N product clearance as the

product amount decays when its synthesis is stopped in comparison to the control.In

addition as mentioned in the Results section we have observed significantvariability

of the aggregate clearance efficiency using different DS batches andcurrent

experiments are trying to pin down the cause.(Figure 14A, 14B, and 15A). We

further measured the effect of DS on EGFP-12xQ/N insoluble fraction by filter-trap

assay, where insoluble aggregates are not able to pass through the membrane unlike

the soluble GFP used as a negative control (Figure15B). DS-treated sample showed

aggregate content reduction of 0.6-fold in comparison to the control (Figure 15B,

GFP line) giving the confirmation of the DS induces aggregates clearance.

As the EGFP-12xQ/N aggregates were shown to be labelled with autophagosome-

lysosome markers, we aimed at investigation if autophagy is involved in enhanced

113

aggregates clearance upon DS addition. We checked the autophagy marker LC3-I to

LC3-II conversion, as an indicator of the autophagic activity in DS treated cells in

comparison to the untreated ones. In fact, we observed an increase of the autophagic

response upon DS treatment (Figure 16, LC3-I/II). As mentioned before, another

important player in autophagy is p62/SQSTM1, an aggregate bound protein that

recruits the autophagy machinery to the aggregates(Myeku & Figueiredo-Pereira

2011). Moreover, p62 is degraded by autophagy whereas, it is found accumulated

when the autophagic flux is inhibited. P62 level was significantly decreased in DS

samples compared to the control (Figure 16, p62). Taken together our data supports

the idea that DS plays an active role in the clearance of TDP-43 aggregates through

the adaptor protein p62 within the autophagy.

Suppression of the autophagy promoting genes such as Atg7 or Atg5, could be a tool

to check the direct involvement of autophagy in the DS enhenced EGFP-12xQ/N

clearance. For this purpose, we silenced one of the essential proteins for mammalian

autophagy, autophagy related gene 7 (Atg7), and we were able to abolish the DS

clearance action, compared to the respective control (siAtg7 ctrl in Figure 17) giving

the final evidence of the autophagy related DS clearance action.

As mentioned before, “loss of function” theory is putting the TDP-43 nuclear

depletion (as the consequence of TDP-43 retention, aggregation in the cytoplasm) in

the first line for the cause of the neuronal death in TDP-43 proteinopathies. Budini et

al. have shown before that some EGFP-12xQ/N aggregates are able to colocalize

with endogenous TDP-43 wild type suggesting the sequestration of the wild type

protein within the aggregates. In agreement with this, we were able to show

overlapping of the EGFP-12xQ/N signal with the endogenous TDP-43 (Figure 19A)

as well as with the overexpressed flag-TDP-43 wild type (Figure 19B) in some cells.

We therefore confirmed the ability of EGFP-12xQ/N to sequester TDP-43 wild type

in our stable inducible system. We further examined if DS treatment was able to

reduce this sequestration and to release TDP-43 wild type from the insoluble

fraction. We expressed simultaneously EGFP-12xQ/N and the exogenous flag-TDP-

43 wt in the clearance assay and separated the soluble (S) and the insoluble (P)

fractions and immunobloted with antibodies against GFP, flag, actin and tubulin. We

have shown that DS decreased the amount of flag-TDP-43 wt in the insoluble

fraction (P) compared to the control (0.4-fold change), consistenly with the EGFP-

114

12xQ/N aggregates behavior (0.5-fold change). This observation led us to propose

the model where DS treatment enhances the clearance of the aggregated EGFP-

12xQ/N through the autophagy leading to the decreased entrapment of TDP-43 wt

inside the insoluble fraction in the cell.

We further hypothesized that the enhanced clearance of TDP-43 aggregates by DS

could lower the capture of the cytoplasmic TDP-43 wild type, thus compensating its

loss from the nucleus, as TDP-43 is normaly nuclear protein. It was previously

shown the direct involvement of autophagy in rescuing the TDP-43 mislocalization

using rapamycin, suggesting that the cytoplasmic TDP-43 is normally degraded by

autophagy(Caccamo et al. 2009). In the same study it was given an in vivo evidence

that supports the hypothesis that accelerating autophagy may alleviate TDP-43-

mediated toxicity(Caccamo et al. 2009). Thus, we have decided to perform

nuclear/cytoplasmic fractionation to evaluate the level of TDP-43 wild type trapped

in the cytoplasm inuntreated and DS treated cells. We observed an enhanced TDP-43

wild type level in the nuclear fraction upon DS treatment when compared to the

control (Figure 22). Our findings support the scenario where low doses of DS 5000

Damay increase the efficacy of the main degradation pathway for TDP-43 aggregates

and decrease TDP-43 wild type sequestration in the cytoplasm. Projecting this model

to the ALS/FTD patient’s affected neurons where endogenous TDP-43 is captured by

cytoplasmic aggregates, leading to its nuclear loss of function, we suggest that DS

promoted autophagy would decrease TDP-43 sequestration in the cytoplasm and

restore normal nuclear level. This observation leads DS toward its possible

therapeutic use.

Interestingly, when we overexpressed TDP-43 wild type in our aggregation TDP-43

model, we noticed the appearance of small fragments of TDP-43 that correspond to

C-terminal fragments: 35 kDa and 25 kDa. Recent findings have shown that TDP-43

C-terminal fragments form cytoplasmic aggregates and cause cytotoxicity(Zhang et

al. 2009; Nonaka et al. 2009; Igaz et al. 2009; Johnson et al. 2008). Thus, TDP-43

truncation may play an important role in the pathogenesis of ALS, FTLD-U and

other TDP-43 proteinopathies. The degradation of TDP-43 by caspases was reported

to be responsible for generation of the C-terminal fragments due to the intrinsic

caspase cleavage sites in its primary sequence(Zhang et al. 2007; Rohn 2008; Rohn

& Kokoulina 2009). We aimed to confirm the caspase involved TDP degradation in

115

our model by the help of the pan caspase inhibitor Z-VAD (OMe)-FMK. The

generation of proteolytic fragments was inhibited by this caspase inhibitor as shown

in the Figure 20. After confirming that the fragments seen correspond to TDP-43 C-

terminal fragments, we checked if DS could decrease TDP-43 C-fragments amount

compared to the control. We have seen a slight decrease of the TDP-43 C-terminal

fragments in the cytoplasmic fraction in the DS sample compared to the control

(Figure22, anti TDP-43deltaQ/N). Yet, it is important to stand out that DS also could

have such action by decreasing caspase-3 action on TDP-43 degradation. In fact, it

was shown by Zhang et al., that the treatment with staurosporine (1µM), a potent

inducer of apoptosis and caspase-3 activator, leads to the high degradation of TDP-

43 and hence the presence of C-terminal fragments in the cytoplasm(Zhang et al.

2007). It was also reported that DS inhibits staurosporine-induced apoptosis via

decreasing the pro-caspase-9 activation, cleaved caspase-3 levels and the

cytochrome c release into the cytosol(Jing et al. 2011). Nevertheless, more

experiments are required to reveal the mechanism by which DS interferes with the

caspase action on TDP-43 or whether DS while enhancing EGFP-12xQ/N aggregates

clearance clear the C-terminal fragments too.

We further aimed to investigate the molecular bases by which DS would be acting as

an enhancer of EGFP-12xQ/N clearance through the autophagy. HDAC6 has

emerged as an important player in the cellular management of protein

aggregates(Rodriguez-Gonzalez et al. 2008). It is now well known that histone

deacetylase-6 (HDAC6) recruits and binds the polyubiquitinated misfolded protein

cargo to dynein motors for transportation to the aggresome(Kawaguchi et al. 2003), a

process required for autophagic degradation of aberrant aggregated proteins like

Huntingtin(Iwata et al. 2005). In addition, several studies showed that HDAC6

mutations reduce the transportation of autophagosomes to perinuclear regions where

lysosomes are more abundant, resulting in the impairment of autophagosome-

lysosome fusion and reduced clearance of aggregate-prone proteins(Lee et al. 2010).

HDAC6 also colocalizes with α-syn and ubiquitin in the Lewy bodies of PD and

Dementia with Lewy Bodies (Kawaguchi et al. 2003; Boyault et al. 2006). Therefore,

HDAC6 may be a key regulator in the cellular clearance of abnormal proteins.

Here,we sought to investigate whether HDAC6 regulates the aggresome-autophagic

clearance of EGFP-12xQ/N and if DS may interfere with HDAC6. Despite the fact

116

that no HDAC6 protein expression level difference upon DS treatment in comparison

with the control was noticed (Figure 24, lane 1 and lane 3, HDAC6), we decided to

check if HDAC6 activity could be altered by DS. It is known that HDAC6 works as a

tubulin deacetylase protein(Zhang et al. 2003), so we looked if the level of acetylated

tubulin was altered in the DS treated cells in comparison to the non treated control.

In fact, upon DS treatment, we could see reduction of the acetylated tubulin (figure

24, ac-tubulin) that could indicate the effect of DS on HDAC6 activity, rather than

protein expression level. Inhibiting HDAC6 with Trichostatin A (TSA) abolished the

effect of DS on tubulin acetylation (Figure 24, lane 4 and 3, anti ac-tubulin). We

found that both pharmacological inhibition of HDAC6 as well as its depletion with

specific siRNA enhance microtubule acetylation, and eliminate DS clearance effect

on EGFP-12xQ/N aggregates (shown in Figure 23). In order to confirm that DS

action on EGFP-12xQ/N clearance is HDAC6 activity dependent, we reintroduced

wild type (HDAC6-wt) or the catalytically inactivate HDAC6 (HDAC6-DC) into the

HDAC-6 silenced cells. We actually used expression plasmids with silent mutations

in the sequences targeted by the HDAC6 siRNA in order to protect them from

degradation(Kawaguchi et al. 2003). We have shown that adding back HDAC6 wt in

the silenced cells restored the DS clearance effect in the treated cells compared to its

respective control, whereas, catalytically inactive HDAC6 mutant had no effect

(Figure 24, line 7 and line 8). Our findings demonstrated that the observed DS

induced EGFP-12xQ/N product decay is mediated by HDAC6. Moreover, this DS-

mediated autophagy action requires a catalytically deacetylase activity of HDAC6. In

fact, it was suggested that HDAC6 also plays a role in the eventual clearance of the

aggresomes in an autophagy-dependent manner, implying a functional connection

between HDAC6 and autophagy(Iwata et al. 2005; Pandey et al. 2007). In addition,

for HDAC6 has been shown to promote fusion of autophagosomes and lysosomes

and facilitates clearance of protein aggregates(Lee et al. 2010). Even more

specifically HDAC6 catalytic activity has been shown to be essential for turnover of

abberant proteins by autophagy(Pandey et al. 2007).As HDAC6 seems to play a

regulatory role in the clearance of misfolded proteins, activation of HDAC6

deacetylase function could be a promising strategy in processing aberrant protein

aggregation in ALS and other proteinopathies. It seems that in general

pharmacological activation of deacetylation reactions in the cytoplasm is positively

117

correlated with the autophagy induction. Indeed, for resveratrol and a panel of

phenolic compounds was demonstrated to stimulate the deacetylation of cytoplasmic

proteins and protects against neurotoxicity in AD and ALS mouse models through

the autophagic flux promotion(Kim et al. 2007; Pietrocola et al. 2012).

Interestingly, it has been suggested that the defect in autophagosome-lysosome

fusion may be an important factor in the development of neurodegenerative disease,

giving abnormal autophagic structures found in Alzheimer’s disease patients

neurons(Nixon et al. 2005). Taking together DS clearance action through the

HDAC6 and HDAC6 involvement in the autophagosome-lysosome fusion one could

investigate a possible link of DS and the maturation step of autophagy. The construct

ptfLC3 that contains mRFP-EGFP tandem fluorescent-tagged LC3 is showing a GFP

and mRFP signal within autophagosomes, but only mRFP signal after

autophagosomes fusion with lysosomes due to the GFP lost in the acidic lysosomal

environment (Kimura et al. 2007). The number of yellow puncta (colocalization of

GFP and mRFP signals) is reflecting the number of autophagosomes per cell,

whereas the red puncta (mRFP signal) is reflecting the number of lysosomes per cell.

On this way mRFP-EGFP-LC3 reporter could be very useful to discriminate between

acidic (autolysosomes) and neutral (autophagosomes) LC3-positive vesicles. The use

of EGFP-12xQ/N construct in this assay is not possible due to the green signal

overlap with GFP fluorescence from LC3. Our preliminary data with HeLa cells

transfected with mRFP-EGFP-LC3 reporter and treated with DS are showing the

lower number of autophagosomes in favor of autolysosomes (red puncta) in the DS

treated samples in comparison to the control conditions (data not shown). This result

may suggest that autophagic turnover in the cell could be faster upon DS treatment.

This observation is leading to the interpretation where DS may help the cell to clear

TDP-43 aggregates through the more effective autophagosomes fusion to lysosomes

resulting in the more efficient cargo degradation. Further experiments are necessary

to confirm and analyze in details DS tendency toward autophagy flux promotion and

link with HDAC6, but the autophagosome-lysosome fusion stimulation, rather than

autophagy activation is doubtless an attractive therapeutic approach for

neurodegenerative disease.

On the base of our data, we are suggesting the DS as a potential therapeutic agent for

TDP-43 proteinopathies. In this respect, we have decided to investigate DS toxicity

118

on the motor neuronal-like cells. NSC-34 cell line was proposed to be a good model

for neurotoxicity testing (Durham et al. 1993). In order to assess DS toxicity, many

assays could be used, such as cell viability, cell morphology, cell proliferation and

cellular membrane damage. We decided to use the most standardized test that assess

cell viability, MTT reduction assay as an indicator of cell redox activity. NSC-34

cells viability was assessed 48 hours after DS 5000 Da (100 mg/l) addition into the

medium, in the presence or absence of EGFP-12xQ/N aggregates. In three

independent experiments, DS did not show any toxicity in NSC-34 cells (Figure 26A

and 26B). This is in the agreement with our previously published data obtained with

HeLa and CHO cells (Menvielle et al. 2013), where we showed that DS is able to

enhance the cells survival through autophagy induction and apoptosis inhibition. DS-

mediated apoptosis inhibition was also shown in staurosporine-induced apoptosis

model in the cells(Jing et al. 2011).

Dextran sulfate is a semisynthetic analog of the glycosaminoglycan family, which

includes heparin, heparan sulfate, dermatan sulfate, and chondroitin sulfate. For

some low molecular weight heparins was shown to be able to pass through the blood-

brain barrier (Leveugle et al. 1998; Ma et al. 2002) and act as neuroprotective agents

(Pérez et al. 1998; Stutzmann et al. 2002; Ma et al. 2007). Dextran sulfate 5000 Da is

not toxic for the motor neuronal-like cell line NSC-34, that together with the fact that

DS already passed the test of time for human use, experiencing over 40 years of its

consumption in Japan to treat arteriosclerosis and high cholesterol, may be

considered as a promising therapeutic agent to treat TDP-43 proteinopathies in

humans.

119

5. C O N C L USI O NS

Protein clearing systems play a key role in TDP-43 proteinopathies. In particular,

autophagy represents the main pathway which through the clearance of the misfolded

proteins removes the potentially toxic material countering to cell death. There is now

a solid evidence that links directly a defective autophagy to ALS. Many studies have

shown that increasing autophagic flux with pharmacological inducers of autophagy

could counteract the pathophysiology of ALS.

EGFP-12xQ/N model of TDP-43 aggregation is a valuable tool to investigate cellular

protein degradation responses. We have shown that some EGFP-12xQ/N aggregates

were labeled with autophagy-lysosomal pathway markers: LC3, p62, and Lamp1

targeting them for degradation through the autophagy. Next, we have demonstrated

that dextran sulfate 5000 Da was able to promote the clearance of the EGFP-12xQ/N

aggregates via autophagy stimulation. Projecting this model to the ALS patient's

affected neurons where endogenous TDP-43 is captured by cytoplasmic aggregates

leading to its nuclear loss, we have demonstrated DS-induced lower TDP-43 wild

type sequestration inside the insoluble fraction. Additionally, we observed a

reduction of acetylated tubulin in DS-treated cells through the higher HDAC6

activity, indicating the role of HDAC6 in DS-mediated TDP-43 EGFP-12xQ/N

aggregates clearance.

120

In conclusion, we propose a model where DS leads to autophagy activation through

the stimulation of HDAC6 activity. As HDAC6 seems to play a regulatory role in the

clearance of misfolded proteins, its pharmacological activation could be a promising

strategy in processing aberrant protein aggregation in ALS, as it was shown before

for Huntington's disease model(Iwata et al. 2005; Pandey et al. 2007).

Data obtained in this study leads DS toward its possible therapeutic use. We have

shown here that DS 5000 Da is not toxic per se for the motor neuronal-like cell line

NSC-34 and may be considered as a promising therapeutic agent to treat TDP-43

proteinopathies in humans.

6. BIB L I O G R APH Y Ahlberg, J. & Glaumann, H., 1985. Uptake--microautophagy--and degradation of

exogenous proteins by isolated rat liver lysosomes. Effects of pH, ATP, and inhibitors of proteolysis. Experimental and molecular pathology, 42(1), pp.78–88.

Al-Sarraj, S. et al., 2011. p62 positive, TDP-43 negative, neuronal cytoplasmic and intranuclear inclusions in the cerebellum and hippocampus define the pathology of C9orf72-linked FTLD and MND/ALS. Acta neuropathologica, 122(6), pp.691–702.

Andersen, P.M. et al., 2003. Sixteen novel mutations in the Cu/Zn superoxide dismutase gene in amyotrophic lateral sclerosis: a decade of discoveries, defects and disputes. Amyotrophic lateral sclerosis and other motor neuron disorders : official publication of the World F ederation of Neurology, Research Group on Motor Neuron Diseases, 4(2), pp.62–73.

Andersson, M.K. et al., 2008. The multifunctional FUS, EWS and TAF15 proto-oncoproteins show cell type-specific expression patterns and involvement in cell spreading and stress response. BMC cell biology, 9(1), p.37.

Aplin, A. et al., 1992. Cytoskeletal elements are required for the formation and maturation of autophagic vacuoles. Journal of cellular physiology, 152(3), pp.458–66.

Arai, T. et al., 2006. TDP-43 is a component of ubiquitin-positive tau-negative inclusions in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Biochemical and biophysical research communications, 351(3), pp.602–11.

121

Arrasate, M. et al., 2004. Inclusion body formation reduces levels of mutant huntingtin and the risk of neuronal death. Nature, 431(7010), pp.805–10.

Attaix, D. et al., 2001. Regulation of proteolysis. Current opinion in clinical nutrition and metabolic care, 4(1), pp.45–9.

Ayala, Y.M. et al., 2005. Human, Drosophila, and C.elegans TDP43: nucleic acid binding properties and splicing regulatory function. Journal of molecular biology, 348(3), pp.575–88.

Ayala, Y.M. et al., 2008. Structural determinants of the cellular localization and shuttling of TDP-43. Journal of cell science, 121(Pt 22), pp.3778–85.

Balls, M. & Horner, S.A., 1985. The FRAME interlaboratory programme on in vitro cytotoxicology. Food and chemical toxicology : an international journal published for the British Industrial Biological Research Association, 23(2), pp.209–13.

Banks, G.T. et al., 2008. TDP-43 is a culprit in human neurodegeneration, and not just an innocent bystander. Mammalian genome : official journal of the International Mammalian Genome Society, 19(5), pp.299–305.

Banz, Y. et al., 2005. Locally targeted cytoprotection with dextran sulfate attenuates experimental porcine myocardial ischaemia/reperfusion injury. European heart journal, 26(21), pp.2334–43.

Barmada, S.J. & Finkbeiner, S., 2010. Pathogenic TARDBP mutations in amyotrophic lateral sclerosis and frontotemporal dementia: disease-associated pathways. Reviews in the neurosciences, 21(4), pp.251–72.

Berger, Z. et al., 2006. Rapamycin alleviates toxicity of different aggregate-prone proteins. Human molecular genetics, 15(3), pp.433–42.

Bernales, S., McDonald, K.L. & Walter, P., 2006. Autophagy counterbalances endoplasmic reticulum expansion during the unfolded protein response. PLoS biology, 4(12), p.e423.

Bjørkøy, G. et al., 2005. p62/SQSTM1 forms protein aggregates degraded by autophagy and has a protective effect on huntingtin-induced cell death. The Journal of cell biology, 171(4), pp.603–14.

Del Bo, R. et al., 2009. TARDBP (TDP-43) sequence analysis in patients with familial and sporadic ALS: identification of two novel mutations. European journal of neurology : the official journal of the European Federation of Neurological Societies, 16(6), pp.727–32.

Boland, B. et al., 2008. Autophagy induction and autophagosome clearance in neurons: relationship to autophagic pathology in Alzheimer’s disease. The Journal of neuroscience : the official journal of the Society for Neuroscience, 28(27), pp.6926–37.

122

Borroni, B. et al., 2009. Mutation within TARDBP leads to frontotemporal dementia without motor neuron disease. Human mutation, 30(11), pp.E974–83.

Boyault, C. et al., 2007. HDAC6, at the crossroads between cytoskeleton and cell signaling by acetylation and ubiquitination. Oncogene, 26(37), pp.5468–76.

Boyault, C. et al., 2006. HDAC6-p97/VCP controlled polyubiquitin chain turnover. The EMBO journal, 25(14), pp.3357–66.

Brady, O. a et al., 2011. Regulation of TDP-43 aggregation by phosphorylation and p62/SQSTM1. Journal of neurochemistry, 116(2), pp.248–59.

Brooks, B.R. et al., 2000. El Escorial revisited: revised criteria for the diagnosis of amyotrophic lateral sclerosis. Amyotrophic lateral sclerosis and other motor neuron disorders : official publication of the World Federation of Neurology, Research Group on Motor Neuron Diseases, 1(5), pp.293–9.

Brooks, B.R., 1994. El Escorial World Federation of Neurology criteria for the diagnosis of amyotrophic lateral sclerosis. Subcommittee on Motor Neuron Diseases/Amyotrophic Lateral Sclerosis of the World Federation of Neurology Research Group on Neuromuscular Diseases and th. Journal of the neurological sciences, 124 Suppl, pp.96–107.

Bruijn, L.I. et al., 1998. Aggregation and Motor Neuron Toxicity of an ALS-Linked SOD1 Mutant Independent from Wild-Type SOD1. Science, 281(5384), pp.1851–4.

Bruijn, L.I., Miller, T.M. & Cleveland, D.W., 2004. Unraveling the mechanisms involved in motor neuron degeneration in ALS. Annual review of neuroscience, 27, pp.723–49.

Budini, M. et al., 2012. Cellular model of TAR DNA-binding protein 43 (TDP-43) aggregation based on its C-terminal Gln/Asn-rich region. The Journal of biological chemistry, 287(10), pp.7512–25.

Buratti, E. et al., 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. The Journal of biological chemistry, 280(45), pp.37572–84.

Buratti, E. & 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. The Journal of biological chemistry, 276(39), pp.36337–43.

Buratti, E. & Baralle, F.E., 2008. Multiple roles of TDP-43 in gene expression, splicing regulation, and human disease. Frontiers in bioscience : a journal and virtual library, 13, pp.867–78.

123

Caccamo, A. et al., 2010. Molecular interplay between mammalian target of rapamycin (mTOR), amyloid-beta, and Tau: effects on cognitive impairments. The Journal of biological chemistry, 285(17), pp.13107–20.

Caccamo, A. et al., 2009. Rapamycin rescues TDP-43 mislocalization and the associated low molecular mass neurofilament instability. The Journal of biological chemistry, 284(40), pp.27416–24.

Chen, Y. et al., 2011. Expression of human FUS protein in Drosophila leads to progressive neurodegeneration. Protein & cell, 2(6), pp.477–86.

Cheng, I.H. et al., 2007. Accelerating amyloid-beta fibrillization reduces oligomer levels and functional deficits in Alzheimer disease mouse models. The Journal of biological chemistry, 282(33), pp.23818–28.

Chen-Plotkin, A.S., Lee, V.M.-Y. & Trojanowski, J.Q., 2010. TAR DNA-binding protein 43 in neurodegenerative disease. Nature reviews. Neurology, 6(4), pp.211–20.

Chiang, H.L. et al., 1989. A role for a 70-kilodalton heat shock protein in lysosomal degradation of intracellular proteins. Science (New York, N.Y.), 246(4928), pp.382–5.

Chung, C.T., Niemela, S.L. & Miller, R.H., 1989. One-step preparation of competent Escherichia coli: transformation and storage of bacterial cells in the same solution. Proceedings of the National Academy of Sciences, 86(7), pp.2172–2175.

Clausen, T.H. et al., 2010. p62/SQSTM1 and ALFY interact to facilitate the formation of p62 bodies/ALIS and their degradation by autophagy. Autophagy, 6(3), pp.330–44.

Cohen, E. et al., 2009. Reduced IGF-1 signaling delays age-associated proteotoxicity in mice. Cell, 139(6), pp.1157–69.

Corrado, L. et al., 2009. High frequency of TARDBP gene mutations in Italian patients with amyotrophic lateral sclerosis. Human mutation, 30(4), pp.688–94.

Coux, O., Tanaka, K. & Goldberg, A.L., 1996. Structure and functions of the 20S and 26S proteasomes. Annual review of biochemistry, 65, pp.801–47.

Cozzolino, M., Ferri, A. & Carrì, M.T., 2008. Amyotrophic lateral sclerosis: from current developments in the laboratory to clinical implications. Antioxidants & redox signaling, 10(3), pp.405–43.

Da Cruz, S. & Cleveland, D.W., 2011. Understanding the role of TDP-43 and FUS/TLS in ALS and beyond. Current opinion in neurobiology, 21(6), pp.904–19.

124

Cuervo, A.M. et al., 2005. Autophagy and Aging: The Importance of Maintaining “Clean” Cells. Autophagy, 1(3), pp.131–141.

Cuervo, A.M., 2010. Chaperone-mediated autophagy: selectivity pays off. Trends in endocrinology and metabolism: TEM, 21(3), pp.142–50.

Cuervo, A.M. & Dice, J.F., 1996. A receptor for the selective uptake and degradation of proteins by lysosomes. Science (New York, N.Y.), 273(5274), pp.501–3.

D’Ambrogio, A. et al., 2009. Functional mapping of the interaction between TDP-43 and hnRNP A2 in vivo. Nucleic acids research, 37(12), pp.4116–26.

Daoud, H. et al., 2009. Contribution of TARDBP mutations to sporadic amyotrophic lateral sclerosis. Journal of medical genetics, 46(2), pp.112–4.

Van Deerlin, V.M. et al., 2008. TARDBP mutations in amyotrophic lateral sclerosis with TDP-43 neuropathology: a genetic and histopathological analysis. Lancet neurology, 7(5), pp.409–16.

Degenhardt, K. et al., 2006. Autophagy promotes tumor cell survival and restricts necrosis, inflammation, and tumorigenesis. Cancer cell, 10(1), pp.51–64.

Dewey, C.M. et al., 2012. TDP-43 aggregation in neurodegeneration: are stress granules the key? Brain research, 1462(null), pp.16–25.

Díaz-Hernández, M. et al., 2003. Neuronal induction of the immunoproteasome in Huntington’s disease. The Journal of neuroscience : the official journal of the Society for Neuroscience, 23(37), pp.11653–61.

Ding, W.-X. et al., 2007. Linking of autophagy to ubiquitin-proteasome system is important for the regulation of endoplasmic reticulum stress and cell viability. The American journal of pathology, 171(2), pp.513–24.

Dobson, C.M., 2003. Protein folding and misfolding. Nature, 426(6968), pp.884–90.

Dreyfuss, G. et al., 1993. hnRNP proteins and the biogenesis of mRNA. Annual review of biochemistry, 62, pp.289–321.

Dubouloz, F. et al., 2005. The TOR and EGO protein complexes orchestrate microautophagy in yeast. Molecular cell, 19(1), pp.15–26.

Dunkel, P. et al., 2012. Clinical utility of neuroprotective agents in neurodegenerative diseases: current status of drug development for Alzheimer’s, Parkinson's and Huntington's diseases, and amyotrophic lateral sclerosis. Expert opinion on investigational drugs, 21(9), pp.1267–308.

Durham, H.D., Dahrouge, S. & Cashman, N.R., 1993. Evaluation of the spinal cord neuron X neuroblastoma hybrid cell line NSC-34 as a model for neurotoxicity testing. Neurotoxicology, 14(4), pp.387–95.

125

De Duve, C. & Wattiaux, R., 1966. Functions of lysosomes. Annual review of physiology, 28, pp.435–92.

De Duve Christian, 1965. the lysosome. Scientific American, (208), pp.64–78.

Essner E. & Novikoff, A.B., 1961. Localization of acid phosphatase activity in hepatic lysosomes by means of electron microscopy. The Journal of biophysical and biochemical cytology, 9, pp.773–84.

Estes, P.S. et al., 2011. Wild-type and A315T mutant TDP-43 exert differential neurotoxicity in a Drosophila model of ALS. Human molecular genetics, 20(12), pp.2308–21.

Fass, E. et al., 2006. Microtubules support production of starvation-induced autophagosomes but not their targeting and fusion with lysosomes. The Journal of biological chemistry, 281(47), pp.36303–16.

Feiguin, F. et al., 2009. Depletion of TDP-43 affects Drosophila motoneurons terminal synapsis and locomotive behavior. F EBS letters, 583(10), pp.1586–92.

Fengsrud, M. et al., 1995. Ultrastructural and immunocytochemical characterization of autophagic vacuoles in isolated hepatocytes: effects of vinblastine and asparagine on vacuole distributions. Experimental cell research, 221(2), pp.504–19.

Fiesel, F.C. et al., 2010. Knockdown of transactive response DNA-binding protein (TDP-43) downregulates histone deacetylase 6. The EMBO journal, 29(1), pp.209–21.

Finkbeiner, S. et al., 2006. -Disease modifying pathways in neurodegeneration. The Journal of neuroscience : the official journal of the Society for Neuroscience, 26(41), pp.10349–57.

Fornai, F., Longone, P., Ferrucci, M., et al., 2008. Autophagy and amyotrophic lateral sclerosis: The multiple roles of lithium. Autophagy, 4(4), pp.527–30.

Fornai, F., Longone, P., Cafaro, L., et al., 2008. Lithium delays progression of amyotrophic lateral sclerosis. Proceedings of the National Academy of Sciences of the United States of America, 105(6), pp.2052–7.

Fuentealba, R.A. et al., 2010. Interaction with polyglutamine aggregates reveals a Q/N-rich domain in TDP-43. The Journal of biological chemistry, 285(34), pp.26304–14.

Geng, J. & Klionsky, D.J., 2008. The Atg8 and Atg12 ubiquitin-like conjugation systems in macroautophagy. “Protein modifications: beyond the usual suspects” review series. EMBO reports, 9(9), pp.859–64.

Geser, F. et al., 2009. Amyotrophic lateral sclerosis, frontotemporal dementia and beyond: the TDP-43 diseases. Journal of neurology, 256(8), pp.1205–14.

126

Gidalevitz, T. et al., 2009. Destabilizing protein polymorphisms in the genetic background direct phenotypic expression of mutant SOD1 toxicity. H. Orr, ed. PLoS genetics, 5(3), p.e1000399.

Giordana, M.T. et al., 2010. TDP-43 redistribution is an early event in sporadic amyotrophic lateral sclerosis. Brain pathology (Zurich, Switzerland), 20(2), pp.351–60.

Gitcho, M.A. et al., 2009. TARDBP 3’-UTR variant in autopsy-confirmed frontotemporal lobar degeneration with TDP-43 proteinopathy. Acta neuropathologica, 118(5), pp.633–45.

Glicksman, M.A., 2011. The preclinical discovery of amyotrophic lateral sclerosis drugs. Expert opinion on drug discovery, 6(11), pp.1127–38.

Gomes, C., Escrevente, C. & Costa, J., 2010. Mutant superoxide dismutase 1 overexpression in NSC-34 cells: effect of trehalose on aggregation, TDP-43 localization and levels of co-expressed glycoproteins. Neuroscience letters, 475(3), pp.145–9.

Gong, Y.H. et al., 2000. Restricted expression of G86R Cu/Zn superoxide dismutase in astrocytes results in astrocytosis but does not cause motoneuron degeneration. The Journal of neuroscience : the official journal of the Society for Neuroscience, 20(2), pp.660–5.

Gräßler, J. et al., 2013. Differential effects of lipoprotein apheresis by lipidfiltration or dextran sulfate adsorption on lipidomic profile. Atherosclerosis. Supplements, 14(1), pp.151–5.

Grune, T. et al., 2004. Decreased proteolysis caused by protein aggregates, inclusion bodies, plaques, lipofuscin, ceroid, and “aggresomes” during oxidative stress, aging, and disease. The international journal of biochemistry & cell biology, 36(12), pp.2519–30.

Gutierrez, M.G. et al., 2004. Rab7 is required for the normal progression of the autophagic pathway in mammalian cells. Journal of cell science, 117(Pt 13), pp.2687–97.

Habib, A.A. & Mitsumoto, H., 2011. Emerging drugs for amyotrophic lateral sclerosis. Expert opinion on emerging drugs, 16(3), pp.537–58.

Hahn, 1898. Das proteolytische Enzym des Hefepresssaftes. Berichte der Deutschen Chemischen Gesellschaft, 31, pp.200–201.

Hanson, K.A. et al., 2010. Ubiquilin modifies TDP-43 toxicity in a Drosophila model of amyotrophic lateral sclerosis (ALS). The Journal of biological chemistry, 285(15), pp.11068–72.

127

Hara, T. et al., 2008. FIP200, a ULK-interacting protein, is required for autophagosome formation in mammalian cells. The Journal of cell biology, 181(3), pp.497–510.

Hara, T. et al., 2006. Suppression of basal autophagy in neural cells causes neurodegenerative disease in mice. Nature, 441(7095), pp.885–9.

Harding, T.M., 1995. Isolation and characterization of yeast mutants in the cytoplasm to vacuole protein targeting pathway. The Journal of Cell Biology, 131(3), pp.591–602.

Hars, E.S. et al., 2007. Autophagy regulates ageing in C. elegans. Autophagy, 3(2), pp.93–5.

Haverkamp, L.J., Appel, V. & Appel, S.H., 1995. Natural history of amyotrophic lateral sclerosis in a database population. Validation of a scoring system and a model for survival prediction. Brain : a journal of neurology, 118 ( Pt 3, pp.707–19.

Hayes-Punzo, A. et al., 2012. Gonadectomy and dehydroepiandrosterone (DHEA) do not modulate disease progression in the G93A mutant SOD1 rat model of amyotrophic lateral sclerosis. Amyotrophic lateral sclerosis : official publication of the World F ederation of Neurology Research Group on Motor Neuron Diseases, 13(3), pp.311–4.

Heinemeyer, W. et al., 1991. Proteinase yscE, the yeast proteasome/multicatalytic-multifunctional proteinase: mutants unravel its function in stress induced proteolysis and uncover its necessity for cell survival. The EMBO journal, 10(3), pp.555–62.

Herman, P.K. & Emr, S.D., 1990. Characterization of VPS34, a gene required for vacuolar protein sorting and vacuole segregation in Saccharomyces cerevisiae. Molecular and cellular biology, 10(12), pp.6742–54.

Hershko, A., Ciechanover, A. & Varshavsky, A., 2000. Basic Medical Research Award. The ubiquitin system. Nature medicine, 6(10), pp.1073–81.

Hilt, W. & Wolf, D.H., 1996. Proteasomes: destruction as a programme. Trends in Biochemical Sciences, 21(3), pp.96–102.

Hochstrasser, M., 1995. Ubiquitin, proteasomes, and the regulation of intracellular protein degradation. Current opinion in cell biology, 7(2), pp.215–23.

Hosokawa, N. et al., 2009. Nutrient-dependent mTORC1 association with the ULK1-Atg13-FIP200 complex required for autophagy. Molecular biology of the cell, 20(7), pp.1981–91.

Huang, C.-C. et al., 2014. Metabolism and mis-metabolism of the neuropathological signature protein TDP-43. Journal of cell science, p.jcs.136150–.

128

Hubbard, V.M. et al., 2012. Selective autophagy in the maintenance of cellular homeostasis in aging organisms. Biogerontology, 13(1), pp.21–35.

Ichimura, Y. et al., 2008. Selective turnover of p62/A170/SQSTM1 by autophagy. Autophagy, 4(8), pp.1063–6.

Igaz, L.M. et al., 2009. Expression of TDP-43 C-terminal Fragments in Vitro Recapitulates Pathological Features of TDP-43 Proteinopathies. The Journal of biological chemistry, 284(13), pp.8516–24.

Iwata, A. et al., 2005. HDAC6 and Microtubules Are Required for Autophagic Degradation of Aggregated Huntingtin * . , 280(48), pp.40282–40292.

Jaeger, P.A. & Wyss-Coray, T., 2009. All-you-can-eat: autophagy in neurodegeneration and neuroprotection. Molecular neurodegeneration, 4(1), p.16.

Jing, Y. et al., 2011. Dextran sulfate inhibits staurosporine-induced apoptosis in Chinese hamster ovary (CHO) cells: Involvement of the mitochondrial pathway. Process Biochemistry, 46(1), pp.427–432.

Johansen, T. & Lamark, T., 2011. Selective autophagy mediated by autophagic adapter proteins. Autophagy, 7(3), pp.279–96.

Johnson, B.S. et al., 2008. A yeast TDP-43 proteinopathy model: Exploring the molecular determinants of TDP-43 aggregation and cellular toxicity. Proceedings of the National Academy of Sciences of the United States of America, 105(17), pp.6439–44.

Johnson, B.S. et al., 2009. TDP-43 is intrinsically aggregation-prone, and amyotrophic lateral sclerosis-linked mutations accelerate aggregation and increase toxicity. The Journal of biological chemistry, 284(30), pp.20329–39.

Johnston, J.A. et al., 2000. Formation of high molecular weight complexes of mutant Cu, Zn-superoxide dismutase in a mouse model for familial amyotrophic lateral sclerosis. Proceedings of the National Academy of Sciences of the United States of America, 97(23), pp.12571–6.

Johnston, J.A., Ward, C.L. & Kopito, R.R., 1998. Aggresomes: a cellular response to misfolded proteins. The Journal of cell biology, 143(7), pp.1883–98.

Kabashi, E. et al., 2010. Gain and loss of function of ALS-related mutations of TARDBP (TDP-43) cause motor deficits in vivo. Human molecular genetics, 19(4), pp.671–83.

Kabashi, E. et al., 2008. TARDBP mutations in individuals with sporadic and familial amyotrophic lateral sclerosis. Nature genetics, 40(5), pp.572–4.

129

Kabeya, Y. et al., 2000. LC3, a mammalian homologue of yeast Apg8p, is localized in autophagosome membranes after processing. The EMBO journal, 19(21), pp.5720–8.

Kabuta, T., Suzuki, Y. & Wada, K., 2006. Degradation of amyotrophic lateral sclerosis-linked mutant Cu,Zn-superoxide dismutase proteins by macroautophagy and the proteasome. The Journal of biological chemistry, 281(41), pp.30524–33.

Kalmar, B. et al., 2008. Late stage treatment with arimoclomol delays disease progression and prevents protein aggregation in the SOD1 mouse model of ALS. Journal of neurochemistry, 107(2), pp.339–50.

Katsumata, K. et al., 2010. Dynein- and activity-dependent retrograde transport of autophagosomes in neuronal axons. Autophagy, 6(3), pp.378–85.

Kawaguchi, Y. et al., 2003. The deacetylase HDAC6 regulates aggresome formation and cell viability in response to misfolded protein stress. Cell, 115(6), pp.727–38.

Kayed, R. et al., 2003. Common structure of soluble amyloid oligomers implies common mechanism of pathogenesis. Science (New York, N.Y.), 300(5618), pp.486–9.

Kim, D. et al., 2007. SIRT1 deacetylase protects against neurodegeneration in models for Alzheimer’s disease and amyotrophic lateral sclerosis. The EMBO journal, 26(13), pp.3169–79.

Kim, J. et al., 2013. Differential regulation of distinct Vps34 complexes by AMPK in nutrient stress and autophagy. Cell, 152(1-2), pp.290–303.

Kimura, S., Noda, T. & Yoshimori, T., 2007. Dissection of the autophagosome maturation process by a novel reporter protein, tandem fluorescent-tagged LC3. Autophagy, 3(5), pp.452–60.

King, A. et al., 2011. Ubiquitinated, p62 immunopositive cerebellar cortical neuronal inclusions are evident across the spectrum of TDP-43 proteinopathies but are only rarely additionally immunopositive for phosphorylation-dependent TDP-43. Neuropathology : official journal of the Japanese Society of Neuropathology, 31(3), pp.239–49.

Kirisako, T. et al., 1999. Formation process of autophagosome is traced with Apg8/Aut7p in yeast. The Journal of cell biology, 147(2), pp.435–46.

Kirkin, V. et al., 2009. A role for NBR1 in autophagosomal degradation of ubiquitinated substrates. Molecular cell, 33(4), pp.505–16.

Kitamura, A. et al., 2006. Cytosolic chaperonin prevents polyglutamine toxicity with altering the aggregation state. Nature cell biology, 8(10), pp.1163–70.

130

Klionsky, D.J. et al., 2007. How shall I eat thee? Autophagy, 3(5), pp.413–6.

Klionsky, D.J. & Emr, S.D., 2000. Autophagy as a Regulated Pathway of Cellular Degradation. Science, 290(5497), pp.1717–21.

Köchl, R. et al., 2006. Microtubules facilitate autophagosome formation and fusion of autophagosomes with endosomes. Traffic (Copenhagen, Denmark), 7(2), pp.129–45.

Komander, D., 2009. The emerging complexity of protein ubiquitination. Biochemical Society transactions, 37(Pt 5), pp.937–53.

Komatsu, M. et al., 2007. Homeostatic levels of p62 control cytoplasmic inclusion body formation in autophagy-deficient mice. Cell, 131(6), pp.1149–63.

Komatsu, M. et al., 2005. Impairment of starvation-induced and constitutive autophagy in Atg7-deficient mice. The Journal of cell biology, 169(3), pp.425–34.

Komatsu, M. et al., 2006. Loss of autophagy in the central nervous system causes neurodegeneration in mice. Nature, 441(7095), pp.880–4.

Kopito, R.R., 2000. Aggresomes, inclusion bodies and protein aggregation. Trends in cell biology, 10(12), pp.524–30.

Korolchuk, V.I., Mansilla, A., et al., 2009. Autophagy inhibition compromises degradation of ubiquitin-proteasome pathway substrates. Molecular cell, 33(4), pp.517–27.

Korolchuk, V.I., Menzies, F.M. & Rubinsztein, D.C., 2009. A novel link between autophagy and the ubiquitin-proteasome system. Autophagy, 5(6), pp.862–3.

Korolchuk, V.I., Menzies, F.M. & Rubinsztein, D.C., 2010. Mechanisms of cross-talk between the ubiquitin-proteasome and autophagy-lysosome systems. F EBS letters, 584(7), pp.1393–8.

Kraft, C., Peter, M. & Hofmann, K., 2010. Selective autophagy: ubiquitin-mediated recognition and beyond. Nature cell biology, 12(9), pp.836–41.

Kroemer, G., Mariño, G. & Levine, B., 2010. Autophagy and the integrated stress response. Molecular cell, 40(2), pp.280–93.

Kühnlein, P. et al., 2008. Two German kindreds with familial amyotrophic lateral sclerosis due to TARDBP mutations. Archives of neurology, 65(9), pp.1185–9.

Kuo, P.-H. et al., 2009. Structural insights into TDP-43 in nucleic-acid binding and domain interactions. Nucleic acids research, 37(6), pp.1799–808.

131

Kwiatkowski, T.J. et al., 2009. Mutations in the FUS/TLS gene on chromosome 16 cause familial amyotrophic lateral sclerosis. Science (New York, N.Y.), 323(5918), pp.1205–8.

Kwong, L.K. et al., 2007. TDP-43 proteinopathy: the neuropathology underlying major forms of sporadic and familial frontotemporal lobar degeneration and motor neuron disease. Acta neuropathologica, 114(1), pp.63–70.

Lagier-Tourenne, C., Polymenidou, M. & Cleveland, D.W., 2010. TDP-43 and FUS/TLS: emerging roles in RNA processing and neurodegeneration. Human molecular genetics, 19(R1), pp.R46–64.

Laird, A.S. et al., 2010. Progranulin is neurotrophic in vivo and protects against a mutant TDP-43 induced axonopathy. PloS one, 5(10), p.e13368.

Laird, F.M. et al., 2008. Motor neuron disease occurring in a mutant dynactin mouse model is characterized by defects in vesicular trafficking. The Journal of neuroscience : the official journal of the Society for Neuroscience, 28(9), pp.1997–2005.

Lamark, T. & Johansen, T., 2012. Aggrephagy: selective disposal of protein aggregates by macroautophagy. International journal of cell biology, 2012, p.736905.

Lashuel, H.A. & Lansbury, P.T., 2006. Are amyloid diseases caused by protein aggregates that mimic bacterial pore-forming toxins? Quarterly reviews of biophysics, 39(2), pp.167–201.

Lee, J.-Y. et al., 2010. HDAC6 controls autophagosome maturation essential for ubiquitin-selective quality-control autophagy. The EMBO journal, 29(5), pp.969–80.

Leigh, P.N., 2007. Chapter 13 Amyotrophic lateral sclerosis. Handbook of clinical neurology, 82, pp.249–78.

Lelouard, H. et al., 2002. Transient aggregation of ubiquitinated proteins during dendritic cell maturation. Nature, 417(6885), pp.177–82.

Leveugle, B. et al., 1998. Heparin oligosaccharides that pass the blood-brain barrier inhibit beta-amyloid precursor protein secretion and heparin binding to beta-amyloid peptide. Journal of neurochemistry, 70(2), pp.736–44.

Li, W., Li, J. & Bao, J., 2012. Microautophagy: lesser-known self-eating. Cellular and molecular life sciences : CMLS, 69(7), pp.1125–36.

Li, Y. et al., 2010. A Drosophila model for TDP-43 proteinopathy. Proceedings of the National Academy of Sciences of the United States of America, 107(7), pp.3169–74.

132

Li, Liang; Zhang, Xiaojie; Le, W., 2008. Altered macroautophagy in the spinal cord of SOD1 mutant mice. Autophagy, 4(3), pp.290–3.

Liang, C. et al., 2008. Beclin1-binding UVRAG targets the class C Vps complex to coordinate autophagosome maturation and endocytic trafficking. Nature cell biology, 10(7), pp.776–87.

Limpert, A.S., Mattmann, M.E. & Cosford, N.D.P., 2013. Recent progress in the discovery of small molecules for the treatment of amyotrophic lateral sclerosis (ALS). Beilstein journal of organic chemistry, 9, pp.717–32.

Logroscino, G. et al., 2010. Incidence of amyotrophic lateral sclerosis in Europe. Journal of neurology, neurosurgery, and psychiatry, 81(4), pp.385–90.

Long, J. et al., 2008. Ubiquitin recognition by the ubiquitin-associated domain of p62 involves a novel conformational switch. The Journal of biological chemistry, 283(9), pp.5427–40.

Lu, Y., Ferris, J. & Gao, F.-B., 2009. Frontotemporal dementia and amyotrophic lateral sclerosis-associated disease protein TDP-43 promotes dendritic branching. Molecular brain, 2, p.30.

Ma, Q. et al., 2007. Heparin oligosaccharides as potential therapeutic agents in senile dementia. Current pharmaceutical design, 13(15), pp.1607–16.

Ma, Q. et al., 2002. The blood-brain barrier accessibility of a heparin-derived oligosaccharides C3. Thrombosis research, 105(5), pp.447–53.

Mackenzie, I.R.A., 2007. The neuropathology of FTD associated With ALS. Alzheimer disease and associated disorders, 21(4), pp.S44–9.

Maday, S., Wallace, K.E. & Holzbaur, E.L.F., 2012. Autophagosomes initiate distally and mature during transport toward the cell soma in primary neurons. The Journal of cell biology, 196(4), pp.407–17.

Martinez-Vicente, M. et al., 2008. Dopamine-modified alpha-synuclein blocks chaperone-mediated autophagy. The Journal of clinical investigation, 118(2), pp.777–88.

McCord, J.M. & Fridovich, I., 1969. Superoxide dismutase. An enzymic function for erythrocuprein (hemocuprein). The Journal of biological chemistry, 244(22), pp.6049–55.

McDonnell, M.E. et al., 2012. Riluzole prodrugs for melanoma and ALS: design, synthesis, and in vitro metabolic profiling. Bioorganic & medicinal chemistry, 20(18), pp.5642–8.

McEwan, D.G. & Dikic, I., 2011. The Three Musketeers of Autophagy: phosphorylation, ubiquitylation and acetylation. Trends in cell biology, 21(4), pp.195–201.

133

Menvielle, J.P. et al., 2013. Dual role of dextran sulfate 5000 Da as anti-apoptotic and pro-autophagy agent. Molecular biotechnology, 54(2), pp.711–20.

Merlini, G. et al., 2001. Protein aggregation. Clinical chemistry and laboratory medicine : CCLM / FESCC, 39(11), pp.1065–75.

Mijaljica, D., Prescott, M. & Devenish, R.J., 2011. Microautophagy in mammalian cells: revisiting a 40-year-old conundrum. Autophagy, 7(7), pp.673–82.

Miller, R.G. et al., 2007. Riluzole for amyotrophic lateral sclerosis (ALS)/motor neuron disease (MND). Cochrane database of systematic reviews (Online), (1), p.CD001447.

Miquel, E. et al., 2012. Modulation of astrocytic mitochondrial function by dichloroacetate improves survival and motor performance in inherited amyotrophic lateral sclerosis. S. T. Ferreira, ed. PloS one, 7(4), p.e34776.

Mitsumoto, H., 1997. Can we treat amyotrophic lateral sclerosis? Rinshō shinkeigaku = Clinical neurology, 37(12), p.1150.

Mitsumoto, H., 1997. Diagnosis and progression of ALS. Neurology, 48(Issue 4, Supplement 4), p.2S–8S.

Mizushima, N. et al., 1998. A protein conjugation system essential for autophagy. Nature, 395(6700), pp.395–8.

Mizushima, N. et al., 2008. Autophagy fights disease through cellular self-digestion. Nature, 451(7182), pp.1069–75.

Mizushima, N., 2004. Methods for monitoring autophagy. The international journal of biochemistry & cell biology, 36(12), pp.2491–502.

Mizushima, N., 2005. The pleiotropic role of autophagy: from protein metabolism to bactericide. Cell death and differentiation, 12 Suppl 2, pp.1535–1541.

Montie, H.L. et al., 2009. Cytoplasmic retention of polyglutamine-expanded androgen receptor ameliorates disease via autophagy in a mouse model of spinal and bulbar muscular atrophy. Human molecular genetics, 18(11), pp.1937–50.

Myeku, N. & Figueiredo-Pereira, M.E., 2011. Dynamics of the degradation of ubiquitinated proteins by proteasomes and autophagy: association with sequestosome 1/p62. The Journal of biological chemistry, 286(25), pp.22426–40.

Nakai, A. et al., 2007. The role of autophagy in cardiomyocytes in the basal state and in response to hemodynamic stress. Nature medicine, 13(5), pp.619–24.

Nandi, D. et al., 2006. The ubiquitin-proteasome system. Journal of biosciences, 31(1), pp.137–55.

134

Nedelsky, N.B., Todd, P.K. & Taylor, J.P., 2008. Autophagy and the ubiquitin-proteasome system: collaborators in neuroprotection. Biochimica et biophysica acta, 1782(12), pp.691–9.

Neumann, M. et al., 2009. Phosphorylation of S409/410 of TDP-43 is a consistent feature in all sporadic and familial forms of TDP-43 proteinopathies. Acta neuropathologica, 117(2), pp.137–49.

Neumann, M. et al., 2006. Ubiquitinated TDP-43 in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Science (New York, N.Y.), 314(5796), pp.130–3.

Nezis, I.P. et al., 2008. Ref(2)P, the Drosophila melanogaster homologue of mammalian p62, is required for the formation of protein aggregates in adult brain. The Journal of cell biology, 180(6), pp.1065–71.

Nixon, R.A., 2006. Autophagy in neurodegenerative disease: friend, foe or turncoat? Trends in neurosciences, 29(9), pp.528–35.

Nixon, R.A. et al., 2005. Extensive involvement of autophagy in Alzheimer disease: an immuno-electron microscopy study. Journal of neuropathology and experimental neurology, 64(2), pp.113–22.

Nonaka, T. et al., 2009. Truncation and pathogenic mutations facilitate the formation of intracellular aggregates of TDP-43. Human molecular genetics, 18(18), pp.3353–64.

Oeda, T. et al., 2001. Oxidative stress causes abnormal accumulation of familial amyotrophic lateral sclerosis-related mutant SOD1 in transgenic Caenorhabditis elegans. Human molecular genetics, 10(19), pp.2013–23.

Ogata, M. et al., 2006. Autophagy is activated for cell survival after endoplasmic reticulum stress. Molecular and cellular biology, 26(24), pp.9220–31.

Okamoto, K. et al., 1991. Reexamination of granulovacuolar degeneration. Acta Neuropathologica, 82(5), pp.340–345.

Olzmann, J.A. et al., 2007. Parkin-mediated K63-linked polyubiquitination targets misfolded DJ-1 to aggresomes via binding to HDAC6. The Journal of cell biology, 178(6), pp.1025–38.

Otomo, A., Pan, L. & Hadano, S., 2012. Dysregulation of the autophagy-endolysosomal system in amyotrophic lateral sclerosis and related motor neuron diseases. Neurology research international, 2012, p.498428.

Ou, S.H. et al., 1995. Cloning and characterization of a novel cellular protein, TDP-43, that binds to human immunodeficiency virus type 1 TAR DNA sequence motifs. Journal of virology, 69(6), pp.3584–96.

135

Øverbye, A., Brinchmann, M.F. & Seglen, P.O., 2007. Proteomic Analysis of Membrane-Associated Proteins from Rat Liver Autophagosomes. Autophagy, 3(4), pp.300–322.

Pan, T. et al., 2008. Neuroprotection of rapamycin in lactacystin-induced neurodegeneration via autophagy enhancement. Neurobiology of disease, 32(1), pp.16–25.

Pandey, U.B. et al., 2007. HDAC6 rescues neurodegeneration and provides an essential link between autophagy and the UPS. Nature, 447(7146), pp.859–63.

Pankiv, S. et al., 2007. p62/SQSTM1 binds directly to Atg8/LC3 to facilitate degradation of ubiquitinated protein aggregates by autophagy. The Journal of biological chemistry, 282(33), pp.24131–45.

Pasquali, L. et al., 2009. Autophagy, lithium, and amyotrophic lateral sclerosis. Muscle & nerve, 40(2), pp.173–94.

Pérez, M. et al., 1998. Sulphated glycosaminoglycans prevent the neurotoxicity of a human prion protein fragment. The Biochemical journal, 335 ( Pt 2, pp.369–74.

Pesiridis, G.S., Lee, V.M.-Y. & Trojanowski, J.Q., 2009. Mutations in TDP-43 link glycine-rich domain functions to amyotrophic lateral sclerosis. Human molecular genetics, 18(R2), pp.R156–62.

Peters, J.-M., 1994. Proteasomes: protein degradation machines of the cell. Trends in Biochemical Sciences, 19(9), pp.377–382.

Phillips, J.P., 1989. Null Mutation of Copper/Zinc Superoxide Dismutase in Drosophila Confers Hypersensitivity to Paraquat and Reduced Longevity. Proceedings of the National Academy of Sciences, 86(8), pp.2761–2765.

Pickford, F. et al., 2008. The autophagy-related protein beclin 1 shows reduced expression in early Alzheimer disease and regulates amyloid beta accumulation in mice. The Journal of clinical investigation, 118(6), pp.2190–9.

Pietrocola, F. et al., 2012. Pro-autophagic polyphenols reduce the acetylation of cytoplasmic proteins. Cell cycle (Georgetown, Tex.), 11(20), pp.3851–60.

Polymenidou, M. et al., 2011. Long pre-mRNA depletion and RNA missplicing contribute to neuronal vulnerability from loss of TDP-43. Nature neuroscience, 14(4), pp.459–68.

Punnonen, E.L. & Reunanen, H., 1990. Effects of vinblastine, leucine, and histidine, and 3-methyladenine on autophagy in Ehrlich ascites cells. Experimental and molecular pathology, 52(1), pp.87–97.

Qin, Z.-H. et al., 2003. Autophagy regulates the processing of amino terminal huntingtin fragments. Human molecular genetics, 12(24), pp.3231–44.

136

Ravikumar, B., 2002. Aggregate-prone proteins with polyglutamine and polyalanine expansions are degraded by autophagy. Human Molecular Genetics, 11(9), pp.1107–1117.

Ravikumar, B. et al., 2005. Dynein mutations impair autophagic clearance of aggregate-prone proteins. Nature genetics, 37(7), pp.771–6.

Ravikumar, B. et al., 2004. Inhibition of mTOR induces autophagy and reduces toxicity of polyglutamine expansions in fly and mouse models of Huntington disease. Nature genetics, 36(6), pp.585–95.

Ritson, G.P. et al., 2010. TDP-43 mediates degeneration in a novel Drosophila model of disease caused by mutations in VCP/p97. The Journal of neuroscience : the official journal of the Society for Neuroscience, 30(22), pp.7729–39.

Rodriguez-Gonzalez, A. et al., 2008. Role of the aggresome pathway in cancer: targeting histone deacetylase 6-dependent protein degradation. Cancer research, 68(8), pp.2557–60.

Rohn, T.T., 2008. Caspase-cleaved TAR DNA-binding protein-43 is a major pathological finding in Alzheimer’s disease. Brain research, 1228, pp.189–98.

Rohn, T.T. & Kokoulina, P., 2009. Caspase-cleaved TAR DNA-binding protein-43 in Pick’s disease. International journal of physiology, pathophysiology and pharmacology, 1(1), pp.25–32.

Rosen, D.R. et al., 1993. Mutations in Cu/Zn superoxide dismutase gene are associated with familial amyotrophic lateral sclerosis. Nature, 362(6415), pp.59–62.

Rothstein, J.D., Martin, L.J. & Kuncl, R.W., 1992. Decreased glutamate transport by the brain and spinal cord in amyotrophic lateral sclerosis. The New England journal of medicine, 326(22), pp.1464–8.

Rowland, L.P., 1998. Diagnosis of amyotrophic lateral sclerosis. Journal of the neurological sciences, 160 Suppl , pp.S6–24.

Rowland, L.P., 2001. How amyotrophic lateral sclerosis got its name: the clinical-pathologic genius of Jean-Martin Charcot. Archives of neurology, 58(3), pp.512–5.

Rubin, D.M. & Finley, D., 1995. Proteolysis: The proteasome: a protein-degrading organelle? Current Biology, 5(8), pp.854–858.

Rubinsztein, D.C., 2006. The roles of intracellular protein-degradation pathways in neurodegeneration. Nature, 443(7113), pp.780–6.

Rutherford, N.J. et al., 2008. Novel mutations in TARDBP (TDP-43) in patients with familial amyotrophic lateral sclerosis. G. A. Cox, ed. PLoS genetics, 4(9), p.e1000193.

137

Ryu, H. et al., 2005. Sodium phenylbutyrate prolongs survival and regulates expression of anti-apoptotic genes in transgenic amyotrophic lateral sclerosis mice. Journal of neurochemistry, 93(5), pp.1087–98.

Sakai, Y. et al., 1998. Peroxisome degradation by microautophagy in Pichia pastoris: identification of specific steps and morphological intermediates. The Journal of cell biology, 141(3), pp.625–36.

Sarkar, S. et al., 2005. Lithium induces autophagy by inhibiting inositol monophosphatase. The Journal of cell biology, 170(7), pp.1101–11.

Sarkar, S. et al., 2007. Small molecules enhance autophagy and reduce toxicity in Huntington’s disease models. Nature chemical biology, 3(6), pp.331–8.

Sasaki, S., 2011. Autophagy in spinal cord motor neurons in sporadic amyotrophic lateral sclerosis. Journal of neuropathology and experimental neurology, 70(5), pp.349–59.

Saudou, F. et al., 1998. Huntingtin Acts in the Nucleus to Induce Apoptosis but Death Does Not Correlate with the Formation of Intranuclear Inclusions. Cell, 95(1), pp.55–66.

Schmid, B. et al., 2013. Loss of ALS-associated TDP-43 in zebrafish causes muscle degeneration, vascular dysfunction, and reduced motor neuron axon outgrowth. Proceedings of the National Academy of Sciences of the United States of America, 110(13), pp.4986–91.

Scotter, E.L. et al., 2014. Differential roles of the ubiquitin proteasome system (UPS) and autophagy in the clearance of soluble and aggregated TDP-43 species. Journal of cell science, p.jcs.140087–.

Seibenhener, M.L. et al., 2004. Sequestosome 1/p62 is a polyubiquitin chain binding protein involved in ubiquitin proteasome degradation. Molecular and cellular biology, 24(18), pp.8055–68.

Sephton, C.F. et al., 2010. TDP-43 is a developmentally regulated protein essential for early embryonic development. The Journal of biological chemistry, 285(9), pp.6826–34.

Shaw, C.E., Al-Chalabi, A. & Leigh, N., 2001. Progress in the pathogenesis of amyotrophic lateral sclerosis. Current Neurology and Neuroscience Reports, 1(1), pp.69–76.

Shaw, P.J., 2005. Molecular and cellular pathways of neurodegeneration in motor neurone disease. Journal of neurology, neurosurgery, and psychiatry, 76(8), pp.1046–57.

Shiina, Y. et al., 2010. TDP-43 dimerizes in human cells in culture. Cellular and molecular neurobiology, 30(4), pp.641–52.

138

Simonsen, A. et al., 2008. Promoting basal levels of autophagy in the nervous system enhances longevity and oxidant resistance in adult Drosophila. Autophagy, 4(2), pp.176–84.

Simonsen, A. & Tooze, S.A., 2009. Coordination of membrane events during autophagy by multiple class III PI3-kinase complexes. The Journal of cell biology, 186(6), pp.773–82.

Smith, R.E. & Farquhar, M.G., 1966. Lysosome function in the regulation of the secretory process in cells of the anterior pituitary gland. The Journal of cel l biology, 31(2), pp.319–47.

La Spada, A.R. et al., 1991. Androgen receptor gene mutations in X-linked spinal and bulbar muscular atrophy. Nature, 352(6330), pp.77–9.

Spilman, P. et al., 2010. Inhibition of mTOR by rapamycin abolishes cognitive deficits and reduces amyloid-beta levels in a mouse model of Alzheimer’s disease. P. F. Ferrari, ed. PloS one, 5(4), p.e9979.

Staats, K.A. et al., 2013. Rapamycin increases survival in ALS mice lacking mature lymphocytes. Molecular neurodegeneration, 8(1), p.31.

Stefanis, L. et al., 2001. Expression of A53T mutant but not wild-type alpha-synuclein in PC12 cells induces alterations of the ubiquitin-dependent degradation system, loss of dopamine release, and autophagic cell death. The Journal of neuroscience : the official journal of the Society for Neuroscience, 21(24), pp.9549–60.

Stein, M.-P. et al., 2005. Interaction and functional analyses of human VPS34/p150 phosphatidylinositol 3-kinase complex with Rab7. Methods in enzymology, 403, pp.628–49.

Strong, M.J. et al., 2007. TDP43 is a human low molecular weight neurofilament (hNFL) mRNA-binding protein. Molecular and cellular neurosciences, 35(2), pp.320–7.

Stutzmann, J.-M. et al., 2002. Neuroprotective profile of enoxaparin, a low molecular weight heparin, in in vivo models of cerebral ischemia or traumatic brain injury in rats: a review. CNS drug reviews, 8(1), pp.1–30.

Szeto, J. et al., 2006. ALIS are stress-induced protein storage compartments for substrates of the proteasome and autophagy. Autophagy, 2(3), pp.189–99.

Takagi, T. et al., 2005. Dextran sulfate suppresses cell adhesion and cell cycle progression of melanoma cells. Anticancer research, 25(2A), pp.895–902.

Tan, J.M.M. et al., 2008. Lysine 63-linked ubiquitination promotes the formation and autophagic clearance of protein inclusions associated with neurodegenerative diseases. Human molecular genetics, 17(3), pp.431–9.

139

Tanaka, M. et al., 2004. Aggresomes formed by alpha-synuclein and synphilin-1 are cytoprotective. The Journal of biological chemistry, 279(6), pp.4625–31.

Tanji, K. et al., 2012. p62/sequestosome 1 binds to TDP-43 in brains with frontotemporal lobar degeneration with TDP-43 inclusions. Journal of neuroscience research, 90(10), pp.2034–42.

Tdp-, P. et al., 2011. Regulation of autophagy by neuropathological protein TDP-43. The Journal of biological chemistry, 286(52), pp.44441–8.

Thumm, M. et al., 1994. Isolation of autophagocytosis mutants of Saccharomyces cerevisiae. F EBS letters, 349(2), pp.275–80.

Tooze, J. et al., 1990. In exocrine pancreas, the basolateral endocytic pathway converges with the autophagic pathway immediately after the early endosome. The Journal of cell biology, 111(2), pp.329–45.

Tsubamoto, Y. et al., 1994. Dextran sulfate, a competitive inhibitor for scavenger receptor, prevents the progression of atherosclerosis in Watanabe heritable hyperlipidemic rabbits. Atherosclerosis, 106(1), pp.43–50.

Tsukada, M. & Ohsumi, Y., 1993. Isolation and characterization of autophagy-defective mutants of Saccharomyces cerevisiae. F EBS Letters, 333(1-2), pp.169–174.

Udan, M. & Baloh, R.H., 2011. Implications of the prion-related Q/N domains in TDP-43 and FUS. Prion, 5(1), pp.1–5.

Urushitani, M. et al., 2010. Synergistic effect between proteasome and autophagosome in the clearance of polyubiquitinated TDP-43. Journal of neuroscience research, 88(4), pp.784–97.

Vabulas, R.M. & Hartl, F.U., 2005. Protein synthesis upon acute nutrient restriction relies on proteasome function. Science (New York, N.Y.), 310(5756), pp.1960–3.

Valdmanis, P.N. & Rouleau, G.A., 2008. Genetics of familial amyotrophic lateral sclerosis. Neurology, 70(2), pp.144–52.

Vance, C. et al., 2009. Mutations in FUS, an RNA processing protein, cause familial amyotrophic lateral sclerosis type 6. Science (New York, N.Y.), 323(5918), pp.1208–11.

Verdel, A. & Khochbin, S., 1999. Identification of a new family of higher eukaryotic histone deacetylases. Coordinate expression of differentiation-dependent chromatin modifiers. The Journal of biological chemistry, 274(4), pp.2440–5.

Verhoef, L.G.G.C. et al., 2002. Aggregate formation inhibits proteasomal degradation of polyglutamine proteins. Human molecular genetics, 11(22), pp.2689–700.

140

Voigt, A. et al., 2010. TDP-43-mediated neuron loss in vivo requires RNA-binding activity. M. B. Feany, ed. PloS one, 5(8), p.e12247.

Wang, H.-Y. et al., 2004. Structural diversity and functional implications of the eukaryotic TDP gene family. Genomics, 83(1), pp.130–9.

Wang, I.-F. et al., 2012. Autophagy activators rescue and alleviate pathogenesis of a mouse model with proteinopathies of the TAR DNA-binding protein 43. Proceedings of the National Academy of Sciences of the United States of America, 109(37), pp.15024–9.

Wang, I.-F., Wu, L.-S. & Shen, C.-K.J., 2008. TDP-43: an emerging new player in neurodegenerative diseases. Trends in molecular medicine, 14(11), pp.479–85.

Wang, X. et al., 2010. Degradation of TDP-43 and its pathogenic form by autophagy and the ubiquitin-proteasome system. Neuroscience letters, 469(1), pp.112–6.

Ward, W.F., 2002. Protein degradation in the aging organism. Progress in molecular and subcellular biology, 29, pp.35–42.

Warraich, S.T. et al., 2010. TDP-43: a DNA and RNA binding protein with roles in neurodegenerative diseases. The international journal of biochemistry & cel l biology, 42(10), pp.1606–9.

Watanabe, M. et al., 2001. Histological evidence of protein aggregation in mutant SOD1 transgenic mice and in amyotrophic lateral sclerosis neural tissues. Neurobiology of disease, 8(6), pp.933–41.

Watson, M.R. et al., 2008. A Drosophila Model for Amyotrophic Lateral Sclerosis Reveals Motor Neuron Damage by Human SOD1. Journal of Biological Chemistry, 283(36), pp.24972–24981.

Webb, J.L. et al., 2003. Alpha-Synuclein is degraded by both autophagy and the proteasome. The Journal of biological chemistry, 278(27), pp.25009–13.

Webb, J.L., Ravikumar, B. & Rubinsztein, D.C., 2004. Microtubule disruption inhibits autophagosome-lysosome fusion: implications for studying the roles of aggresomes in polyglutamine diseases. The international journal of biochemistry & cell biology, 36(12), pp.2541–50.

Welchman, R.L., Gordon, C. & Mayer, R.J., 2005. Ubiquitin and ubiquitin-like proteins as multifunctional signals. Nature reviews. Molecular cell biology, 6(8), pp.599–609.

Winton, M.J. et al., 2008. Disturbance of nuclear and cytoplasmic TAR DNA-binding protein (TDP-43) induces disease-like redistribution, sequestration, and aggregate formation. The Journal of biological chemistry, 283(19), pp.13302–9.

Wong, E. & Cuervo, A.M., 2010. Autophagy gone awry in neurodegenerative diseases. Nature neuroscience, 13(7), pp.805–11.

141

Wooten, M.W. et al., 2008. Essential Role of Sequestosome 1/p62 in Regulating Accumulation of Lys63-ubiquitinated Proteins. Journal of Biological Chemistry, 283(11), pp.6783–6789.

Worms, P.M., 2001. The epidemiology of motor neuron diseases: a review of recent studies. Journal of the neurological sciences, 191(1-2), pp.3–9.

Xilouri, M. & Stefanis, L., 2011. Autophagic pathways in Parkinson disease and related disorders. Expert reviews in molecular medicine, 13, p.e8.

Yim, M.B. et al., 1996. A gain-of-function of an amyotrophic lateral sclerosis-associated Cu,Zn-superoxide dismutase mutant: An enhancement of free radical formation due to a decrease in Km for hydrogen peroxide. Proceedings of the National Academy of Sciences of the United States of America, 93(12), pp.5709–14.

Yoo, Y.-E. & Ko, C.-P., 2012. Dihydrotestosterone ameliorates degeneration in muscle, axons and motoneurons and improves motor function in amyotrophic lateral sclerosis model mice. L. Mei, ed. PloS one, 7(5), p.e37258.

Yoo, Y.-E. & Ko, C.-P., 2011. Treatment with trichostatin A initiated after disease onset delays disease progression and increases survival in a mouse model of amyotrophic lateral sclerosis. Experimental neurology, 231(1), pp.147–59.

Yorimitsu, T. & Klionsky, D.J., 2005. Autophagy: molecular machinery for self-eating. Cell death and differentiation, 12 Suppl 2(S2), pp.1542–52.

Yu, W.H. et al., 2005. Macroautophagy--a novel Beta-amyloid peptide-generating pathway activated in Alzheimer’s disease. The Journal of cell biology, 171(1), pp.87–98.

Zhang, T. et al., 2012. Caenorhabditis elegans RNA-processing protein TDP-1 regulates protein homeostasis and life span. The Journal of biological chemistry, 287(11), pp.8371–82.

Zhang, X. et al., 2011. Rapamycin treatment augments motor neuron degeneration in SOD1 G93A mouse model of amyotrophic lateral sclerosis. Autophagy, 7(4), pp.412–425.

Zhang, Y. et al., 2003. HDAC-6 interacts with and deacetylates tubulin and microtubules in vivo. The EMBO journal, 22(5), pp.1168–79.

Zhang, Y. et al., 2007. Progranulin mediates caspase-dependent cleavage of TAR DNA binding protein-43. The Journal of neuroscience : the official journal of the Society for Neuroscience, 27(39), pp.10530–4.

Zhang, Y.-J. et al., 2009. Aberrant cleavage of TDP-43 enhances aggregation and cellular toxicity. Proceedings of the National Academy of Sciences of the United States of America, 106(18), pp.7607–12.

142

Zhang, Y.-J. et al., 2010. Phosphorylation regulates proteasomal-mediated degradation and solubility of TAR DNA binding protein-43 C-terminal fragments. Molecular neurodegeneration, 5, p.33.


Recommended