studies on islet amyloid polypeptide aggregation: from model

110
Linköping University Medical Dissertations No. 1254 Studies on Islet Amyloid Polypeptide Aggregation: From Model Organism to Molecular Mechanisms Sebastian W Schultz Department of Clinical and Experimental Medicine Linköping University, Sweden Linköping 2011

Upload: phamtuyen

Post on 31-Dec-2016

220 views

Category:

Documents


0 download

TRANSCRIPT

Linköping  University  Medical  Dissertations  No.  1254                  Studies  on  Islet  Amyloid  Polypeptide  Aggregation:  

From  Model  Organism  to  Molecular  Mechanisms  

 

 

Sebastian  W  Schultz  

 

 

 

 

 

 

Department  of  Clinical  and  Experimental  Medicine  

Linköping  University,  Sweden  

Linköping  2011  

 

 

                                                         ©  Sebastian  W  Schultz      Cover:  Drosophila  brain;  green:  cell  nuclei  of  ventral  lateral  neurons,  red:  neuropil        During   the   course   of   the   research   underlying   this   thesis,   Sebastian  W   Schultz  was  enrolled   in  Forum  Scientium,  a  multidisciplinary  doctoral  programme  at  Linköping  University,  Sweden.      Printed  by  LiU-­‐Tryck,  Linköping,  Sweden,  2011      ISBN  978-­‐91-­‐7393-­‐099-­‐4  ISSN  0345-­‐0082  

                                       

Der  Weg  ist  das  Ziel                                                    

Supervisor    Gunilla  T  Westermark,  Professor  Department  of  Medical  Cell  Biology  Uppsala  University,  Sweden            Opponent    Anne  Simonsen,  Associate  Professor  Department  of  Biochemistry  University  of  Oslo,  Norway                                                          

Preface  

This  thesis  is  based  on  the  following  papers,  which  are  referred  to  in  the  text  by  their  roman  numerals:        

I. Paulsson  JF,  Schultz  SW,  Kohler  M,  Leibiger  I,  Berggren  PO,  Westermark  GT.  Real-­‐time  monitoring   of   apoptosis   by   caspase-­‐3-­‐like   protease  induced   FRET   reduction   triggered   by   amyloid   aggregation.   2008,  Exp  Diabetes  Res  2008:  865850.  

      A  free,  coloured  version  of  this  paper  can  be  downloaded  from:       www.hindawi.com/journals/edr/2008/865850/        

II. Schultz   SW,   Nilsson   KP,   Westermark   GT.   Drosophila   melanogaster    as   a   model   system   for   studies   of   islet   amyloid   polypeptide    aggregation.  2011,  PLoS  One  6:e20221.  

    A  free,  coloured  version  of  this  paper  can  be  downloaded  from:     www.plosone.org/article/info%3Adoi%2F10.1371%2Fjournal.pone.0020221        

III. Schultz   SW,   Gu   X,   Rusten   TE,   Alenius   M,  Westermark   GT.  HIAPP   and  hproIAPP   trigger   selective   autophagy   and   inhibit   the   neuro-­‐protective  effect  of  autophagy.  Manuscript.  

                             

Abstract  

The  proper  folding  of  a  protein  into  its  defined  three-­‐dimensional  structure  is  one  of  the  many   fundamental  challenges  a  cell  encounters.  A  number  of   tightly  controlled  pathways  have  evolved  to  assist  in  the  proper  folding  of  a  protein,  but  also  to  aid  in  the   removal   of   misfolded   proteins.   Despite   the   presence   of   these   pathways  accumulation   of   misfolded   proteins   can   still   occur.   Amyloid   deposits   consist   of  misfolded  proteins  with   a   characteristic   highly   ordered   fibrillar   structure   that  will  exert   affinity   for   the   amyloid   dye   Congo   red   and   has   a   unique   X-­‐ray   diffraction  pattern.   Currently   27   different   proteins   have   been   identified   as   amyloid   forming  proteins   in   human,   however   the   exact   role   of   amyloid   in   the   pathogenesis   of   the  connected  disease  is  most  often  unclear.  Islet  amyloid  is  made  up  of  the  beta  cell  derived  hormone  islet  amyloid  polypeptide  (IAPP)   and   is   associated  with   the   development   of   type   2   diabetes.   Propagation   of  IAPP-­‐fibrils   is  believed   to  be  one   important  cause  of   the  pancreatic  beta  cell  death  detected  in  patients  with  type  2  diabetes.  IAPP  is  a  naturally  occurring  polypeptide  hormone   stored   and   secreted   together   with   insulin.   IAPP   and   insulin   arise   from  posttranslational   processing   of   their   biological   inactive   precursors   proIAPP   and  proinsulin.  In  addition  to  human,  cat  and  monkey  IAPP  will  form  amyloid  deposits  in  conditions  resembling  human  type  2  diabetes.  However,  IAPP  from  mouse  and  rat  do  not  form  amyloid  as  a  result  of  the  differences  in  amino  acid  sequence.  My  main  research  goal  was  to  establish  a  unique  model  system  suitable  to  study  the  effects  of  proIAPP  and  IAPP  aggregation.   I  selected  Drosophila  melanogaster  due   to  its   many   suitable   characteristics   as   a   model   organism   and   its   superior   genetic  toolbox.   I   have   demonstrated   that   over-­‐expression   of   hproIAPP   and   hIAPP   in   the  central   nervous   system   (CNS)   results   in   aggregate   formation   in   the   brain   and  neighbouring  fat  body.  Consistent  with  previous  studies,  expression  of  mIAPP  does  not   result   in   the   formation  of   aggregates.  To   investigate   the   intracellular   effects  of  hproIAPP   and  hIAPP   aggregation   on   a   specific   population   of   neurons,  we   targeted  the   expression   of   these   peptides   specifically   to   16   neurons   in   the   brain,   the   pdf-­‐neurons.   These   pdf-­‐neurons   are   divided   into   2   clusters   of   8   cells   per   brain  hemisphere.   First   I   showed   that   expression   of   aggregation   prone   hIAPP   and  hproIAPP  resulted   in  significant  death  of   the  8  cells,  whereas  expression  of  mIAPP  had  no  such  effect.   In  efforts   to  pinpoint   the  mechanisms  behind   the  observed  cell  death  I  demonstrated  that  hproIAPP  and  hIAPP  both  pass  the  ERs  quality  control  for  protein   folding   and   that   the   initiated   cell   death   does   not   occur   through   classical  apoptosis.   Instead,   selective   autophagy   is   activated   by   hIAPP   and   hproIAPP.   This  activation   counteracts   the   usually   neuro-­‐protective   effects   of   autophagy   and  contributes   to   cell   death.   Strikingly,   I   also   showed   that   Aβ,   the   amyloid   protein  implicated   in   Alzheimer’s   disease,   does   not   exhibit   any   intracellular   toxicity  when  expressed   in   pdf-­‐cells.   This   supports   the   existence   of   separate   toxic   pathways   for  different  amyloid  proteins.    

Popular  scientific  summary  

Proteins  are  one  of   the  building  blocks  of   life.  They  are   important   for  almost  every  process  in  the  cell,  e.g.  forming  a  framework  involved  in  cellular  structure,  activation  of   chemical   reactions   and   mediating   cell   signals   and   cell   interactions.   However,  proteins  have  to  adopt  a  pre-­‐defined  three-­‐dimensional  fold,  referred  to  as  its  native  confirmation,   in   order   to   function.   Because   proteins   are   so   important,   cells   have  developed  highly  sophisticated  and  tightly  controlled  pathways  used  to  assist   their  proper   folding   and   to   remove   misfolded   proteins.   Despite   quality   control,  accumulation   of   misfolded   proteins   can   occur.   Amyloidosis   is   a   group   of   protein  misfolding  diseases.  Hitherto,  27  different  proteins  have  been  identified  as  amyloid  forming   in  man.  Each  amyloid  protein   is  associated  with  a  specific  disease,  but   the  exact   role   for   amyloid   in   the   pathogenesis   of   the   illness   is   unclear.   All   amyloid  deposits  share  certain  characteristics,  they  have  all  affinity  for  amyloid  specific  dyes  and  methods  providing  high-­‐resolution  information  reveal  a  highly  ordered  fibrillar  structure.  The  protein  I  have  been  working  on  is  the  hormone  islet  amyloid  polypeptide  (IAPP)  that   together   with   insulin   and   glucagon   participates   in   the   regulation   of   blood  glucose.   IAPP   can   form   amyloid   in   pancreas   and   this   is   associated   with   type   2  diabetes.   After   food   intake   the   blood   glucose   concentration   raises,   which   leads   to  release  of  insulin  from  beta  cells  in  the  pancreas.  Insulin  facilitates  cellular  uptake  of  sugar  and  thereby  lowers  the  blood  glucose  concentration.  Patients  that  suffer  from  type   2   diabetes   cannot   produce   sufficient   amounts   of   insulin   and   they   develop  chronic  elevated  blood  sugar  level.  One  reason  for  the  decreased  insulin  secretion  is  the  replacement  of  beta  cells  by  IAPP-­‐amyloid,  and  it  is  believed  that  islet  amyloid  is  responsible  for  this  cell  reduction  and  contributes  to  insulin  deficiency.    One  question  that  still  remains  to  be  answered  is  -­‐  how  does  IAPP-­‐amyloid  mediate  cell   death?   Since   IAPP   and   insulin   are   produced   by   the   same   cells,   death   can   be  initiated  from  the  inside  or  from  the  outside  of  the  cell.  For  my  work  I  have  set  up  a  new   Drosophila   melanogaster   (fruit   fly)   model   to   study   effects   of   aggregation   of  human  IAPP  and  its  precursor  proIAPP.  I  have  produced  transgenic  flies  that  secrete  human  IAPP  or  proIAPP  and  shown  that  expression  of  these  proteins  in  the  fly  head  results  in  aggregation  (paper  II).  In  paper  III,  I  limited  IAPP  and  proIAPP  expression  to  a  subset  of  16  neurons,  and  showed   that   this  caused  cell  death.  The  mechanism  behind  intracellular  cell  death  was  studied  in  detail  and  I  was  able  to  show  that  the  autophagy   (self-­‐eating)   pathway   was   selectively   triggered   by   human   IAPP   and  human   proIAPP.   Gained   evidence   indicates   that   activation   of   this   self-­‐eating  (autophagy)   pathway  decreases   the   normal   protective  mechanism  of   this   pathway  and  thereby  contributes  to  cell  death.  I  have  included  studies  on  Aβ,  the  protein  that  forms  amyloid  in  patients  with  Alzheimer’s  disease.  Aβ  expression  in  the  16  cells  did  not   result   in   cell   death.   Instead,   comparison   of   Aβ   and   IAPP/proIAPP   expression  revealed  that  amyloid  proteins  use  different  pathways  to  exhibit  their  toxicity.    

                           

   

TABLE  OF  CONTENTS  ABBREVIATIONS  .................................................................................................................  1  

INTRODUCTION  ...................................................................................................................  3  

PROTEIN  FOLDING  AND  MISFOLDING  ................................................................................................  4  

AMYLOID  AND  AMYLOIDOSIS  ..............................................................................................................  5  

History  and  definitions  ..............................................................................................................  5  

Amyloid  and  diseases  .................................................................................................................  6  

Structure  of  amyloid  ...................................................................................................................  8  

Non-­‐fibrillar  components  in  amyloid  deposits  .................................................................  9  

Amyloid  formation  ....................................................................................................................  10  

Toxic  effects  .................................................................................................................................  11  

Functional  amyloid  ...................................................................................................................  12  

ISLET  AMYLOID  POLYPEPTIDE  (IAPP)  ...........................................................................................  13  

General  introduction  ................................................................................................................  13  

Prohormone  processing  ..........................................................................................................  15  

IAPP  and  type  2  diabetes  ........................................................................................................  17  

IAPP  fibril  formation  ................................................................................................................  18  

Transgenic  animal  models  with  hIAPP  .............................................................................  21  

Aβ  .........................................................................................................................................................  22  

Alzheimer’s  disease  ...................................................................................................................  22  

Aβ  and  IAPP  .................................................................................................................................  23  

DROSOPHILA  MELANOGASTER  AS  MODEL  SYSTEM  ........................................................................  25  

History  of  Drosophila  as  model  system  .............................................................................  25  

Huge  genetic  toolbox:  Gal4/UAS  system  ...........................................................................  26  

Drosophila  models  for  protein  aggregation  ....................................................................  28  

MOLECULAR  PATHWAYS  CONNECTED  TO  PROTEIN  MISFOLDING  ...............................................  31  

ER-­‐stress  and  Unfolded  protein  response  (UPR)  ...........................................................  31  

Apoptosis  .......................................................................................................................................  37  

Autophagy  ....................................................................................................................................  41  

AIMS  OF  THE  THESIS  .......................................................................................................  51  

   

 

MATERIAL  AND  METHODS  ............................................................................................  53  

WORKING  WITH  DROSOPHILA  .........................................................................................................  54  

P-­‐element  insertion  ...................................................................................................................  54  

Survival  assay  ..............................................................................................................................  54  

DETECTION  METHODS  .......................................................................................................................  55  

Immunofluorescence  –  tissue  preparation  ......................................................................  55  

Congo  Red  or  pFTAA  .................................................................................................................  55  

Image  processing  .......................................................................................................................  56  

RESULTS  AND  DISCUSSION  ............................................................................................  57  

EXTRACELLULAR  AMYLOID  FORMATION  INDUCES  APOPTOSIS  (PAPER  I)  .................................  58  

CHARACTERISATION  OF  A  NEW  DROSOPHILA  MODEL  FOR  STUDIES  OF  IAPP  AGGREGATION  

(PAPER  II)  ............................................................................................................................................  60  

HPROIAPP  AND  HIAPP  TRIGGER  SELECTIVE  AUTOPHAGY  (PAPER  III)  ...................................  64  

GENERAL  DISCUSSION  AND  FUTURE  PERSPECTIVES  ...........................................  69  

ACKNOWLEDGEMENTS  ...................................................................................................  73  

REFERENCES  .......................................................................................................................  77  

 

 

     

   

  1  

Abbreviations  Aβ     amyloid-­‐β  peptide  AD     Alzheimer’s  disease  AGE     advanced  glycation  end-­‐products  Alfy     PI3P-­‐binding  autophagy-­‐linked  FYVE  domain  protein  ApoE     apolipoprotein  E  APP     Aβ  precursor  protein  ASK1     apoptosis  signal  regulation  kinase-­‐1  ATG     autophagy-­‐related  genes  ATF6     activating  transcription  factor-­‐6  Bchs     blue  cheese  Bcl-­‐2     B  cell  lymphoma-­‐2  BiP     binding  immunoglobulin  protein  CGRP     calcitonin  gene-­‐related  peptide  CHOP     C/EBP  homologous  protein  CMA     chaperone  mediated  autophagy  CPE     Carboxypeptidase  E  CRLR     calcitonin-­‐receptor-­‐like-­‐receptor  CSF     cerebrospinal  fluid  CT     calcitonin  CTR-­‐2     calcitonin  receptor  2  CVT     cytosol-­‐to-­‐vacuole  targeting  EDEM     ER  degradation-­‐enhancing  α1,2-­‐mannosidase  like  protein  EM     electron  microscopy  EOFAD     early-­‐onset  FAD  ER     endoplasmic  reticulum  ERAD     ER  associated  degradation  ERAF     ER  associated  folding  ERdj     ER-­‐resident  J-­‐domains  ERManI     ER  degradation  α1,2-­‐mannosidase  I  ESCRT     endosomal  sorting  complex  required  for  transport  FAD     familial  form  of  Alzheimer’s  disease  FADD     Fas-­‐associated  death  domain  GAGs     Glycosaminoglycans  GFP     green  fluorescent  protein  GS     glycogen  synthase  GSK3α     glycogen  synthase  3α  HDAC     histone  deacteylase  HFNs     human  fetal  neurons  hIAPP     human  IAPP  

   

 2  

HS     heparin  sulphate  Hsc     heat  shock  cognate  Hsf1     heat  shock  factor-­‐1  Hsp     heat  shock  protein  HSPG     heparan  sulphate  proteoglycan  HSR     heat  shock  response  Htt     Huntingtin  IAPP     islet  amyloid  polypeptide  IDE     insulin  degrading  enzyme  IRE1     inositol-­‐requiring  protein-­‐1  JNK     c-­‐Jun  N-­‐terminal  kinase  LAMP     lysosome-­‐associated  membrane  type  protein  LC3     microtubule  associated  protein  1  light  chain  3  mIAPP     murine  IAPP  MVBs     multivesicular  bodies  NEFA     non-­‐esterified  fatty  acids  NFT     neurofibrillary  tangles  NMR     nuclear  magnetic  resonance  OST     oligosaccharyltransferase  PAM     peptidyl  amidating  monooxygenase  PC     prohormone  convertase  PD     Parkinson’s  disease  PE     phosphatidylethanolamine  PERK     protein  kinase  RNA-­‐like  ER  kinase  PI3K     phosphatidylinositol  3-­‐kinase  PI3P     phosphatidylinositol  (3,4,5)-­‐trisphosphate  Poly-­‐Q     polyglutamine  PS1     presenilin-­‐1  RAMP     receptor  activity-­‐modifying  protein  ROS     reactive  oxygen  species  SAP     serum  amyloid  P  SDS     sodium  dodecyl  sulphate  TNFR1     tumor  necrosis  factor  receptor  1  TTR     transthyretin  TUNEL     terminal  deoxynucleotidyl  transferase  dUTP  nick  labelling  UAS     upstream  activating  sequence  ULK     Unc-­‐51-­‐like  kinase  UGGT     UDP-­‐glucose:glycoprotein  glucosyltransferase  UPR     unfolded  protein  response  UPRE     unfolded  protein  response  element  UPS     ubiquitin-­‐proteasome  system  Xbp1     X-­‐box  binding  protein-­‐1  YFP     yellow  fluorescent  protein  

   

   

                                     

Introduction

Introduction    

   

 4  

 

Protein  folding  and  misfolding  

One  of   the  most   fundamental  processes   in  biology   is   the  ability  of  a  protein  to   fold  into   its   defined   three-­‐dimensional   structure.   The   function   of   a   protein   is   tightly  coupled  to  this  defined  conformation.  Already  in  the  1950’s  Anfinsen  pointed  out  the  relationship   between   the   amino   acid   sequence   of   the   enzyme   ribonuclease   and   its  functional   conformation.   This   functional   conformation   could   be   destroyed   by   the  addition  of  8  M  urea  and  the  reducing  agent  β-­‐mercaptoethanol  but  as  soon  as  urea  was   removed   and   the   protein   re-­‐oxidized,   it   reassembled   into   its   native   structure.  The  free  energy  gained  in  this  assembly  drives  the  refolding  process  [1].  As  tribute  to  his  work  on  ribonuclease  Anfinsen  was  awarded  the  Nobel  Prize  in  1972.    The   native   state   of   a   protein   is   thought   to   be   the   most   stable   structure   under  physiological  conditions.  However  it  was  for  long  not  clear  how  this  structure  could  be  adopted  and  there  was  no  reasonable  explanation  for  the  Levinthal  paradox  [2].  The  basic   concept   introduced  by  Levinthal   is   that   the   search   for   the  proper   three-­‐dimensional  structure  is  a  random  “trial  and  error”  event.  If  a  protein  of  100  amino  acids  had  to  try  all  of  its  putative  conformations  (each  taking  10-­‐11  seconds  to  find)  the   calculated   time   for   this   exceeds   the   age   of   our   universe.   However,   from  experiments   we   now   know   that   folding   occurs   in   the   order   of   milliseconds   to  seconds.   This   time   discrepancy   is   known   as   the   Levinthal   paradox   [3].   Today,   the  current  concept   is   that  a  polypeptides  search   for   its  native  structure   is   following  a  “folding   funnel”   or   “folding   landscape”   with   the   native   structure   as   the   lowest  accessible  point.  Because,   on   average  native-­‐like   interactions   are  more   stable   than  non-­‐native   ones,   not   all   possible   conformations   have   to   be   tested,   instead   it   is  sufficient  to  test  a  small  number  of  possible  conformations.  The  shape  of  this  energy  landscape   is   encoded   in   the   amino-­‐acid   sequence   [4].   The   crowded   intracellular  milieu  with  a  protein  concentration  of  300-­‐400  mg/ml  complicates  protein  folding,  since  it  increases  the  risk  for  undesirable  interactions  with  other  molecules  [4,5].  A  way   to   circumvent   this   problem   is   the   engagement   of   folding   catalysts   and  chaperones.  They  function  either  by  accelerating  slow  folding  steps  or  by  protecting  partially  folded  proteins  from  misfolding  [6,7].  Despite  all  cellular  efforts  to  optimize  folding  can  protein  misfolding  occur.  In   fact,   accumulation   of   misfolded   proteins   can   have   detrimental   effects   on   the  organism,   and   is   indeed   linked   to   many   diseases,   including   amyloidosis.   This  dissertation   deals   with   various   aspects   of   misfolded   proteins   with   focus   on   the  amyloid   forming   islet  amyloid  polypeptide  (IAPP),  and  the  consequences   that  arise  when  cells  are  exposed  to  misfolded  IAPP.      

Introduction    

   

  5  

 

Amyloid  and  amyloidosis  

History  and  definitions  

In   1854   the   German   physician   Rudolph   Virchow   was   the   first   to   use   the   term  amyloid  (from  Latin  amylum  =  starch)  to  describe  the  macroscopic  changes  he  found  in   some  human  organs   after   they  had  been   treated  with   iodine   and   sulphuric   acid  [8].  At  this  time,  this  staining  method  was  widely  used  by  botanists  to  demonstrate  cellulose  [9].  Already  five  years  later,  Friedreich  and  Kekulé  were  able  to  show  that  amyloid   isolated   from   the   spleen  was   not   “starch-­‐like”  material   but   instead   it  was  mainly  made   up   by   protein   [10].  With   time,   new   staining  methods   evolved   and   in  1922   Bennhold   introduced   the   cotton   dye   Congo   red   as   a   histological   dye   for  amyloid   [11].   In   1927   Divry   and   Florkin   showed   that   Congo   red   emits   green  birefringence  when  observed  in  cross-­‐polarized  light  [12].  A  standardized  Congo  red  staining  protocol  was  introduced  in  1962  and  this  is  still  in  use  [13,14].  The  property  of   amyloid   to   emit   green   birefringence  when   stained  with   Congo   red   suggested   a  highly   ordered   structure,   which   was   confirmed   by   Cohens   and   Calkins   electron  microscopy   studies   on   amyloid   fibrils.   They   showed   that   amyloid   is   made   up   of  unbranched   fibrils   with   a   diameter   of   approximately   10   nm   and   undetermined  length  [15].  Further  research  revealed  that  all  amyloid  fibrils  are  made  up  of  smaller  sub-­‐elements,   named   protofibrils,   a   finding   that   proved   to   be   independent   on   the  protein  constituent  of  the  amyloid  [16].  X-­‐ray  diffraction  analysis  was  used  by  Eanes  at  al.  to  define  the  well-­‐ordered  cross-­‐β-­‐sheet  pattern  of  amyloid  fibrils  [17].    In  order  to  be  defined  as  amyloid,  following  criteria  have  to  be  fulfilled:  

1. In  vivo  deposited  material  2. Affinity   for   Congo   red   and   presentation   of   green   birefringence   when  

viewed  in  polarized  light  3. The   characteristic   fibrillar   structure   when   investigated   with   an   electron  

microscope  4. A  specific  X-­‐ray  diffraction  pattern  of  the  fibril  

 All  stated  criteria  follow  the  consensus  reached  at  the  meeting  of  the  Nomenclature  Committee   of   the   International   Society   of   Amyloidosis   in   November   2006.   During  this  meeting  one  previous  characteristic  of  amyloid  was  actually  revised.  Due  to  the  increasing   evidence   of   intracellular   amyloid,   the  definition   of   amyloid   is   no   longer  limited  to  extracellular  material  [18].    

Introduction    

   

 6  

Amyloid  and  diseases  

Today,  at  least  27  different  proteins  have  been  identified  to  form  amyloid  in  humans  and  the  heterogeneous  group  of  diseases  associated  with  such  deposits  is  referred  to  as   amyloidosis   [19].   Each   type   of   amyloidosis   is   characterised   by   a   distinct   fibril  protein   [18].   Despite   the   common   structural   features   of   amyloid   fibrils   exhibit  amyloid  proteins  only  modest  primary,   secondary  and   tertiary  structure  homology  [20,21].  Dependant  on  the  amyloid  distribution  the  disease  is  divided  into  localized  and  systemic  amyloidosis.    Amyloid   that   appears   at   a   single   site   or   in   one   tissue   type   is   called   localized  amyloidoses.   Typically,   these   deposits   occur   in   close   proximity   of   the   amyloid  protein   expression   site.   Localized   amyloidosis   are   often   linked   to   ageing,   e.g.   Aβ  deposition  in  Alzheimer’s  disease  or  IAPP  in  type  2  diabetes.    Amyloid  diseases  with  deposits  that  affect  several  organs  are  referred  to  as  systemic  amyloidoses.   The   amyloid   precursor   in   systemic   amyloidosis   is   a   plasma   protein.  Examples  of  systemic  amyloidosis  are  reactive  amyloidosis  or  secondary  amyloidosis  with  protein  AA  deposits  or  AL-­‐amyloidosis  with  light  chain  deposits  [18].      

Introduction    

   

  7  

Table  1:  Amyloid  fibril  proteins  and  their  precursors  in  human  [19].  

Amyloid  protein   Precursor  

Systemic   (S),  or   localized  (L)  

Syndrome  or  involved  tissue  

AL   Immunoglobulin   light  chain  

S,  L   Primary  Myeloma-­‐associated  

AH   Immunoglobulin   heavy  chain  

S,  L   Primary  Myeloma-­‐associated  

Aβ2M   β2-­‐microglobulin   S  L?  

Hemodialysis-­‐associated  Joints  

ATTR   Transthyretin   S   Familial  Senile  systemic  

AA   (Apo)serum  AA   S   Secondary,  reactive  AApoAI   Apolipoprotein  AI   S  

L  Familial  Aorta,  meniscus  

AApoAII   Apolipoprotein  AII   S   Familial  AApoAIV   Apolipoprotein  AIV   S   Sporadic,  associated  with  ageing  AGel   Gelsolin   S   Familial  (Finnish)  ALys   Lysozyme   S   Familial  AFib   Fibrinogen  α-­‐chain   S   Familial  ACys   Cystatin  C   S   Familial  ABri   ABriPP   S   Familial  dementia,  British  ALect2   Leukocyte   chemotactic  

factor  2  S   Mainly  kidney  

ADan   ADanPP   L   Familial  dementia,  Danish  Aβ   Aβ   protein   precursor  

(AβPP)  L   Alzheimer’s  disease,  ageing  

APrP   Prion  protein   L   Spongiform  encephalopathies  ACal   (Pro)calcitonin   L   C-­‐cell  thyroid  tumors  AIAPP   Islet   amyloid  

polypeptide  (also  called:  amylin)  

L   Islets   of   Langerhans   (type   2  diabetes)  Insulinomas  

AANF   Atrial  natriuretic  factor   L   Cardiac  atria  APro   Prolactin   L   Ageing  pituitary  

Prolactinomas  AIns   Insulin   L   Iatrogenic  AMed   Lactadherin   L   Senile  aortic,  arterial  media  AKer   Kerato-­‐epithelin   L   Cornea,  familial  ALac   Lactoferrin   L   Cornea  AOaap   Odontogenic  

ameloblast-­‐associated  protein  

L   Odontogenic  tumors  

ASemI   Semenogelin  I   L   Vesicula  seminalis    

Introduction    

   

 8  

Structure  of  amyloid  

The   high-­‐resolution   structures   of   different   in   vitro   assembled   amyloid-­‐like   fibrils  have  been  solved.  The  primary  building  block  of  the  fibrils,  the  actual  protein,  gives  rise   to   two,   or   more,   β-­‐strands   that   run   perpendicular   to   the   fiber   axis.   Amyloid  fibrils  are  easily  identified  when  viewed  in  an  electron  microscope  [22].  The  highly  ordered,   repetitive   composition   of   the   fibrils   give   rise   to   a   characteristic   X-­‐ray  diffraction  pattern  with  an   inter-­‐β-­‐strand  distance  of  4.7Å  and  a  distance  of  6-­‐11Å  between   stacked   β-­‐sheets.   Association   of   2-­‐6   protofilaments,   each   2.5-­‐3.5   nm   in  diameter,  forms  fibrils  (see  Figure  1).  By  twisting  around  one  another  along  the  fiber  axis,   these   protofilaments   contribute   to   the   rigidity   of   the   amyloid   fibril   [23].  Amyloid   fibrils   from   the   same   protein   are   able   to   form   different   morphologies,  depending  on  the  surrounding  conditions  [24].  Solid-­‐state  NMR  and  EM  images  have  supported   the   idea  of   structural  polymorphism   in  amyloids   [25,26].  Different   local  minima   in   the  energy   landscape  of   the  unfolded  amyloid  protein  are  accounted   for  this  diversity   in  vivo   [27].  The  structural  heterogeneity  of   fibrils   includes  degree  of  twisting,  the  number  of  filaments  per  fibril,  and  the  diameter  or  mass  per  length  of  the  fibrils  [25,26].  

Figure  1:   Structure   of   the   amyloid   fibril.  The  β-­‐strands  of   the  amyloid  protein  are  stacked  perpendicular  to  the  fiber  axis.  The  intermolecular  distance  of  β-­‐strands  of  neighbouring  units  is  4.7Å.  Two  to  six  protofilaments  twist  around  each  other  and  give  rise  to  the  mature  amyloid  fibril.  

Introduction    

   

  9  

Non-­‐fibrillar  components  in  amyloid  deposits  

The  major   amyloid   constituent   is   the   disease-­‐specific   fibril   protein.   In   addition   to  this   fibril   protein   other,   non-­‐fibrillar   components   are   present,   such   as  Glycosaminoglycans,   Serum   amyloid   P   (SAP)   component   and   Apolipoprotein   E  (ApoE).    Glycosaminoglycans   (GAGs)   are   negatively   charged   heteropolysaccharides  composed   of   repeating   disaccharide   units.   The   structure   of   the   repeating  disaccharide   unit   defines   the   five   GAG   classes,   namely   heparin/heparin   sulphate  (HS),  chondroitin  sulphate,  dermatan  sulphate,  hyaluronan,  and  keratan  sulphate.  All  GAGs   except   for   hyaluronan   are   usually   found   covalently   linked   to   a   protein  backbone   and   this   complex   is   then   called   proteoglycan.   In   the   light   of  amyloidogenesis   are   heparan   sulphate   and   the   heparan   sulphate   proteoglycan  (HSPG)   perlecan   the   best   studied   GAG   and   proteoglycan.   Numerous   in   vitro  experiments  showed  the  potential  of  GAGs  and  HSPGs  to  promote  fibril  formation  by  increasing  the  β-­‐sheet  content  of  the  amyloidogenic  protein.  It  is  also  reported  that  HS   is   involved   in   processing   of   the   amyloid   precursor   proteins   and   thereby  influencing   fibril   formation  kinetics  and/or   toxicity   [28,29].  Experiments   in  animal  models   affirm   an   active   role   for  HS   in   amyloidogenesis   [30,31].   The   interaction   of  GAGs  and  amyloid  is  a  target  for  drug  therapy  [32,33,34].    Serum   amyloid   P   component   belongs   to   the   pentraxin   superfamily   and   binds  amyloid  fibrils  in  an  calcium-­‐dependent  manner  [35].  The  binding  of  SAP  to  amyloid  fibrils  is  suggested  to  prevent  proteolysis  of  amyloid  fibrils  [36].  Due  to  its  high  and  specific   affinity,   radiolabelled   SAP   is   used   to   monitor   amyloid   deposits   in   a   non-­‐invasive  manner  [37].    Apolipoprotein   E  has  been  detected   in  association  to  numerous  amyloid  deposits,  including   IAPP   derived   islet   amyloid   and   amyloid   deposits   of   Alzheimer’s   disease  [38].  However,  the  exact  role  of  ApoE  in  amyloidogenesis  is  unclear.  Polymorphisms  in   the   APOE   gene,   ε2,   ε3,   and   ε4   strongly   alter   the   likelihood   of   developing  Alzheimer’s   disease   and   cerebral   amyloid   angiopathy.   It   has   been   suggested   that  ApoE  modulates  Aβ  metabolism  and  accumulation,  although  there  are  contradictive  results  on  plaque  density  or  number  depending  on  the  APOE  genotype.  Differential  effects  of  APOE  isoforms  on  lipid  metabolism  have  been  assigned  a  role  in  synaptic  plasticity  and  neurodegeneration,  independent  of  interactions  with  Aβ  [39].      

Introduction    

   

 10  

Amyloid  formation  

In  vitro,  many  proteins  are  capable  of  forming  amyloid-­‐like  fibrils  if  exposed  to  low  pH,  high  temperature,  high  pressure,  and/or  presence  of  co-­‐solvents  that  all  reflect  unphysiological   circumstances   [40].   In   case   of   some   globular   proteins,   such   as  lysozyme,   superoxide   dismutase   1,   and   transthyretin,   denaturing   conditions   are  close   to   physiological,   but   despite   this   can   amyloid-­‐like   fibrils   form   in   vitro.   It   is  thought  that  aggregation  in  these  cases  is  a  direct  consequence  of  fluctuations  from  the   native   state   or   other   local   unfolding   events,   and   does   not   require   global  unfolding   [41].   Amyloid-­‐like   fibril   formation   is   in   general   thought   to   occur   via   a  nucleation-­‐dependant   mechanism,   resembling   crystallisation   kinetics   [42,43].   A  typical   feature   of   a   nucleation-­‐dependant  mechanism   is   the  presence  of   a   lag   time  before   bigger   aggregates   are   detectable.   During   the   lag   phase   monomers   self-­‐assemble  and  form  oligomers  that  can  act  as  nuclei  for  further  fibrillization.  The  self-­‐assembly   of   monomers   requires   partially   unfolding   of   the   protein   and   is  thermodynamically   unfavourable   [44].   This   step   only   occurs   if   a   critical  concentration   is   exceeded.   The   lag   phase   is   followed   by   an   elongation   phase.  During  this  period  protofibrils  are  formed  that  rapidly  assemble  into  fibrils  and  grow  as   long   as   the   concentration   of   available   monomers/oligomers   is   sufficient.  Equilibrium  of  monomers  and  fibrils  characterises  the  final  plateau  phase.  The  time  span  of   the   lag  phase   can  be   significantly   reduced  by   addition  of   nuclei   in   form  of  preformed   oligomers   and/or   fibrils,   a  mechanism   referred   to   as   “seeding”   [43,45]  (see  Figure  2).  Seeding  is  also  an  in  vivo  finding  [46,47,48,49].    

 Figure  2:     Illustration  of  kinetics  of  amyloid   formation.  Addition  of  preformed  fibrils  and  protein  aggregates  can  shorten  the  lag  phase  (seeding  effect).      Events   that   can   lead   to   nucleation   in   vivo   are   interactions   between   the   amyloid  protein  and  cell  membranes,  increased  protein  synthesis  and  deficiencies  in  protein  clearance  [50]  (see  Toxic  effects).  

Introduction    

   

  11  

Toxic  effects  

In   general,   diseases   associated  with   amyloid   are   of   late   onset   and   actual   deposits  have   degenerative   effects   [45].   The   role   of   amyloid   in   different   diseases   has   been  subject   of   discussion   over   a   long   period   and   during   the   last   decade   many   new  insights   into   structural   properties   of   amyloid   fibril   precursor   species   have   shed   a  new   light   on   how   to   think   about   amyloid   cytotoxicity.   In   2006,   the   year   this   PhD  thesis  was  initiated,  it  was  believed  that  amyloid  cytotoxicity  is  coupled  to  common  mechanism  independent  of  protein  or  peptide.  Until  then,  several  in  vitro  studies  had  shown  that  oligomeric  species  and/or  protofibrils  of  several  amyloid  proteins  were  able  to  permeabilize  cell  membranes,  resulting   in  cell  dysfunction  [51,52,53,54,55].  In   the   same  year  Cohen  et   al.  were   able   to  demonstrate   in   a  C.  elegans  model   that  protofibrils  of  Aβ  were  toxic,  whereas  high  molecular  weight  Aβ  aggregates  were  not  [56].  Today,  oligomers  are  still  seen  as  the  major  cause  for  cytotoxicity.  Over  the  last  few   years   there   has   been   growing   evidence   for   the   concept   that   the   same  amyloidogenic  peptide/protein  can  give  rise  to  structurally  different  oligomers  and  structural  distinct  fibrils.  This  led  to  the  proposal  of  an  aggregation  energy  landscape  with  several  local  energy  minima  corresponding  to  distinguishable  oligomeric  states  [50].   But   toxicity   is   not   only   thought   to   be   dependent   on   the   structure   of   the  oligomeric   species   but   also   on   the   biophysical   and   biochemical   properties   of   the  interacting   membrane.   Anionic   surfaces   (e.g.   anionic   phospholipid-­‐rich   liposomes,  glycosaminoglycans)   seem   to  play   an   important   role   as  potent   triggers   for  protein  fibrillization.  Also  mature  fibrils  can  be  ascribed  certain  toxicity  since  the  deposited  amyloid   can   be   massive   and   affect   exchange   of   oxygen   and   nutrients.   Moreover,  mature  fibrils  might  contribute  to  cytotoxicity  by  leakage  of  toxic  oligomers  [50].    When  it  comes  to  IAPP  it  is  still  unclear  if  toxic  oligomers  exist  in  vivo.  In  vitro,  beta  cell  toxicity  has  been  shown  in  the  presence  of  freshly  solubilized  IAPP  and  this  leads  to  activation  of  apoptosis  [55,57,58].  On  the  other  hand  have  different  studies  shown  that  even  pre-­‐formed  IAPP  fibrils  induce  beta  cell  death  [59,60].  A  recent  study  could  show   that   there  exists  a   significant   relation  between   the  amount  of  deposited   islet  amyloid   and  measured  beta   cell   apoptosis.  This   latter   study   strongly   suggests   that  islet  amyloid  deposition  contributes  to  beta  cell  death  [61].  The   inhibitory  effect  of  amyloid   inhibitors   on   beta   cell   death   further   challenges   the   concept   of   toxic  oligomers   (reviewed   in   [62]).   The   oligomeric   state   might   be   transient,   and   this  complicates  the   interpretation  of   the   in  vitro  assays  where  cells  are   incubated  with  oligomers.   If   cells   are   incubated   for   longer   times   with   oligomers,   these   oligomers  might  alter   their   structure  and  start   fibrillization.  So   in  order   to  be  able   to  ascribe  toxicity  to  oligomers  it  is  crucial  to  make  sure  that  these  oligomers  are  stable.    An  alternative  pathway  for   IAPP  toxicity  has  been  suggested  by  Engel  et  al..   In   this  model,  IAPP  binds  to  membranes,  which  results  in  fibril  growth,  significant  changes  of   membrane   curvature   and   will   over   time   lead   to   physical   breakage   of   the  membrane.   Notably,   the   kinetic   profile   of   hIAPP   fibril   formation   matched   that   of  membrane   leakage   [63].   The   model   of   membrane   interaction   as   crucial   step   in  

Introduction    

   

 12  

mediating   toxicity  might  be  of   general  nature.  Membranes   can   serve  as  a   template  that   allow   orientation   of   monomers   in   a   way   that   favour   aggregation   [64].   In  addition  membrane  interaction  of  amyloidogenic  proteins  can  lead  to  increased  local  protein   concentration   and   thereby   catalyse   aggregation   [65].   Finally,   it   has   been  shown  that  membranes  have  the  ability  to  alter  the  conformation  of  a  protein  and  in  this  way  induce  aggregation  [66,67].  Taken  together  results  from  different  studies  that  all  tried  to  identify  toxic  species  of  amyloidogenic   proteins,   it   becomes   clear   that   aggregation   pathways   have   a  major  influence   on   how   toxicity   is   mediated.   Since   these   aggregation   pathways   not  necessarily   are   the   same   for   different   amyloid-­‐related   peptides,   we   have   to  reconsider   the   concept   that   there   exists   a   general   mechanism   that   accounts   for  toxicity.    In  parallel  to  the  attempt  of  identifying  a  toxic  amyloid  species,  several  groups  have  started   to   look   at   molecular   pathways   that   might   be   altered   upon   protein  aggregation   and   subsequent   amyloid   formation.   Several   pathways,   such   as  autophagy,   endoplasmic   reticulum   associated   degradation   (ERAD)   and   unfolded  protein   response   (UPR),   have   been   identified   to   be   triggered   upon   protein  aggregation   (intra-­‐  and  extracellular)  and  a  more  detailed  overview  of  our  current  knowledge  how  these  pathways  influence  cell  survival  is  given  in  a  separate  section  of  this  introduction  (see  Molecular  pathways  connected  to  protein  misfolding).  

Functional  amyloid  

Since   many,   structurally   unrelated   proteins   are   capable   of   forming   amyloid-­‐like  fibrils  in  vitro,  it  has  been  speculated  that  amyloid  structures  have  been  a  prominent  fold   in   early   life   [68].   In   coherence   with   this   speculation,   the   field   of   functional  amyloid  has  evolved  over  the  last  decade.  Originally   it  was  hypothesised  that  some  organisms   have   during   evolution   taken   advantage   of   the   widespread   potential   of  proteins   to   fold   in   a   stable,   amyloid-­‐like   manner   [69].   Today,   several   functional  amyloid  structures  are  reported  in  lower  organisms,  including  curly  and  chaplins  in  bacteria  [70,71],  Sup32p  and  Ure2p  in  fungi  [72,73],  and  chorion  in  insects  [74].  In  aplysia   (sea   slug)   conversion   of   CBEP   to   an   amyloid-­‐like   structure   has   been  suggested   to   play   a   functional   role   in   memory   storage   [75].   In   humans   Mα,   a  component  of  Pmel17,  has  been  described  to  play  a  role  as  functional  amyloid  as  it  serves   as   template   for  melanin   and   thereby   is   involved   in  melanin   polymerisation  [76].  Maji  et  al.  suggested  in  2009  that  peptide  and  protein  hormones  are  stored  in  secretory   granules   in   an   amyloid   like   aggregation   state   [77].   Their   hypothesis   is  based   on   different   in   vitro   experiments   in   which   they   showed   how   31   of   42  investigated   protein   hormones   formed   amyloid-­‐like   structures   at   pH   5.5   in   the  presence  of  heparin  -­‐  conditions  that  mimic  the  environment  of  secretory  granules.  In   addition,   they   also   investigated  mouse   pituitary   tissue   and  were   able   to   detect  

Introduction    

   

  13  

amyloid   like   structures.   The   proposed   working   model   is   that   either   a   critical  concentration   in   the   Golgi   per   se   and/or   processing   of   prohormones   can   trigger  amyloid   formation.  As  a   result,  hormones  can  be  packed   in  secretory  granules  at  a  highest  density  possible  and  even  be  stored  over  long  periods  due  to  high  stability  of  the  amyloid  entity.  At  the  same  time,  the  secretory  granules  could  serve  as  an  “inert”  membrane   container   protecting   the   cell   of   putative   toxic   effects   of   the   formed  amyloid.   In   this  model,   amyloid   fibrils  will   be   destabilized   once   they   are   released  from  the  secretory  granules  and  are  exposed  to  pH  7.4  [77].  Unfortunately,  neither  insulin  nor  IAPP  were  part  of  the  investigated  protein  hormones  though.  It  is  known  however,  that  IAPP  fibrils  are  extremely  stable  and  generally  need  harsh  conditions  for   depolymerisation   [78].   It   is   questionable   if   secreted   IAPP   fibrils   are   able   to  dissolve  once  they  are  secreted  from  β-­‐cells.            

Islet  amyloid  polypeptide  (IAPP)  

General  introduction  

Eugene  Opie   reported   in  1901  a  hyaline   substance   to   replace  areas  of   the   islets  of  Langerhans  in  autopsy  material  from  a  patient  with  type  2  diabetes  [79].  Already  in  1973   the   characteristic   interaction   of   extracellular   amyloid   fibrils   with   β-­‐cell  membranes  was  described  [80].  But   it  was  not  until  1986  the  amyloid  protein  was  sequenced   and   for   the   first   time   fully   characterised   as   37   amino   acid   residue  polypeptide   [81,82].   This   peptide   was   initially   being   called   islet   amyloid   peptide  (IAP)  and  later  islet  amyloid  polypeptide  (IAPP).  Short  after  the  very  first  description  of   IAPP,  a  second  report  was  published  describing   the  same  polypeptide  naming   it  diabetes  associated  peptide  (DAP)  and  later  amylin  [83].    The  gene  for  IAPP  consists  of  3  exons  of  which  exon  one  is  non-­‐coding.  It  is  situated  on   the   short   arm   of   chromosome   12   and   has   a   promoter   region   similar   to   the  promoter  region  of  insulin  [84,85,86,87].  IAPP  belongs  to  the  calcitonin  gene  peptide  family   together   with   calcitonin   (CT),   calcitonin   gene-­‐related   peptide   (CGRP),  intermedin  and  adrenomedullin  [88].  Sequence  homology  of  hIAPP  with  CGRP-­‐I  and  II  is  43-­‐46%  and  with  human  CT  20%  [89,90].    IAPP  is  mainly  expressed  in  the  beta  cells   in  the  islets  of  Langerhans.  Here,   IAPP  is  stored   in   secretory   granules   together   with   insulin   and   those   hormones   are   co-­‐

Introduction    

   

 14  

secreted  upon  stimulation  [91,92,93].  The  intra-­‐granular  concentration  of  IAPP  is  1-­‐4   mM   and   the   insulin   concentration   is   10-­‐40   times   higher   [94,95].   The   plasma  concentration   of   IAPP   ranges   between  2-­‐10  pM   [96].   Expression   of   IAPP  has   been  found   in  mammals,   avian,   and   the  bony   fish   [94,97,98,99,100,101,102].   In   rodents,  expression  of   IAPP  was  also   reported   in  delta   cells   in   the   islets  of  Langerhans,   the  gastrointestinal   tract,   in   sensory   neurons   and   in   the   central   nervous   system  [103,104,105].      Over   the   years   several   different   biological   functions   have   been   ascribed   to   IAPP.  These   functions   include   auto-­‐   and   paracrine   effects   in   the   islets   of   Langerhans,  actions   as   a   satiety   peptide   in   the   brain,   antagonising   insulin   action   in   skeletal  muscles  and  also  a  role  in  calcium  homeostasis  in  regard  to  bone  mass.  Each  of  these  different  functions  is  briefly  highlighted  below.    Auto-­‐   and   paracrine   effects   of   IAPP   are   reported   to   regulate   insulin   secretion.  Autocrine  actions  include  a  dual  role  for  IAPP  on  insulin  secretion.  Transgenic  mice  that   are   deficient   for   IAPP   show   normal   basal   levels   of   circulating   insulin   and  glucose.  However,  these  knock-­‐out  mice  have  increased  insulin  responses  and  blood  glucose   elimination   upon   glucose   administration   when   compared   to   wild   type  controls.  It  can  be  concluded  that  usually  IAPP  limits  the  degree  of  glucose-­‐induced  insulin  secretion  [106].  Studies  about  5  years  later  gave  a  more  differentiated  picture  of  IAPPs  role  in  insulin  secretion.  Akesson  et  al.  detected  a  modest  increase  of  basal  insulin   secretion   in   the   presence   of   low   IAPP   concentrations   (10-­‐10   –   10-­‐6   M)   and  physiological  glucose  concentrations  (7  mM).   In  contrast,  high   IAPP  concentrations  (10-­‐6   –   10-­‐5  M)   inhibited   glucose   stimulated   (10  mM  &  16.7  mM)   insulin   secretion  [107].  In  addition  it  has  been  shown  that  IAPP  acts  in  a  paracrine  manner  on  alpha-­‐  and   delta-­‐cells   and   suppresses   glucagon   and   somatostatin   release,   respectively  [107,108].  The  observed   inhibitory   effect   of   IAPP  on  glucagon   release  was   already  seen  at  low  concentrations  (10-­‐10  and  10-­‐8  M)  [107].      Today,  IAPP  has  also  been  identified  as  a  satiety  hormone.  This  action  was  a  matter  of  discussion  but  the  identification  of  receptor  activity-­‐modifying  proteins  (RAMPs)  was  a  major  break-­‐through  [109,110].  McLatchie  et  al.   showed  that  RAMPs,  single-­‐transmembrane-­‐domain  proteins,   can  bind   to   the  Calcitonin-­‐receptor-­‐like   receptor  (CRLR).   This   binding   and   hence   newly   formed   RAMP:CRLR   complex   has   high  affinities   for   substrates   that   do   not   bind   CRLR   alone.   If   any   of   the   three   RAMPs  (RAMP-­‐1,   -­‐2,  or-­‐3)  binds  to  calcitonin  receptor  2  (CTR-­‐2),  a  class  of  receptors  with  affinity  for  IAPP  is  formed  [109,111,112,113].  It  is  not  clear  if  effects  of  IAPP  in  the  brain   are   due   to   local   expression   in   neurons   or   if   IAPP   crosses   the   blood-­‐brain  barrier  [114].  Effects  of  IAPP  on  glycaemic  control  have  also  led  to  the  development  of  pramlintide  (symlin).  Pramlintide  is  a  hIAPP  analogue  with  proline  substitutions  at  position  25,  28,   and  29.  The  proline   substiutions   abrogate   the   capacity   to   form  amyloid   fibrils.  

Introduction    

   

  15  

This  hIAPP  analogue   is   today  an  approved  drug   for  use   in  conjugation  with   insulin  therapy  in  patients  with  type  1  or  type  2  diabetes.  Furthermore  reveal  preliminary  data  a  weight  loss  in  obese  patients  with  and  without  diabetes  upon  symlin  intake  [115,116,117].    A   recent   study   in   rats   suggests   a   role   for   IAPP   in   maternal   regulations   and   IAPP  mRNA  was  up-­‐regulated  in  the  preoptic  area  of  the  hypothalamus  of  lactating  dams  [118].  IAPP  also  has  been  attributed  a  role  in  reducing  pain  [119,120].  In  skeletal  muscles  IAPP  has  been  found  to  inhibit   insulin-­‐stimulated  incorporation  of  glucose   into  glycogen.  The  effect   is  described   to  occur  via   inhibition  of  glycogen  synthase  (GS)  and  activation  of  glycogen  phosphorylase  (GP),   [121,122].   Insulin  on  the   other   hands   stimulates   dephosphorylation   of   GS   thereby   promoting   glycogen  synthesis.   These   effects   of   IAPP   that   are   contrary   to   insulin   action   on   skeletal  muscles,  are  accounted  for  playing  a  role  in  developing  insulin  resistance  [123].    Finally,   I   want   to   mention   IAPPs   effect   on   calcium   homeostasis.   Infusion   of   IAPP  decreases  circulating  levels  of  calcium  in  humans  [124].  Mice  deficient  for  IAPP  show  a  50%  reduction  in  bone  mass  when  compared  to  wild-­‐type  littermates;  an  effect  due  to  increased  bone  resorption  mediated  by  IAPP  [125].    

Prohormone  processing  

Biological   mature   human   IAPP   derives   from   proteolytic   cleavage   of   the   89   amino  acid   hormone  preproIAPP.  The   first   22   amino   acids   account   for   the   signal   peptide  and   are   cleaved   off   after   entrance   into   the   endoplasmic   reticulum   (ER)   [126].   The  remaining,  67  amino  acid  long,  proIAPP  enters  the  secretory  pathway  and  there  it  is  cleaved  at  its  C-­‐terminal  and  N-­‐terminal  site,  giving  rise  to  mature  IAPP  (see  Figure  3)  [126,127,128,129].    Processing   of   proIAPP   is   sequential   and   occurs   first   at   the   C-­‐terminal   site   where  prohormone   convertase   (PC)   1/3   cleaves   at   di-­‐basic   amino   acid   residues   K50-­‐R51  [128,130].   In   the   secretory   granules   PC2   removes   the  N-­‐terminal   flanking   peptide  processing  after  di-­‐basic  residues  K10-­‐R11  [129,130].  Notably,  in  absence  of  PC  1/3  is  PC  2  capable   to  cleave  at   the  C-­‐terminal  processing  site.  This  redundancy  does  not  work   the   other   way   round.   Removal   of   the   N-­‐terminal   flanking   can   solely   be  achieved   by   PC   2   [128].   Carboxypeptidase   E   (CPE)   removes   the   dibasic   residues  lysine  and  arginine  at  the  C-­‐terminus  of  processed  proIAPP.  The  exposed  glycine  is  carboxyamidated   by   the   peptidyl   amidating   monooxygenase   (PAM)   complex.  Presence   of   active   CPE   is   also   necessary   in   order   to   facilitate   processing   at   the  N-­‐terminal  site  by  PC  2  [131].    

Introduction    

   

 16  

Both  prohormone  convertase  are  produced  as  precursor  molecules  themselves  and  have  to  undergo  cleavage  events  in  order  to  become  fully  active.  PC  1/3  is  first  auto-­‐catalytically   cleaved   at   its   N-­‐terminus.   This   occurs   already   in   the   ER.   In   mature  secretory  granules  PC  1/3  is  additionally  cleaved  at  the  C-­‐terminal.  This  site-­‐specific  maturation   of   PC   1/3  may   explain   the   observed   granule-­‐specific   processing   by   PC  1/3.   At   the   same   time,   a   partial   activation   of   PC   1/3   in   the   late   TGN   was  demonstrated   and   it   was   shown   that   C-­‐terminal   cleavage   of   proIAPP   is   already  initiated  in  the  TGN  before  entering  secretory  granules  [130,132,133,134].  Sorting  of  PC2  starts  in  the  ER  where  7B2  binds  to  proPC  2  and  enables  relocalization  of  proPC  2   to   the   TGN.   The   binding   of   proPC   2   to   7B2   requires   proper   folding   of   proPC   2  [135,136].  ProPC  2  is  finally  cleaved  in  the  secretory  granule,  a  prerequisite  to  gain  enzymatic  function  [134,137].      In  mature   IAPP,   a  disulphide  bridge   is  present  between  cysteine  2  and  7   [94,138].  Several  studies  show  that  impaired  processing  of  proIAPP  influences  fibril  formation  of   IAPP   [139,140,141,142,143].   The   implications   of   incomplete   processing   of  proIAPP  on  fibril  formation  are  discussed  below  (see  IAPP  fibril  formation).    

Figure  3:     Schematic   drawing   of   prohormone   processing.   Both   proIAPP   and  proinsulin  are  sequentially  processed  by  the  prohormone  convertases  2  and  1/3.  

 The   prohormone   convertases   that   process   proIAPP   also   sequentially   cleave  proinsulin  into  insulin.  Initially  PC  1/3  cleaves  at  two  arginines  at  position  31  and  32  (R31-­‐R32),  separating  the  B-­‐chain  from  the  C-­‐peptide  [144,145].  PC  1/3  cleavage  gives  rise  to  the  transient  intermediate  des-­‐31,32  proinsulin.  Thereafter,  PC  2  removes  the  C-­‐peptide   from   the  A-­‐chain  after   residues   lysine  64  and  arginine  65   [146].  Dibasic  residues  at  the  cleavage  sites  are  removed  by  CPE  [147].    

Introduction    

   

  17  

IAPP  and  type  2  diabetes  

Diabetes  is  today  classified  into  different  types:  type  1  and  type  2  diabetes,  maturity  onset   diabetes   in   young   (MODY),   ketosis   prone   diabetes   (KPD),   and   latent  autoimmune  diabetes  (LADA)  [148].  Type  1  diabetes  leads  to  insulin  deficiency  due  to   destruction   of   beta   cells   by   an   auto-­‐immune   reaction   and   usually   debuts   at   a  young   age.   The   majority   of   individuals   with   diabetes   suffer   from   type   2   diabetes  [149].    Deposition  of   IAPP-­‐derived   islet  amyloid   is  closely  associated  with  type  2  diabetes.  At  autopsy,   amyloid  can  be   found   in  a  great  majority   (up   to  95%)  of  patients  with  type   2   diabetes   [150,151,152].   Contrary,   in   healthy,   age  matched   individuals,   only  10-­‐20%   subjects   show   amyloid   deposition   in   the   pancreas,   and   the   amyloid   load  found   is   much   lower   when   compared   to   patients   with   type   2   diabetes   [152,153].  Type  2  diabetes  is  a  heterogeneous  disease  that  is  characterized  by  hyperglycaemia  [154].  The  prevalence  for  type  2  diabetes  increases  with  age  and  the  combination  of  several   risk   factors,   such   as   genetic   predisposition,   physical   inactivity,   and  obesity  have  been  pointed  out  as  determinants  in  developing  this  heterogeneous  disease.  An  initial  event   in   the  pathogenesis  of   type  2  diabetes   is  peripheral   insulin  resistance,  which  will  be  compensated  for  by  elevated  insulin  secretion  from  the  pancreatic  beta  cells  [155].  However,  type  2  diabetes  is  not  manifested  as  long  as  beta  cells  are  able  to   keep   insulin   levels   high   enough.   The   transition   of   beta   cells   not   being   able   to  secrete  sufficient  amount  of  insulin  and  thereby  leading  to  type  2  diabetes  is  referred  to  as   “beta   cell  decompensation”  or   “beta  cell   failure”   [156,157].  There   is  evidence  that   beta   cell   loss   via   increased   apoptosis   is   important   for   the   onset   of   type   2  diabetes  [152,158].  One  question   that   has  been  matter   of   debate   over   a   long   time   concerns   the   actual  role  of  islet  amyloid  in  type  2  diabetes.  Looking  at  a  pancreatic  section  from  a  patient  with   type   2   diabetes   where   almost   all   islets   are   replaced   by   amyloid   it   is   very  tempting  to  conclude  that  amyloid  most  certainly  will  affect  the  amount  of  secreted  insulin.  However,   today   it   is   impossible   to  say   if   the   formation  of   islet  amyloid   is  a  cause   or   a   consequence   of   type   2   diabetes.   The   circumstance   that   not   all   patients  with   type   2   diabetes   develop   islet   amyloid   is  most   likely   due   to   the  multifactorial  nature   of   the   disease.   It   does   not   exclude   the   possibility   that   aggregation   and  fibrillization   of   IAPP  might   be   of   major   importance   in   the   development   of   type   2  diabetes  in  a  considerable  large  subgroup  of  patients.  The  lack  of  in  vivo  techniques  to   track   islet   amyloid   formation   in   humans   leaves   the   question   if   islet   amyloid  deposition  precedes  type  2  diabetes  unanswered  [159].  Such  chronological  order  has  been   observed   in   baboons   though,   where   islet   amyloid   appeared   before  development  of  disease.  In  these  animals  the  amount  of  amyloid  correlated  well  with  raised  blood  glucose  levels,  a  good  indicator  for  the  progression  of  the  disease  [160].    In   animals   that   “spontaneously”  develop   type  2  diabetes,   i.e.  monkey   and   cat,   islet  amyloid  can  be  found.  This  is  in  contrast  to  animals  like  rats  and  mice,  which  do  not  spontaneously  develop  type  2  diabetes  and  neither  deposit  islet  amyloid.  As  a  matter  

Introduction    

   

 18  

of  fact  murine  IAPP  (mIAPP)  lacks  the  capability  to  form  amyloid  fibrils  (in  vivo  and  in   vitro)   [161].   The   reason   for   this   can   be   found   in   the   presence   of   three   proline  substitutions   in  mIAPP.  All   three  prolines  are   situated  between  amino  acid  20  and  29.   Synthetic   hIAPP   20-­‐29   is   extremely   amyloidogenic,   the   corresponding  murine  not  at  all  [162].  Proline  is  a  known  beta-­‐sheet  breaker  and  is  absent  in  the  primary  sequence  of  hIAPP.    However,  transgenic  mice  expressing  human  IAPP  (hIAPP)  that  were  crossbred  with  ob/ob  or  Agoutivy  mice  respectively  did  develop  islet  amyloid  in  response  to  insulin  resistance  with  hyperglycaemia  and  a  type  2  diabetic  phenotype  [163,164].  Increased  demands  of  insulin  secretion  also  imply  elevated  production  of  IAPP  since  both   hormones   are   co-­‐secreted.   It   is   possible   that   this   rise   in   IAPP   concentration  initiates   oligomerisation   and   IAPP   fibril   formation.   Numerous   in   vivo   studies   have  shown  that  IAPP  fibril  formation  can  cause  beta  cell  death  and  some  of  these  studies  identified  apoptosis  as  death  mechanism  [58,142,165,166,167].    Besides   the   presence   of   deposited   islet   amyloid   are   hyperproinsulinemia   and  elevated   levels   of   circulating   des   31-­‐32   proinsulin   hallmarks   of   type   2   diabetes  [168].  As  aberrant  processing  of  proinsulin  becomes  more  frequent,  is  also  aberrant  processing   of   proIAPP   expected,   a   circumstance   that   is   thought   to   accelerate  formation  of  islet  amyloid  (see  IAPP  fibril  formation).  In  Asian  population  a  serine  to  glycine  substitution  at  position  20  (S20G)  has  been  reported  and  is  associated  with  early  onset  of  type  2  diabetes  and  increased  risk  for  developing  diabetes  [169,170].  In  vitro  this  mutation  is  more  prone  to  form  amyloid-­‐like  fibrils  than  the  wild-­‐type  counterpart  [171,172].  

IAPP  fibril  formation  

In   order   to   understand   more   about   the   role   of   IAPP-­‐derived   amyloid   in   type   2  diabetic  patients  we  have   to   find  an  answer   to   the  question:   “why  does   IAPP   form  amyloid   in  patients  with  type  2  diabetes?”.  At   the  same  time  one  can  rephrase  this  fundamental  question  and  ask:  “which  mechanisms  prevent  fibril  formation  of  IAPP  under   normal   conditions?”.   Below,   some   of   the   results   are   presented   that   contain  clues  to  these  puzzling  questions.    The   structure   of   the   IAPP   monomer   is   not   determined.   This   is   due   to   the  circumstance   that   IAPP   in   an   aqueous   environment   spontaneously   aggregates   into  insoluble  fibrils  within  a  few  hours.  Structural  data  on  IAPP  in  monomeric  form  are  obtained  by  either  analysing  murine  IAPP,  the  addition  of  SDS,  or  the  binding  of  IAPP  to  a  membrane  or  insulin.  Results  from  several  NMR  experiments  suggest  IAPP  to  be  an  unfolded  protein,  however  residues  in  the  region  8-­‐19  can  dynamically  adopt  an  α-­‐helical  structure  [167,173,174].    

Introduction    

   

  19  

The   formation  of   a  N-­‐terminal  helix   is   thought   to  be   stabilized  by   interaction  with  insulin  [174].  More  recently,  three  different  conformational  preferences  for  hIAPP  in  solution  have  been  calculated  using  molecular  simulations  and  infrared  experiments.  Regarding   to   this   study,   the   most   stable   structure   of   hIAPP   is   an   extended  antiparallel  β-­‐hairpin  with  the  turn  region  comprising  residue  20-­‐23.  A  slightly  less  stable   structure   has   an   α-­‐helical   segment   spanning   residues   9-­‐17   and   a   short  antiparallel   β-­‐sheet   including   residues   24-­‐28   and   31-­‐35.   The   least   favourable  conformation  is  a  random  coil  structure  [175].  The  identification  of  several  stable  IAPP  structures  in  solution  is  very  interesting  for  several   reasons.   It  has  ben   suggested   that   IAPP  has   to   form  monomeric  β-­‐hairpins  that  can  aggregate  and  lead  to  fibril   formation  [176].  The  structure  described  to  be  most  stable  contains  such  a  β-­‐hairpin  and  the  turn  region  in  this  β-­‐hairpin  coincides  with  the  turn  region  found  in  protofilaments  of  IAPP  fibrils  [177].  This  could  explain  the  fast  aggregation  of  IAPP  in  vitro.    It  is  known  that  insulin  is  found  in  the  secretory  granules  together  with  IAPP.  Even  though  those  two  hormones  are  found  at  different  intra-­‐granular  localizations  -­‐  IAPP  resides  in  the  halo  region  of  secretory  granules,  whereas  insulin  form  a  crystal  in  the  core   region   –   I   want   to   mention   some   in   vitro   data   from   insulin-­‐IAPP   interaction  studies   as   they  exemplify  how   inhibition  of   IAPP  aggregation   could  work   [178].   In  vitro   insulin  can  inhibit  fibril  formation  of  IAPP  [178,179,180].  Insulin  is  thought  to  interact   with   IAPP   by   keeping   IAPP   in   its   α-­‐helical   conformation   [174].   Taken  together   these   findings   give   rise   to   a  model   in   which   IAPP   is   very   prone   to   form  amyloid  in  absence  of  an  inhibitor  impeding  the  formation  of  the  preferred  β-­‐hairpin  structure.  On  the  other  hand,  insulin  can  serve  as  inhibitor  for  IAPP  fibril  formation  by  stabilizing  another  naturally  occurring  structure  of  IAPP,  in  which  IAPP  forms  an  α-­‐helix   at   its   N-­‐terminus.   Parenthetically,   I   want   to   point   out   that   this   suggested  structure  of   IAPP  stabilized  by   insulin  resembles  very  much  the  structure  assigned  for  murine  IAPP  [175].    When   looking   closer   at   interactions   of   IAPP   with   insulin   one   can   estimate   the  complexity  of  how  the  environment  of  beta  cell   secretory  granules   influences   IAPP  fibril   formation   and/or   toxic   effects   of   such   events.   As   already  mentioned,   several  studies   have   shown   in   vitro   an   inhibitory   effect   of   insulin   on   IAPP   fibrillization  [178,179,180,181].   Brender   et   al.   recently   published   results   showing   insulin   to   be  capable   of   preventing   fiber-­‐dependant   membrane   disruption,   but   in   this   study  insulin  could  neither  prevent  the  formation  of  small  oligomers  on  the  membrane  nor  the  initial  phase  of  membrane  disruption  before  fibrillogenesis  [182].    Mice   expressing   hIAPP   but   not  mIAPP   (+hIAPP/-­‐mIAPP)   fed   on   a   diet   high   on   fat  develop   islet  amyloid.   In   these  mice   intra-­‐granular  amyloid   fibrils   can  be  detected.  This   led   to   the   assumption   that   IAPP   initially   forms   amyloid   intracellular   in   the  secretory   granules   resulting   in   cell   death.   Once   these   cell   have   disappeared   the  degradation   resistant   amyloid   will   be   found   extracellularly   and   act   as   seed   for  further   amyloid   formation   from   exocytosed   IAPP   [142].   These   in  vivo   data   further  

Introduction    

   

 20  

highlight   the   importance  of   intra-­‐granular  events   in  promoting  and   inhibiting   fibril  formation.  One   important   intra-­‐granular   event   thought   to   influence   fibril   formation   is   the  processing  of  proIAPP.  The  aberrant  processing  of  proIAPP  by  PC  1/3  and  PC  2  has  been   investigated   in   different   cell   lines   with   unique   prohormone   convertases  expression  profiles.  Only  cell   lines   in  which  proIAPP  was  not  completely  processed  contained   intracellular   amyloid   [143].   Our   group   also   was   able   to   detect  unprocessed  proIAPP  in  intra-­‐cellular  amyloid  fibrils  in  vivo  [142].  In  addition,  it  has  been   reported   that   processing   of   proIAPP   alters   the   binding   capacity   of   heparan  sulphate  to  the  amyloidogenic  protein.  A  binding  site  for  heparan  sulphate  (HS)  was  identified   in   the   N-­‐terminal   region   of   the   prohormone   –   a   region   not   present   in  processed   IAPP   [141].   Biophysical   studies   with   unprocessed   proIAPP   and  mature  IAPP   have   supported   these   findings.   This   HS   –   proIAPP   interaction   might   lead   to  local  high  concentrations  of   the  amyloidogenic  protein  and   initiate  oligomerisation  and  subsequently  fibril  formation.  Fibrils  formed  of  proIAPP  are  competent  to  seed  fibril  formation  of  mature  hIAPP  [139,140].  Heparan  sulphate  synthesis  is  thought  to  occur   in   the  Golgi,  generally  allowing   for   intra-­‐granular  HS-­‐proIAPP   interactions   in  vivo   [183].   These   results   overlap   with   the   above-­‐described   model   of   IAPP   fibril  formation  at  membranes  being  responsible  for  amyloid  toxicity  [63].  A   direct   interaction   of   IAPP   with   cell   membranes   has   been   suggested   to   locally  increase   IAPP   concentrations   and   cause   fibril   formation.   In   all   these   models  negatively   charged   lipids   increase   the   propensity   of   membranes   to   interact   with  IAPP  (reviewed  in  [167]).      Even   though   IAPP   fibrils   can   be   found   intracellular   it   is   yet   unclear   where   the  primary  step  of  IAPP  aggregation  occurs.  Especially  with  regards  to  IAPP  amyloid  in  connection  to  type  2  diabetes,  I  want  to  mention  three  extracellular  factors  that  are  associated  with  islet  amyloid  formation.  Chronic   hyperglycaemia   leads   to   non-­‐enzymatic   glycation   of   proteins   and   is  referred   to   as   advanced   glycation   end-­‐products   (AGE)   [184].   Non-­‐enzymatic  glycation  occurs  mainly  on  proteins  with  a   low  turnover  rate  such  as  haemoglobin  and  collagen.  The  short  half-­‐life  of  IAPP  (approximately  30  min)  therefore  makes  it  unlikely  that  AGE-­‐IAPP  exists   in  vivo  [185].  However,  deposited  islet  amyloid  might  be  glycated.  In  vitro,  such  AGE-­‐IAPP  amyloid-­‐like  fibrils  are  more  prone  to  seed  IAPP  fibril   formation   that   non-­‐glycated   IAPP   amyloid-­‐like   fibrils   [186].   These   results  argue  for  a  role  of  hyperglycaemia-­‐induced  glycation  in  amyloid  propagation  rather  than  initiation  of  IAPP  fibrillization.  Patients   with   type   2   diabetes   have   elevated   levels   of   non-­‐esterified   fatty   acids  (NEFAs)   in   plasma,   and   there   is   a   linear   correlation   with   levels   of   blood   glucose  [187,188].   In   this   context   it   is   noteworthy   that   transgenic  mice   expressing   human  IAPP  have   to   be   fed   a   diet   high   in   fat   in   order   to   develop   islet   amyloid.   The  mere  over-­‐expression  of  human  IAPP  in  mice  is  not  sufficient  for  islet  amyloid  formation  [189].  When  studied   in  vitro,  NEFAs  accelerate   IAPP   fibril   formation  without  being  

Introduction    

   

  21  

incorporated   into   fibrils   themselves   [190].   But   NEFAs   might   not   just   catalyse  fibrillization   of   IAPP   by   direct   interactions.   The  NEFAs   palmitine   and   oleate   dose-­‐dependently   induced   IAPP   expression   in   the   murine   pancreatic   β-­‐cell   line   MIN6  [191].  As  mentioned  above,  several  reports  assign  GAGs  and  HSPGs  an   important  role   in  catalysing  fibril  formation  of  amyloidogenic  proteins  [31,183].  As  HSPGs  are  a  major  component  of  extracellular  matrix  and  basement  membranes,  one  can  speculate  that  secreted   proIAPP   can   bind   to   proteoglycans   present   in   the   basement  membranes,  increase  local  IAPP  concentration  there  and  thereby  facilitating  fibril  formation.  This  is  supported  by  frequent  occurrence  of  perivascular  amyloid  deposition  [192,193].    

Transgenic  animal  models  with  hIAPP  

Today,   in   vivo-­‐detection   of   human   islet   amyloid   is   not   possible.   Collected   human  material  derived   from  either  autopsy  or   from  surgical   resection  does  not  allow   for  longitudinal   studies.  About  15   years   ago   several   groups   independently   established  transgenic  mouse  strains  as  models  for  islet  amyloidogenesis.  Mouse  strains  had  to  express  hIAPP  since  mIAPP  is  not  amyloidogenic.  In  most  strains  is  the  expression  of  hIAPP   under   control   of   the   rat   insulin   I   or   II   promoter   [192,194,195,196].   In   the  model  generated  by  Yagui  et  al.  hIAPP  expression  is  regulated  by  the  human  insulin  promoter   [197].   It   soon  became   clear   that   the  mere  hIAPP  overexpression  did  not  lead   to   amyloid   deposition  despite   elevated  plasma   concentrations   of   hIAPP   (2-­‐15  times).   These   mice   were   also   normoglycaemic   and   normoinsulinemic  [192,195,196,197].  However,  one  strain  contained  intracellular  amyloid  fibrils  in  the  beta  cell  secretory  granules  and  de  Koning  et  al.  found  hIAPP  in  secondary  lysosomes  [192,197].   Transgenic   mice   that   were   hemizygous   for   hIAPP   and   treated   with  growth   hormone   and   dexamethasone   contained   small   intra-­‐   and   extracellular  amyloid   deposits   [194].   When   male   mice   from   this   strain   were   homozygous   for  hIAPP  they  spontaneously  developed  hyperglycaemia,  diabetes  with  beta  cell  death  and  intra-­‐  and  extracellular  amorphous  IAPP  aggregates  by  8-­‐14  weeks  of  age.  After  20  weeks,   the  majority   of   the  male  mice   had   detectable   amounts   of   islet   amyloid.  Deposits  were  often  found  in  close  vicinity  to  the  nucleus  and  were  neighboured  by  swollen   mitochondria   [198].   An   interesting   observation   was   done   in   the   lab   of  Steven   Kahn.   When   hIAPP   transgenic   mice   were   fed   a   diet   high   in   fat,   the   mice  became  hyperglycaemic  and  in  80%  of  male  and  11%  of  female  mice  older  than  13  months   were   extensive   amounts   of   amyloid   found   [199].   This   emphasizes   the  importance  of  additional   factors   for   IAPP   fibrillization   [195,199].  Oophorectomy   in  this  strain  did  increase  the  occurrence  of  amyloid  in  female  mice  to  64%,  suggesting  a  protective   role  of  ovarian  products  on   IAPP   fibrillization   [200].  Crossbreeding  of  hIAPP  transgenic  mice  with  insulin  resistant  traits  also  promoted  amyloid  formation  and   persistent   hyperglycaemia   [163,164].   A   hIAPP   strain   deficient   for   mIAPP  

Introduction    

   

 22  

(+hIAPP/-­‐mIAPP)  developed  islet  amyloid  when  fed  a  diet  high  in  fat  [189].  In  these  mice   fibrils   made   up   of   proIAPP   were   found   in   the   halo   region   of   the   secretory  granules  [142].  Today  not  only  mice  but  also  rats  are  established  that  are  transgenic  for   hIAPP.   Rats   transgenic   for   hIAPP   (HIP-­‐rat)   that   are   hemizygous   for   hIAPP  develop  diabetes  within  5-­‐10  months,  accompanied  by  the  presence  of  extracellular  amyloid,  decreased  beta  cell  mass,  and  increased  beta  cell  apoptosis  [201].            

Aβ  

Alzheimer’s  disease  

Alzheimer’s  disease  (AD)   is   the   leading  cause  of  dementia   in   the  ageing  population  and   clinical   symptoms   for   AD   include   cognitive   alterations,   memory   loss   and  behavioural   changes.   Neurodegeneration,   initially   characterized   by   synaptic   injury  and   followed   by   neuronal   loss   has   a   causal   role   in   the   development   of   AD   [202].  Histopathologically  AD  is  hallmarked  by  the  presence  of  amyloid-­‐β   (Aβ)  containing  plaques  and  neurofibrillary  tangles  (NFT),  composed  of  hyper-­‐phosphorylated  forms  of  the  microtubule-­‐associated  protein  tau  [203].  Almost  105  years  have  passed  since  Alois   Alzheimer   presented   the   first   case   of   Alzheimer’s   disease   and   extensive  research   has   been   undertaken   since   then   in   order   to   understand   the  mechanisms  lying  behind  AD.  By   the   time   this   text   is  written   there  are  58212  articles   listed  on  Pubmed  all   in   one  way  or   the  other  dealing  with  AD.   Still   the  precise  mechanisms  leading   to   neurodegeneration   in   AD   are   not   completely   clear.   The  majority   of   AD  patients  develop  a   sporadic   form,  with  age  as  a  main   risk   factor  and  disease  onset  between  60  and  70  years  [204].  About  10-­‐15%  of  patients  have  a  genetically  linked  familial   form   of   AD   (FAD),   with  mutations   in   genes   such   as   Aβ   precursor   protein  (APP),  tau  and  presenilin-­‐1  (PS1).  FAD  patients  often  have  an  earlier  onset  of  disease  [202].  Many  of   the  mutations   found   in  patients  with  FAD   lead   to  Aβ  production  or  aggregation  and  hence  Aβ  has  been  ascribed  an  important  role  in  the  development  of  the  disease.  Today,  this  role  is  more  defined  in  the  amyloid  cascade  hypothesis  that  comprises   deposition   of   Aβ   as   trigger   for   neuronal   dysfunction   and   death   in   the  brain.  Aβ   is  a  product  of   step-­‐wise  cleavage  of  APP  and  comes   in  different   lengths,  depending   on   the   cleavage   pattern   of   α-­‐,   β-­‐,   and   γ-­‐secretase   [205].   Of   these   Aβ  peptides,   Aβ40   is   the   most   abundant   peptide   but   Aβ42   seems   to   be   essential   for  initiating   Aβ   aggregation   and   is   considered   central   to   the   amyloid   cascade  hypothesis   [206].   The   ratio   of   these   two   peptides   measured   in   the   cerebrospinal  

Introduction    

   

  23  

fluid  (CSF)  can  be  a  useful  measure  to  confirm  the  diagnosis  of  AD.  The  smaller  the  relative   portion   of   Aβ1-­‐42   is,   the   higher   is   the   risk   for   developing   the   disease  [207,208].   This   counter-­‐intuitive   finding   is   explained   by   the   idea   that   Aβ1-­‐42   is  trapped  in  amyloid  deposits  and  therefore  cannot  transit   from  the  brain  to  the  CSF  (“amyloid  sink”  hypothesis)  [206].    Even   though  mechanisms   leading   to   selective   neuronal   death   are   still   debated,   Aβ  oligomers   have   emerged   as   potential   culprit   in   causing   neurodegeneration   by  interference  with  synaptic  function  [202,209].  Oligomers  as  toxic  species  also  might  explain   the   finding   of  most   clinico-­‐pathological   studies,  which   fail   to   find   a   strong  correlation  between  Aβ  amyloid  plaque  burden  and  AD  severity  [210].  A  mutation  in  APP  (E693G),  known  as  the  arctic  mutation,  leads  to  early  onset  AD  and  the  formed  Aβ42   E22G   peptide   has   shown   to   form   increased   amounts   of   protofibrils   in   vitro  [211].    Mouse  models  transgenic  for  mutant  APP  in  combinations  with  mutant  PS1  (PS1  is  part   of   the   γ-­‐secretase   complex   and   involved   in   the   proteolytic   cleavage   of   APP),  recapitulate   several   of   the   neuropathological   characteristics,   but   it   is   notable   that  many  of  these  mouse  models  don´t  suffer  neurodegeneration  [212].  Lately,   the   second   hallmark   of   AD,   the   presence   of   NFTs   made   up   of   hyper-­‐phosphorylated   tau   has   gained   more   attention   again.   The   distribution   and  abundance  of  NFTs  correlates  well  with  clinical  symptoms  of  AD  [213].  Today  a  “tau  axis  hypothesis”   is   discussed   in   the  Alzheimer   field.  Regarding   to   this  model   there  are   three   different   ways   of   how   Aβ   and   tau   toxicity   interact   with   each   other   and  thereby   cause   AD.   In   a   hierarchical   model   is   Aβ   acting   on   tau   (hyper-­‐phosphorylation   of   tau   due   to   Aβ   amyloid   formation),   which   in   its   turn   mediates  toxicity  in  neurons.  In  a  second  model  is  tau  mediating  Aβ  toxicity.  Synergistic  toxic  effects  of  tau  and  Aβ  are  summarized  in  a  third  model  [203].  

Aβ  and  IAPP  

Aβ   and   IAPP   form   insoluble   amyloid   aggregates   in   AD   and   type   2   diabetes  respectively.   Several   epidemiological   studies   have   established   a   link   between   the  two  diseases,   showing   that  patients  with   type  2  diabetes  have  an   increased  risk   to  develop   AD   [214]   and   vice   versa   [215,216].   Both   proteins   are   thought   to   be  dependant   on   proper   degradation   and   clearance   rate   in   order   to   avoid   fibril  formation  and  interestingly,  insulin-­‐degrading  enzyme  (IDE)  degrades  Aβ  as  well  as  IAPP   [217,218].   In  vivo   and   in  vitro,  Aβ   and   IAPP  compete   for   IDE  binding  and  can  thereby   influence   the   respective   degradation   efficiency   [219,220].   Hence,  modulating   the   clearance   capacity   of   IDE   has   been   discussed   as   beneficial   in   the  treatment  of  AD  and  type  2  diabetes.  Still  the  mechanistic  basis  for  the  correlation  of  AD  and  type  2  diabetes   is  unclear.  Aβ  and  IAPP  have  25%  sequence  homology  and  50%   sequence   similarity   (see   Figure   4)   [221].   This   structural   similarity   prompted  

Introduction    

   

 24  

O’Nuallain  et  al.  to  test  whether  preformed  fibrils  of  one  of  the  two  peptides  (Aβ  or  IAPP)  can  serve  as  seed  for  the  other  peptide  and  accelerate  fibril  formation  (“cross-­‐seeding”).  Interestingly,  fibrils  made  up  of  Aβ1-­‐40  or  Aβ1-­‐42  respectively  have  the  same  capacity  in  seeding  IAPP  fibrillization  as  preformed  IAPP  fibrils  have.  Fibrils  made  up  of   IAPP  have  only  very  poor   seeding   capacity  on  Aβ   [221].  Therefore   it   is  unlikely  that  circulating  IAPP  fibrils  initiate  Aβ  aggregation  in  the  brain.  Jhamandas  et  al.  have  recently  shown  another  connection  between  Aβ  and  IAPP.  In  their   study   in   primary   cultures   of   human   fetal   neurons   (HFNs)   they   found   Aβ   to  mediate  toxicity  via  the  IAPP  receptor  (CTR  and  RAMP3)  and  that  down-­‐regulation  of  this  receptor  has  a  neuro-­‐protective  effect  [222].      

Figure  4:   Sequence  alignment  of   IAPP  and  Aβ40.   IAPP  and  Aβ40  share  a  25%  sequence   identity   (long   bars)   and   50%   sequence   homology   (short   bars)   (adopted  from  [221]).  

   

Introduction    

   

  25  

 

Drosophila  melanogaster  as  model  system  

History  of  Drosophila  as  model  system  

History   has   shown,   that   many   results   obtained   from   simple   organisms   like  Drosophila  melanogaster  have   invaluable   impact  on  the  understanding  of  biological  correlations  in  vertebrates  and  as  a  matter  of  fact  several  biological  concepts  as  we  know   them   today  would   lack   fundamental   insights  without   research   conducted   in  fruit   flies.   In   the   following   section   I   want   to   list   some   selected   examples   of   how  Drosophila  research  helped  us  to  get  a  better  understanding  of  biology.    Research  in  Drosophila  started  in  1908,  when  T.H.  Morgan  chose  this  model  for  his  studies  on  heredity.  In  1910  Morgan  had  found  an  eye  phenotype  (white  eyes)  that  was  sex-­‐linked  and  he  postulated  the  information  for  the  phenotype  to  be  found  on  the   X-­‐chromosome.   Besides   the   eye   colour,   several   other   X-­‐chromosome   derived  phenotypes,  e.g.  yellow  body  colour,  vermilion  eyes,  miniature  wings,  were  identified  over  the  next  years.  In  1913  Sturtevant  published  a  paper  in  which  he  put  different  genes   in   a   linear   order   [223].   This   and   other   results   were   the   basis   for   a  revolutionary  chromosome  theory  of  heredity  that   in  1933  was  awarded  the  Nobel  prize  (reviewed  in  [224]).      Notch  mutations  were   first   found   in  Drosophila   and  reported  already   in  1915.  The  systemic  search  for  mutations  causing  similar  phenotypes  as  Notch  mutations  led  to  the  identification  of  the  Notch  signalling  pathway.  In  the  1990s  it  became  clear  that  the  Notch  pathway  as  it  was  described  in  Drosophila  is  conserved  in  vertebrates.  The  Notch  pathway  has  a   fundamental   role   in  developmental  neurobiology  as   it   affects  almost   every   aspect   of   neurogenesis   and  differentiation   of   neurons   in   vertebrates,  both  in  the  developing  and  adult  brain  (reviewed  in  [225]).    Another   Nobel   prize   awarded   discovery   came   in   1927   when   Muller   was   able   to  demonstrate   the   mutagenic   effects   of   ionizing   radiation   on  Drosophila   [226].   The  possibility   of   inducing   mutations   was   expanded   in   1968   when   chemical  mutagenesis   with   ethyl   methane   sulfonate   was   introduced   to   the   scientific  community,   giving   rise   to   a   new   set   of   experiments   accelerating   the   functional  identification   of   many   new   genes   [227].   The   systematic   genome-­‐wide   mutation  screen  conducted  by  Nüsslein-­‐Vollhard  and  Wieschaus  led  to  the  discovery  of  most  major   signalling   pathways,   including   Hedgehog,   Tumor   growth   factor-­‐β ,   and  

Introduction    

   

 26  

Wingless.  In  1995  Nüsslein-­‐Vollhard  and  Wieschaus  were  awarded  the  Nobel  prize  in  recognition  for  their  work  (reviewed  in[225]).    Even   several   insights   into   behaviour   derive   from   discoveries   made   in  Drosophila.  Main  regulators  of  circadian  rhythm  were  first  identified  in  flies  and  the  isolation  of  the   genes   period,   timeless   and   clock   respectively,   was   an   essential   initiator   for  functionally  dissecting  the  circadian  rhythm  that   is  conserved  from  flies  to  humans  (reviewed  in  [225]).  The  identification  of  cAMP  in  learning  and  memory  is  another  discovery   first   made   in   Drosophila   [228].   Whole   families   of   novel   channels   in  vertebrates,   such   as   transient   receptor   potential   (TRP)   channels,   basic  understandings  of  synaptic  transmission  are  all  based  on  discoveries  first  made  in  fruit  flies  (reviewed  in  [225]).      In  2000  the  genome  of  Drosophila  melanogaster  was  published  in  a  special  edition  of  the  Science  magazine  and  today  we  know  that  77%  of  human  disease-­‐related  genes  have  a  homologue   in  Drosophila  and  complicated  molecular  mechanisms,   including  apoptosis,  autophagy,  and  unfolded  protein  response  are  widely  conserved  between  human  and  fruit  fly  [229,230,231,232,233,234].    The  Drosophila   field  working  on  diseases  that  are  coupled  to  protein  aggregation  is  still   very   young   but   has   already   shown   to   be   of   immense   value   (see   “Drosophila  models   for  protein  aggregation”).  One  of  many  success  stories   includes  research   in  flies   related   to   Parkinson’s   disease   that   has   provided   compelling   evidence   that  parkin   and   PINK1   are   components   of   a   pathway   that   is   involved   in   regulation   of  mitochondrial   remodelling   and   that   mitochondrial   dysfunction   is   a   cause   of  Parkinson’s  disease  (reviewed  in  [225]).  

Huge  genetic  toolbox:  Gal4/UAS  system  

The  model  organism  Drosophila  melanogaster  offers  a  vast  variety  of  genetical  tools  to   investigate   all   sorts   of   scientific   questions.   This   toolbox  was   expanded   in   1993  when  Andrea  Brand  and  Norbert  Perrimon  published   their  ground-­‐breaking  paper  introducing  the  Gal4/UAS  system  [235].  Until  then  it  was  only  possible  to  manipulate  gene   expression   in   Drosophila   by   either   driving   the   expression   by   a   heat   shock  promoter  or  tissue  specific  promoters.  The  heat  shock  promoter  approach  allows  for  inducible   expression,   but   also   carries   disadvantages   like   ubiquitous   ectopic  expression,   basal   expression   from   the   heat   shock   promoter,   and   the   fact   that   heat  shock  itself  can  induce  altered  phenotypes  [236,237,238].  The  use  of  tissue-­‐specific  promoters  allows  for  targeted  expression,  but  there  are  limitations  in  the  availability  of  cloned  and  characterized  promoters  that  can  be  used  for  this  purpose.  In  addition,  this  technique  does  not  allow  for  expression  of  genes  coding  for  toxic  products  [235].  All   these   obstacles   can   be   bypassed   with   the   bipartite   Gal4/UAS   system.   In   this  

Introduction    

   

  27  

system,   any   gene   of   interest   is   placed   downstream   of   an   upstream   activating  sequence   (UAS)   and   is   incorporated   into   the   fly   genome.   Flies  with   such   insertion  represent  the  transgenic  responder  line.  The  UAS  contains  five  tandem-­‐arranged  and  optimized   Gal4   binding   sites.   Gal4   is   a   Saccharomyces   cerevisiae   derived  transcription   factor,  not   found   in  Drosophila,   that’s  why  transgenic  responder   lines  maintain  the   inserted  gene  of   interest   in  a   transcriptionally  silent  state.  Mating  the  transgenic   responder   line   with   flies   expressing   Gal4   in   a   cell   or   tissue   specific  pattern   (Gal4   driver   line)   generates   a   progeny   that   contains   the   UAS   controlled  transgene  in  all  cells  but  expression  of  the  transgene  is  restricted  to  cells  that  contain  Gal4  (see  Figure  5)  [235].  Today,  there  are  thousands  of  lines  available  that  contain  Gal4   or   different   UAS   controlled   genes   [225,239,240,241].   Some   of   the   UAS  dependant  gene  insertions  contain  fluorescent  markers,  e.g.  GFP,  YFP,  mCherry,  and  can   be   used   to   display   cellular   compartments   or   proteins.   This   is   achieved   by  addition  of  a  localisation  signal,  e.g.  nuclear  location  sequence  (nls)  that  directs  the  expressed  protein  to  the  nucleus,  CD8  that  directs  the  protein  to  membranes,  or  by  tagging  a  protein  with  a  fluorophore,  e.g.  UAS-­‐mCherry-­‐atg8a.  During  the  last  5  years  several  groups   independently  generated  UAS  controlled,   transgenic  RNAi   lines   that  allow  conditional  gene  inactivation  of  any  Drosophila  gene  (reviewed  in  [242]).  It  is  also   possible   to   combine   several   UAS-­‐dependant   elements   in   the   same   fly   and  thereby  design  rather  advanced  experiments.    

Figure  5:   Schematic   illustration   of   the   bipartite   Gal4/UAS   system.  Transcription   of   a   UAS   transgene   responder   requires   the   presence   of   the   yeast  transcription   factor   Gal4.   Crossing   of   flies   with   a   tissue   specific   Gal4   expression  pattern  (Gal4  driver)  and  the  UAS  responder  line  generates  new  F1  flies.  Expression  of  the  UAS  responder  in  this  F1  generation  is  driven  by  tissue-­‐specific  Gal4  (modified  after  [240]).    

Introduction    

   

 28  

The  Gal4   expression   is   temperature   regulated   and   expression   levels   increase  with  raising  temperature  [240].  Other   modifiers   of   the   Gal4/UAS   system   include   the   temperature   sensitive  expression   of   the   Gal4   inhibitor   Gal80   and   systems   with   hormone-­‐induced   Gal4  expression.  The  Gal4/UAS  system  can  even  be  used  for  highly  sophisticated  “Mosaic  Analysis   with   a   Repressible   Cell   Marker”   (MARCM)   experiments,   where   directed  mosaic  clones  are  generated  [243].  

Drosophila  models  for  protein  aggregation  

Drosophila   melanogaster   has   become   a   popular   model   for   studies   of   protein  aggregation   and   for   diseases   associated   with   aggregated   proteins,   such   as  Alzheimer’s   disease   (Aβ   and   tau),   familial   amyloidotic   polyneuropathy  (transthyretin,  TTR),  Parkinson’s  disease  (α-­‐synuclein),  Huntington’s  disease  (poly-­‐glutamine  expanded  huntingtin),  and  prion  disease  (HaPrP).  Greve  et  al.  published  an  Alzheimer’s  disease  (AD)  model  in  which  Aβ  production  is  up-­‐regulated  by   simultaneously   overexpression  of   the  human  amyloid  β-­‐precursor  protein   APP,   human   β-­‐secretase,   and  Drosophila   presenilin   (dPsn).   The   latter   two  proteins  process  APP   into  Aβ-­‐peptide.  These   flies  produced  modest   levels  of  Aβ40,  Aβ42,   an   additional   Aβ-­‐peptide   (δ-­‐Aβ),   and   intracellular   fragments   of   APP.  Consequences   of   this   overexpression   were   age-­‐dependant   neurodegeneration,  formation   of   β-­‐amyloid   plaques   and   shortened   life   span.   The   neurodegeneration  phenotype  was  enhanced  in  flies  carrying  mutations  in  dPsn  that  are  associated  with  early-­‐onset   familial   AD   (EOFAD)   [244].   In   parallel,   three   groups   independently  generated   transgenic  Drosophila   strains   expressing   human   Aβ40,   Aβ42,   and   Aβ42  E22G   (arctic  mutation).   The   produced   peptides   all   contained   an   N-­‐terminal   signal  peptide   targeting   the   peptide   for   secretion   [245,246,247].   Only   the   expression   of  either   Aβ42   or   Aβ42   E22G   resulted   in   amyloid   deposition   in   the   fly   brain.   Aβ42  caused   short-­‐term   memory   defects,   and   at   later   stages,   locomotor   defects,   age-­‐dependant   neurodegeneration   and   premature   death.   The   arctic   mutation   in   Aβ42  strongly  enhanced  all   these  phenotypes.   Iijima  et  al.  reported  necrotic  cell  death  as  cause   for   neurodegeneration   in   their   flies   [245,247].   Expression   of   Aβ-­‐peptides   in  the  eye  caused  retinal  degeneration  [245,246].  These  models  set  the  stage  for  several  follow  up  studies  trying  to  understand  more  about  the  mechanisms  underlying  intra-­‐neuronal   Aβ42   accumulation   and   neurodegeneration   [248].   In   addition,   these  models   were   used   in   the   search   for   new   genetic   and   pharmacological   modifiers  preventing  the  phenotypes  caused  by  Aβ42.  Some  of  the  approaches  were  to  alter  β-­‐  and  γ-­‐secretase  activity,  enhance  Aβ-­‐degrading  enzymes  like  neprilysin  and  insulin-­‐degrading  enzyme  (IDE),  or  directly  interact  with  Aβ  aggregation,  by  feeding  Congo  red,  or  ligands  that  stabilize  Aβ   in  an  α-­‐helical  structure  [248].  The  addition  of  such  ligands  had  protective  effects  and  prolonged  survival,   increased   locomotor  activity,  

Introduction    

   

  29  

and   reduced   neurodegeneration   [249].   Computational   studies   of   predicted   Aβ42  aggregation  caused  by  different  mutations  were  tested  in  Drosophila  and  revealed  a  strong  relation  between  degree  of  neurotoxicity  and  the  propensity  of  Aβ42  to  form  protofibrils  [250].  In  a  genetic  screen  utilizing  1963  EP  transposon  insertions  that  allow  Gal4  to  up-­‐  or  down-­‐regulate  endogenous  Drosophila  genes,  Cao  et  al.  found  23  modifiers  involved  in   the   pathogenesis   of   AD.   Some   of   the   genes   were   involved   in   the   secretory  pathway,  cholesterol  homeostasis,  copper  transport,  and  ubiquitination/proteolysis  (ubiquitin-­‐conjugating  enzyme  E2Q  and  NEP  2).  They  also  had  some  indication  that  the  autophagy  gene  ATG1  is  crucial  in  reducing  Aβ42  toxicity.  When  transcription  or  chromatin   remodelling   was   impaired,   Aβ42   had   more   severe   pathological   effects  [251].   In  a  genome  wide  expression  analysis  and  complementary  genetic  screen  on  Aβ42   flies,   several   oxidative-­‐stress   related   genes   (e.g.   ferritin   and   catalase)   were  identified  to  be  potent  suppressors  of  Aβ42  and  Aβ42  E22G  toxicity  [252].  Ling  et  al.  demonstrated  an  involvement  of  autophagy  in  Aβ42  expressing  flies.  In  their  hands,  reduction  of  autophagy  decreased  toxicity  while   the  opposite  effect  was  seen  upon  autophagy   induction   and   they   suggested   an   autophagic-­‐lysosomal   injury   to   be  involved  in  Aβ42  toxicity  [253].  To  look  closer  at  the  role  for  tau  in  AD  were  fly  models  created  that  express  human  wild-­‐type   or   mutant   forms   of   tau   (R406W   and   V337M)   that   are   associated   with  inherited   frontotemporal   dementia,   and   Parkinsonism   linked   to   chromosome   17  (FTDP-­‐17)   [254,255].   The   expression   of   tau   in   neurons   reduced   lifespan,   caused  neurodegeneration  characterized  by  nuclear  fragmentation  and  vacuole  formation  in  neurons   of   the   cortex   and   neuropil   [255,256].   Expressed   in   the   eye,   tau   evokes   a  “rough”   phenotype   [254].   Although   tau   expression  mimics   the  phenotype   of  AD   in  several  ways,   the   formation  of  neurofibrillary   tangles   (NFT)  was  not   seen   in   these  transgenic  flies  [255].  Flies  expressing  wild-­‐type  TTR,  TTR  V30M,  and  TTR  L55P,  two  mutations  associated  with  FAP  have  been  created.  Expression  of  any  of  the  two  mutant  forms  resulted  in  shortened   lifespan,   neurodegeneration,   and   attenuation   of   locomotor   activity  [257,258].  Recently,  Pokrzywa  et  al.  showed  an  uptake  of  circulating  TTR  by  the  fat  body   and   subsequent   storage   of   TTR   as   quasi-­‐crystalline   nanospherules   as   a  mechanism  to  neutralize  toxic  species  of  TTR  [259].  Research   on   Parkinson’s   disease   has   made   tremendous   advancements   due   to  Drosophila  models  [260].  Here,  I  will  focus  on  the  impact  of  aggregating  α-­‐synuclein.  In  2000  Feany  and  Bender  presented  their  results  of  flies  with  ectopic  expression  of  wild-­‐type   α-­‐synuclein,   α-­‐synuclein   A30P   and   α-­‐synuclein   A53T.   This   ectopic  expression   led   to   loss   of   dopaminergic   neurons,   locomotor   defects   in   adult   flies,  intra-­‐neuronal   inclusions   and   retinal   degeneration   [261].   These   phenotypes   were  confirmed   two   years   later,   and   the   described   phenotypes   could   be   altered   upon  changes   of   chaperone   activity.   Increased   levels   of   Hsp70   suppressed   α-­‐synuclein  toxicity  whereas  the  opposite  was  the  case  if  chaperone  activity  was  reduced  [262].  

Introduction    

   

 30  

Polyglutamine  (poly-­‐Q)-­‐expanded  huntingtin  is  linked  to  Huntington’s  disease  and  a   fly   model   expressing   poly-­‐Q   huntingtin   was   characterised   by   neuronal  degeneration  and  nuclear  inclusions.  Age  of  onset  and  severity  of  neurodegeneration  correlated  with   increases   in   poly-­‐Q   expansion   length   [263].   Changes   in   genes   that  regulate  degradation  of  misfolded  proteins,  such  as  the  E3  ligase  CHIP  (C-­‐terminus  of  Hsc70   interacting   protein)   that   also   is   responsible   for   targeting   tau   for  ubiquitination  and  degradation,  and  genes  related  to  apoptosis,  and  other  signalling  pathways  have  proven  to  strongly  modify  the  degenerative  effects  of  polyQ  proteins.  An   effective   strategy   to   prevent   polyQ-­‐expansion   caused   toxicity   was   the  maintenance   of   acetylation   levels   by   reduction   of   histone   deacteylases   (HDAC)  (reviewed   in   [264]).   One   identified   HDAC,   HDAC6,   has   been   shown   to   serve   as  molecular   link   between   autophagy   and   the   ubiquitin-­‐proteasome   system   (UPS)  [265].  Also  several  other  factors,  like  phosphoinositide-­‐dependant  kinase-­‐1  (PDK1),  p70   ribosomal   S6   kinase   (S6K),   and   endosomal   sorting   complex   required   for  transport   (ESCRT)   complexes,   were   identified   to   modify   autophagy   and   thereby  reduce  polyQ  mediated  toxicity  [266,267].  Drosophila  as  model  system  to  study  prion  diseases  has  been  more  changeling  than  expected.   Expression   of   wild-­‐type   PrP   from   Syrian   hamster   and   different   mutant  forms  of  PrP,  failed  to  induce  neuropathology  in  flies  and  only  very  small  amounts  of  mutant  PrP  was  found  to  accumulate  in  fly  brains.  Flies  producing  mouse  PrP  P101L  had  signs  of  neurodegeneration,  but  contained  no  detergent-­‐insoluble  or  proteinase  K  (PK)-­‐resistant  PrP  conformers.  Finally,  the  crossing  of  a  Gal4  driver  line  optimized  for   high   expression   levels  with   a   UAS-­‐responder   line   transgenic   for  wild   type   PrP  from   Syrian   hamster   (HaPrP)   generated   a   fly   with   neurons   that   underwent  spongiform   vacuolation.   Guanidine-­‐resistance   of   conformationally   changed   HaPrP  and   immunoreactivity   with   the   conformation-­‐dependant   antibody   15B3   further  confirmed   the   validity   of   this   model.   Only   PK-­‐resistant   PrP   was   absent.   Further  studies  with  this  model  ascribed  Hsp70  a  protective  role  and  suggested  a  mechanism  of   how   cytosol-­‐derived   Hsc70   and   HaPrP,   which   is   secreted   via   the   secretory  pathway,   can   interact   with   each   other.   The   idea   is   that   in   older   flies   expressing  HaPrP,   Hsp70   can   colonize   the   lipid   raft,   the   proposed   site   for   PrP   conversion  (reviewed   in   [268]).   Indeed,   it   has   been   shown   that   Hsc70   can   move   across  membranous  structures  into  organelles  or  into  the  extra-­‐cellular  space  [269,270].      

Introduction    

   

  31  

 

Molecular  pathways  connected  to  protein  misfolding  

There  is  a  continuously  protein  turnover  in  a  cell  and  protein  biogenesis  is  an  error-­‐prone  process.  Accumulation  of  damaged  or  misfolded  proteins  can  disturb  cellular  homeostasis  and  provoke  ageing  and  cytotoxicity.   It   is  therefore  necessary  for  cells  to   detect   and   counteract   protein   misfolding.   This   has   led   to   the   evolution   of   a  compartment-­‐specific   quality   control   systems   that   can   initiate   cellular   responses  upon   imbalances   in   protein   folding   [271].   The   response   mechanisms   include  refolding  or  degradation  of  misfolded  proteins  or  the  capability  to  undergo  suicide  if  the  process  of  misfolding  is  too  excessive.  In  this  chapter  I  will  focus  on  the  quality  control   system   and   response   mechanisms   in   the   ER,   programmed   cell   death  (apoptosis),   and   autophagy,   a   mechanisms   shown   to   degrade   misfolded   protein  aggregates.    

ER-­‐stress  and  unfolded  protein  response  (UPR)  

Secretory  proteins  are  directly  released  into  the  ER  after  translation  and  roughly  one  quarter  of   the  proteome  that   traverses   the  secretory  pathway  structurally  matures  in   the  ER   [272].   In   order   to   cope  with   these  high  demands  has  ER  developed   as   a  highly   specialized   and   optimized   compartment   containing   a   high   concentration   of  general   chaperones   and   a   quality   control   system   that   senses   the   folding   state   of  present  proteins  [273].  An  imbalance  in  ER  load  of  unfolded  and  misfolded  proteins  and  the  capacity  to  cope  with  it  is  referred  to  as  ER-­‐stress  [274].  But  how  can  the  ER  sense   the   folding  state  of  a   residing  protein?  There  are  several   systems   for   folding  quality  control  in  the  ER.  Glycoproteins  become  Asparagine  (N)-­‐linked  glycosylated  in   the  ER.   Initially,  oligosaccharyltransferase   (OST)   transfers  glucose3-­‐mannose9-­‐N-­‐acetylglucosamine2   (Glc3Man9GlcNAc2)   to   an   acceptor   sequence   (N-­‐X-­‐S/T).   The  protein-­‐bound   oligosaccharide   is   processed   in   a   cascade   of   reactions   resulting   in  specific   N-­‐glycan   structures   that   direct   proteins   for   folding,   export   or   degradation  [275].   First,   two   glucoses   are   removed   by   glucosidase   I   and   II   generating   a  monoglucosylated   (Glc1Man9GlcNAc2)   glycoprotein,   which   is   recognized   by   the  lectins  calreticulin  and  calnexin.  Removal  of  the  remaining  glucose  by  glucosidase  II  results  in  the  release  of  the  glycoprotein  from  its  interacting  lectin.  Correctly  folded  glycoproteins  can  then  exit  the  ER.  However,  glycoproteins  with  an  incorrect  fold  are  re-­‐glucosylated   by   UDP-­‐glucose:glycoprotein   glucosyltransferase   (UGGT)   and   re-­‐enter   the   on-­‐off   cycle   with   lectins   until   the   glycoprotein   has   reached   its   native  conformation  or  is  targeted  for  degradation.  Glucosylation  plays  an  important  role  in  

Introduction    

   

 32  

quality  control,  as  the  added  glucose  acts  as  a  tag  for  incomplete  folded  proteins.  The  glycoprotein  can  only  exit  this   folding  cycle   if  UGGT  fails  to  re-­‐glucosylate  [273].  In  vitro   it   could   be   shown   that   UGGT   recognizes   partially   folded,  molten-­‐globule   like  conformations.  Exposure  of  hydrophobic  patches  that  usually  are  buried  in  the  core  of   a   folded   protein,   are   important   in   the   recognition   of   such   incomplete   folded  structures   by   UGGT   [276,277].   Removal   of   a   specific   mannose   residue   by   the   ER  α1,2-­‐mannosidase  I  (ERManI)  leads  to  the  recognition  of  the  glycoprotein  by  the  ER  degradation-­‐enhancing   1,2-­‐mannosidase   like   protein   (EDEM)   and   subsequent  targeting   for  ERAD   [273].   ERManI  has   a   relative   low   catalytic   activity   and   thereby  sets  a  time  window  for  productive  glycoprotein  folding  [271].  The  absence  of  an  “N-­‐X-­‐S/T”  acceptor  sequence  makes   it  unlikely   that   IAPP   folding   is  monitored  via   this  pathway.    The  ER  has  also  developed  an  oligosaccharide  independent  control  system  in  which  binding   of   the   immunoglobulin-­‐binding  protein   (BiP)   is   crucial.   BiP   belongs   to   the  Hsp70   chaperone   family   and   is   the   most   abundant   ER-­‐chaperone.   The   primary  function  of  BiP  is  to  assist  protein  folding  and  it  is  of  no  surprise  that  BiP  is  usually  the   first   chaperone   to  bind  a  nascent  polypeptide   chain  entering   the  ER.  However,  BiP  is  not  only  involved  in  ER  associated  folding  (ERAF)  but  has  also  been  suggested  to  target  substrates  for  ERAD.  Today  it  remains  elusive  how  this  distinction  is  done.  However,  some  preliminary  results  point  out  ER-­‐resident  J-­‐domains  (ERdj)  (the  ER  contains  seven  different  ERdjs)   to  be   involved   in  pathway  selection.  ERdjs  catalyse  hydrolysis   of   ATP   by   BiP,   an   important   step   in   the   on-­‐off   cycle   of   BiP   and   its  substrate   [278].   In   general,   it   is   speculated   that   extensive  BiP   binding   of   unfolded  substrates  is  involved  in  ER-­‐stress  sensing  (see  below).  Once  ER-­‐stress  is  sensed  and  induced,   cells   have   different   opportunities   to   counteract   this   imbalance.   Cells   can  reduce   the   protein   load   entering   the   ER   by   lowering   protein   synthesis   and/or  increase   ERs   capacity   to   handle   unfolded   proteins.   This   requires   targeted   up-­‐regulation  of  transcription  of  genes   involved  in  ERAF  and  ERAD,  e.g.  chaperones.   If  ER-­‐stress   is   too   intensive   and   cannot   be   counteracted   by   ERAF   and   ERAD   is   cell  death   triggered.   This   minimizes   the   damage   on   organism   level   due   to   extensive  production  of  misfolded  proteins.   Cellular   reactions   on  ER-­‐stress   are   embraced  by  the  term  unfolded  protein  response  (UPR).  There  is  no  hierarchical  order  in  how  ER-­‐stress   is   counteracted.   Instead   act   three   different   arms   of   UPR   in   parallel,  initiated  by   specific   signal   transducers,   inositol-­‐requiring  protein-­‐1   (IRE1),  protein  kinase   RNA   (PRK)-­‐like   ER   kinase   (PERK),   and   activating   transcription   factor-­‐6  (ATF6)  (see  Figure  6)  [274].                  

Introduction    

   

  33  

 Figure  6:   ER-­‐stress   sensors   and   pathways   of   the   UPR.   ER-­‐stress   activates  unfolded   protein   response   (UPR)   mediated   by   three   distinct   ER-­‐stress   sensors:  ATF6,  PERK  and  IRE1.  Cleaved  ATF6  relocates  to  the  nucleus,  binds  to  an  unfolded  protein   response   element   (UPRE)   and   activates   chaperone   transcription.   PERK  phosphorylates  elF2α,  resulting  in  attenuation  of  protein  synthesis  and  activation  of  UPRE-­‐binding   ATF4.   IRE1   activation   leads   to   transcription   of   ER   chaperones   and  ERAD  components  by  splicing  of  the  transcription  factor  Xbp-­‐1.  Extensive  ER-­‐stress  can  trigger  autophagy  and/or  apoptosis  (adopted  and  modified  from  [231]).  

Introduction    

   

 34  

IRE1   is   a   transmembrane   domain   that   resides   in   the   ER.   Activation   of   IRE1   and  subsequent   signalling   requires   its   oligomerisation   and   is   thought   to   be   a   direct  consequence  of  ER-­‐stress  [279].  IRE1  possesses  structural  similarity  to  the  peptide-­‐binding  domain  of  MHC  complexes,  suggesting  a  direct  interaction  of  IRE1  with  other  proteins   [280].   However,   the   crystal   structure   of   IRE1   shows   that   the   MHC-­‐like  binding  groove  is  obstructed  and  not  accessible  for  protein  interactions  [279].  This  finding   is   in   conflict   with   in   vitro   experiments   showing   direct   interaction   of   IRE1  with  unfolded  proteins  [281].  Hence,  it  is  unclear  if  IRE1  oligomerisation  is  a  direct  outcome   of   binding   to   unfolded   proteins.   Alternatively,   it   has   been   suggested   that  BiP   binds   to   IRE-­‐1   and   thereby   prevents   its   oligomerisation.   High   amounts   of  unfolded   proteins   could   then   withdraw   BiP   from   IRE-­‐1   leading   to   IRE-­‐1  oligomerisation.  However,  this  model  is  challenged  by  BiP  binding  not  to  be  essential  for   IRE1   regulation   [280,282].   Also   a   hybrid   model   has   been   suggested,   which  requires   both   -­‐   dissociation   of   BiP   from,   and   binding   of   unfolded   protein   to   IRE1  [274].  Once  IRE1  oligomerises  it  becomes  trans-­‐autophosphorylated  on  its  cytosolic  side  and  activated   [283].  The  only  substrate   for   IRE1   is  mRNA  of   the   transcription  factor   X-­‐box   binding   protein-­‐1   (Xbp1)   [284,285].   Hence,   IRE1   activity   can   be  monitored   by   detection   of   spliced   Xbp1.   Only   in   its   spliced   form   initiates   Xbp1  transcription   of   several   UPR   genes,   including   chaperones   that   increase   folding  capacity,   genes   for   lipid   synthesis   that   result   in   ER-­‐expansion,   and   ERAD   proteins  [274,286].   Phosphorylated   IRE1   can   also   recruit   Traf2   (tumour   necrosis   factor  receptor-­‐associated   factor-­‐2),   a   complex   that   has   been   linked   to   caspase-­‐12  activation  and  cell  death  [287,288].  The  cellular  response  to  stress  in  the  cytosol   is  termed  heat  shock  response  (HSR)  and  mediates  gene  expression  of  e.g.  chaperones  via  the  heat  shock  factor-­‐1  (hsf1).  It  has  been  shown  that  activation  of  hsf1  can  relief  ER-­‐stress   and   this   led   to   the   identification   of   several   genes   that   are   regulated   by  both,  hsf1  and  Xbp1.  These  results  point  towards  a  cross  talk  between  UPR  and  HSR  [289].  ER-­‐stress  can  also  result  in  oligomerisation  and  trans-­‐autophosphorylation  of  PERK.  This  step   is   followed  by  phosphorylation  of   the  α-­‐subunit  of  eukaryotic   translation  initiation  factor-­‐2  (elF2α),  which  results  in  lower  levels  of  active  elF2.  Low  levels  of  active   elF2  decrease   general   translation   and   thereby   reduce   the   load   from  protein  synthesis.  In  addition,  low  levels  of  active  elF2  also  lead  to  an  increased  translation  of  activating  transcription  factor  4  (ATF4).  ATF4  triggers  the  transcription  of  genes  involved  in  UPR,  e.g.  Xbp1  and  CHOP  (C/EBP  homologous  protein)  [231,274].  The  transmembrane  protein  ATF6  represents  the  third  arm  of  UPR.  Upon  ER-­‐stress  is  ATF  6  transported  to  the  Golgi  and  cleaved  by  Golgi  resident  S1  and  S2  proteases.  The   released   cytosolic  domain  of  ATF6   then   transfers   to   the  nucleus   and  activates  transcription   of   Xbp1,   calreticulin,   and   ER-­‐chaperones   such   as   BiP   [231,285].  Expression  of  the  unspliced  form  of  Xbp1  alerts  cells  for  further  ER-­‐stress  responses  [286].  The   three   different   arms   with   their   own   signalling   cascades   interplay   [274].   The  basic   principle   of   ERAD   comprises   the   recognition   and   unfolding   of   misfolded  

Introduction    

   

  35  

proteins  and  their   translocation  over   the  ER  membrane  to   the  cytosol  where   these  proteins   are   tagged   with   ubiquitin   and   are   targeted   for   proteasomal   degradation  [290].  One   important   aspect   of  ERAD  has   to  be  kept   in  mind  when  discussing   this  mechanism   in   the   context   of   amyloid   prone   proteins:   even   though   the   exact  mechanism  of  retro-­‐translocation  of  proteins  from  the  ER  to  cytosol  still  is  matter  of  debate,   it   is   unlikely   that   larger   aggregates   can   be   transported   this   way.   Indeed,  FRAP   (fluorescence   recovery  after  photobleaching)  experiments  with  a   fluorescent  ERAD   substrate   demonstrated   that   only   non-­‐aggregated   populations   were  eliminated  by  ERAD  [291].  Aggregated  proteins  can  be  isolated  in  the  ER-­‐associated  compartment  (ERAC)  -­‐  a  subcompartment  of  the  ER  [292].  These  subcompartments  were   found   to  be   cleared  by   autophagy  but   the  mechanistic   details   behind   remain  unclear   [274].   Bernales   et   al.   found   autophagic   removal   of   ER   as   response   to  UPR  and  termed  this  process  ER-­‐phagy  [293].  This  process  has  to  occur  in  the  absence  of  ubiquitination  since  the  ER  lacks  the  ubiquitin  system  [274].  Autophagy  induction  as  a  consequence  of  ER-­‐stress  occurs  via  both,  IRE1  activation  and  subsequent  JNK  (c-­‐Jun  N-­‐terminal  Kinase)  signalling,  and  elF2α  phosphorylation  by  PERK  [294].    If  none  of  the  above-­‐mentioned  actions  help  to  ease  ER-­‐stress  can  UPR  also  initiate  cell  death.  The  exact  link  between  ER-­‐stress  and  apoptosis  are  still  elusive,  but  some  key   signalling   events   have   been   identified.   It   is   known   that   the   ER-­‐stress   sensors  PERK   and   IRE1   are   involved   in   triggering   apoptosis.   PERK   can   activate   CHOP   (via  ATF4),  which  in  its  turn  down-­‐regulates  the  anti-­‐apoptotic  factor  B  cell  lymphoma-­‐2  (Bcl-­‐2)   and  up-­‐regulates  proapoptotic  Bim   (BH3-­‐only  member  of   the  Bcl-­‐2   family)  [231].   As   result   of   Bim   and   Bcl-­‐2  modulation,  mitochondria   release   cytochrome   c,  which   leads   to   caspase   3   activation   and   apoptosis   [295].  Drosophila  melanogaster  lacks   an   apparent   CHOP   homologue,   but   substrates   of   CHOP,   such   as   Bcl-­‐2   (in  Drosophila   DEBCL   and   BUFFY)   can   be   found   in   fruit   flies.   There   role   of   ER-­‐stress  induced  apoptosis  in  Drosophila  melanogaster  is  unclear  though  [231].  Activation  of  cell   death   by   IRE1   involves   complex-­‐formation   with   Traf2   and   ASK1   (apoptosis  signal   regulating   kinase-­‐1)   and   activation   of   JNK   signalling   [296].   This   signalling  pathway  is  independent  of  Xbp1  splicing  and  in  vitro  experiments  suggest  decreased  Xbp1-­‐splicing  activity  of  IRE1  at   late  stages  of  ER-­‐stress  when  apoptosis   is   induced  [297].   A   third   connection   between   ER-­‐stress   and   apoptosis   has   been   found   in  caspase-­‐12.   Calcium  activated   calpains   induce   caspase   12  dependant   procaspase-­‐9  cleavage.   Cleaved   caspase   9   in   its   turn   activates   caspase   3   and   thereby   triggers  apoptosis   [231].   Caspase-­‐12   deficient   mice   have   been   shown   to   have   decreased  levels  of  ER-­‐stress  induced  apoptosis,  but  showed  at  the  same  time  no  alterations  in  apoptosis  in  response  to  other  death  stimuli  [298,299].    The  intensity  of  UPR  activation  seems  to  be  critical  for  response  selection  upon  ER-­‐stress.  In  experiments  with  human  cell  lines,  pro-­‐apoptotic  mRNAs  were  less  stable  when   compared  with  mRNA  of   factors   that   facilitate   protein   folding   and   adaption.  This  difference  was  neutralized  under  conditions  of  strong  and  continuous  ER-­‐stress  when  pro-­‐apoptotic   factors  prevailed.  Hence  programmed  cell  death  was  activated  [300].  

Introduction    

   

 36  

The   roles   of   UPR   and   ER-­‐stress   related   responses   in   the   context   of   IAPP   are  controversial.  Expression  of  either  mIAPP  or  hIAPP  increased  the  amounts  of  BiP  in  10  weeks  old  mice.  Specific  for  hIAPP  expressing  mice  was  the  increased  occurrence  of   active   Xbp1   (as   detected   with   Western   Blot),   ubiquitin,   active   caspase-­‐12   and  CHOP,  suggesting  toxicity  that  might  be  due  to  ER-­‐stress  [301].  The  same  group  also  found  high  levels  of  CHOP  in  HiP  rats  and  again  this  was  interpreted  as  induction  of  apoptosis   due   to   ER-­‐stress   [302].   Marchetti   et   al.   found   only  modest   signs   of   ER-­‐stress   in   islets   from   individuals  with   type   2   diabetes.  However,   isolated   islet   from  such  patients   showed   signs   of   ER-­‐stress   in   form  of   increased  mRNA  expression   of  BiP  and  Xbp1  when  cultured  with  elevated  glucose  levels  (11.1  mM/l).  No  ER-­‐stress  was  detected  at  glucose  levels  of  5.5  mM/l  [303].  Accumulation  of  BiP,  spliced  Xbp1,  CHOP,   and   Bcl-­‐2   was   demonstrated   in   islets   from   human   pancreas   of   individuals  with  type  2  diabetes.  ER-­‐stress  was  also  detected  in  isolated  islets  from  db/db  mice  [304].   Taking   in   account   that   these   db/db   mice   are   not   transgenic   for   hIAPP,   one  might  ask  to  which  extent  the  observed  ER-­‐stress  in  diabetic  patients  really  is  due  to  IAPP   aggregation   and   not   to   other   abnormalities   found   in   patients   with   type   2  diabetes.   Treatment   with   exogenous   hIAPP   also   has   been   reported   to   induce   ER-­‐stress  [305].  Even  though  several  of  these  models  for  type  2  diabetes  concluded  that  ER-­‐stress   is   initiated,   the   question   remains   unanswered   if   UPR   is   triggered   by  aggregation  of  hIAPP.  A  study  performed  by  Hull  et  al.  failed  to  find  any  involvement  of  ER-­‐stress  and  questions  if  UPR  is  the  pathway  mediating  toxic  effects  of  hIAPP.  In  this  study  the  authors  looked  at  mRNA  levels  of  BiP,  CHOP,  ATF4,  and  spliced  Xbp1  in   islets   from   human   IAPP   transgenic   mice   cultured   at   different   glucose  concentrations.  As  control  they  used  islets  from  non-­‐transgenic  mice.  There  was  no  relative  increase  in  mRNA  levels  of  any  investigated  ER-­‐stress  marker  in  islets  from  hIAPP   transgenic   mice.   However,   ER-­‐stress   markers   did   increase   in   response   to  thapsigargin,   a   known   ER-­‐stress   inducer.   It   was   shown   by   immunohistochemistry  that  both  BiP  and  activated  Xbp1  were  elevated  in  pancreas  of  mice  fed  a  diet  high  in  fat  and  in  human  pancreas  from  patients  with  type  2  diabetes.  This  staining  was  not  influenced   by   the   presence   or   absence   of   islet   amyloid   though.   A   finding   that  prompted   the   authors   to   discuss   the   possibility   that   other   factors   than   IAPP  aggregation,  such  as  e.g.  elevated   levels  of  circulating  NEFAs,  have  to  be  accounted  for  ER-­‐stress  associated  with   type  2  diabetes   [306].   Immunohistochemical  analysis  of   mice   that   were   transgenic   for   hIAPP   but   lacked   the   gene   for   mIAPP  (+hIAPP/mIAPP-­‐)   and  were   fed  a  diet  high   in   fat   failed   to   find  an  up-­‐regulation  of  CHOP  due  to  hIAPP  expression.  The  antibodies  used  in  this  setup  were  the  same  as  in  the  studies  mentioned  above  (G.  Westermark,  unpublished  results).  These  mice  did  develop  amyloid  though  and  even  intra-­‐granular  amyloid  fibrils  were  found  [142].  A  recent   study   by   Gurlo   et   al.   did   not   directly   examine   ER-­‐stress   in   the   context   of  hIAPP  aggregation  but   rather   investigated  oligomer   formation  of  hIAPP   in   islets  of  transgenic   hIAPP  mice   and   pancreas   tissue   of   patients  with   insulinoma   (with   and  without  type  2  diabetes).  Sections  were  taken  from  tumour-­‐free  pancreas  providing  material   from   individuals   with   and   without   type   2   diabetes.   Oligomeric   hIAPP  

Introduction    

   

  37  

species  were  identified  in  secretory  granules  in  the  beta  cells   from  transgenic  mice  as  well  as  beta  cells  from  patients  with  type  2  diabetes  [307].  The  presence  of  such  oligomers  in  secretory  vesicles  after  the  ER  (only  minor  amounts  were  present  in  the  ER  and  Golgi)   further   indicates   that  hIAPP   can  pass   through   the  ER  without  being  degraded  by  ERAD  or  causing  UPR  in  such  degree  that  apoptosis  is  triggered.  

Apoptosis  

It  is  of  extreme  importance  for  metazoans  to  control  both  cell  survival  and  cell  death.  In   this   context,   apoptosis   (programed   cell   death)   has   a   central   role   during  development   and   homeostasis.   Failure   of   apoptosis   regulation   has   been   linked   to  several   diseases,   e.g.   cancer   (too   little   apoptosis)   and   neurodegenerative   diseases  (removal   of   neurons   due   to   apoptosis   induction).   Central   hallmarks   of   apoptosis  include  caspase-­‐activation,  DNA  fragmentation  (often  visualised  by  TUNEL  (terminal  deoxynucleotidyl   transferase   dUTP   nick   end   labelling)),   membrane   blebbing,   and  nuclear   and   cytoplasmic   condensation   (apoptotic   bodies).   There   are   two   different  pathways   that   can   lead   to   the   activation   of   programed   cell   death;   one   initiated   by  extracellular  factors  (extrinsic  pathway),  and  the  intrinsic  pathway  that  is  triggered  upon  intracellular  death  stimuli  (reviewed  in  [229,308]).    The  extrinsic  pathway  is  activated  via  a  cell  surface  bound  death  receptor,  such  as  the   tumor  necrosis   factor  receptor  1  (TNFR1)  or   the  Fas  receptor.  Death  receptors  oligomerise   upon   binding   of   extracellular   death   ligands,   e.g.   TNFα   or   TNFβ.   Death  receptor   oligomerisation   leads   to   the   binding   of   the   adaptor   protein   FADD   (Fas-­‐associated  death  domain)  forming  a  death-­‐inducing  signalling  complex  (DISC)  at  the  plasma   membrane.   FADD   further   recruits   procaspase-­‐8   or   -­‐10,   which   are   auto-­‐activated  upon   this   relocation.  Once   activated,   the   initiator   caspase  will   cleave   the  activator   caspase-­‐3   or   -­‐7   and   thereby   set   the   stage   for  mobilization   of   DNase   and  induction  of  DNA  fragmentation.  Under  normal  conditions  the  DNase  DFF40/CAD  is  tightly   bound   by   DFF45/ICAD   and   thereby   inhibited.   Once   caspase-­‐3   or-­‐7   are  cleaved  and  activated  they  will  degrade  DFF45/ICAD  and  liberate  DNase  DFF40/CAD  (reviewed  in  [229]).  One  central  event  of  the  intrinsic  pathway  is  the  modulation  of  members  of  the  Bcl-­‐2  protein  family.  Bcl-­‐2  proteins  can  be  divided  into  three  classes;  one  class  with  anti-­‐apoptotic  function  (Bcl-­‐2  (the  protein  giving  name  to  the  whole  family)  and  Bcl-­‐xL)  and   two   classes   comprising   pro-­‐apoptotic   proteins.   These   two   subfamilies   are  classified  as  multi-­‐domain  subfamily  (e.g.  Bax  and  Bak)  and  BH-­‐3  only  subfamily  (e.g.  Bim,  Bid,  Bik).   The   anti-­‐apoptotic   effect   of  Bcl-­‐2  or  Bcl-­‐xL   can  be   ascribed   to   their  interaction  with   the   pro-­‐apoptotic  members   of   their   family.   Interaction   occurs   via  binding   to   the   BH3   domain.   Apoptotic   death   signals   (e.g.   due   to   growth   factor  withdrawal  or  ER-­‐stress)  are  associated  with  inhibition  of  anti-­‐apoptotic  members  of  the  Bcl-­‐2   family  and/or  activation  of  pro-­‐apoptotic  members  of   this  protein   family.  

Introduction    

   

 38  

As   result   of   this   complex   interplay   homo-­‐oligomerise   apoptosis-­‐inducing   Bcl-­‐2  proteins   and   form   pores   in   mitochondria   leading   to   leakage   of   cytochrome   c   and  Smac/DIABLO  to  the  cytosol  [229,309,310].  Cytochrome  c   forms  a  complex  termed  apoptosome   together   with   the   adaptor   protein   Apaf-­‐1.   [311].   The   apoptosome  cleaves  caspase  9,  which  in  its  turn  activates  the  effector  caspase-­‐3  and  -­‐7  and  finally  leads  to  the  dismantling  of  the  cell.  All   caspases   are   inactive   unless   they   are   cleaved   proteolytically.   Initiator   caspases  (caspase-­‐8,   -­‐10,   -­‐9,   and   -­‐2)   are   auto-­‐activated   under   apoptotic   conditions  whereas  effector   caspases   (caspase-­‐3,   -­‐7,   and   -­‐6)   have   to   be   cleaved   by   initiator   caspases.  There  is  a  second  mechanism  of  protection  against  caspase  activity.  The  inhibitor  of  apoptosis   (IAP)   family   negatively   regulates   caspases.   In   order   to   trigger   apoptosis  these  IAPs  have  to  be  inhibited.  Smac/DIABOLO  (released  together  with  cytochrome  c  from  mitochondria)  interacts  and  neutralizes  IAPs  (reviewed  in  [229]).  The   extrinsic   and   intrinsic   pathways   share   some   common   players.   Bid,   a   pro-­‐apoptotic  Bcl-­‐2  family  member  is  activated  upon  DISC  formation  (an  early  step  in  the  extrinsic   pathway)   and   induces   mitochondrial   release   of   pro-­‐apoptotic   factors  [312,313].   The   activation   of   caspase-­‐3   is   also   a   common   feature   of   both   signalling  pathways  (see  Figure  7).      

Introduction    

   

  39  

 

Figure  7:   Apoptotic   pathways   in   mammalian   cells.   Hallmarks   of   the  extracellular   triggered  extrinsic  pathway   include  the   formation  of  a  death-­‐inducing  signalling   complex   (DISC),   activation   of   effector   caspases-­‐3,7   and   DNA  fragmentation.  The  intrinsic  pathway  is  activated  by  intracellular  apoptosis  signals,  leads   to   mitochondrial   release   of   cytochrome   c,   apoptosome   formation,   and  activation   of   initiator   caspase-­‐9   and   effector   caspases-­‐3,7,   subsequent   resulting   in  DNA  fragmentation  (adopted  and  modified  from  [229]).  

 Drosophila   melanogaster   contains   homologues   to   many   of   the   above-­‐mentioned  mammalian   proteins.   And   even   though   certain   protein-­‐protein   interactions  responsible   for   caspase   regulation  are   controlled  differently   in   flies   and  mammals,  

Introduction    

   

 40  

there  still  is  a  big  degree  of  conservation  when  it  comes  to  the  general  control  of  the  pathways.   One   example   of   different   regulations   includes   inhibition   of   caspases   by  IAPs.   In  mammals   is   caspase-­‐9   activity   directly   inhibited   by   XIAP.   The   Drosophila  homologue  DIAP1  shows  no  such  direct  inhibition  of  Dronc  (the  Drosophila  caspase-­‐9  homologue)  but  functions  as  E3  ligase  recognizing  Dronc  and  thereby  targeting  it  for   proteasomal   degradation   [229,314].   So   even   if   the   inhibitory   effect   of   IAPs   is  different,  the  basic  principle  is  the  same:  inactivation  of  caspases  by  IAPs,  which  only  is  countermanded  by  an  apoptotic  signal  and  subsequent  Smac/DIABLO  (Drosophila  homologues:   hid,   grim,   reaper,   and   sickle)   release   from   the  mitochondria.   Drice   is  the  Drosophila  homologue  of  mammalian  caspase-­‐3,  another   important  molecule   in  apoptosis.  There   is   evidence   that   beta   cells   reduction   in   type   2   diabetic   patients   is   due   to  increased  apoptosis  [152,315].  Several  studies  have  suggested  an  apoptotic  beta  cell  death   as   direct   consequence   of   islet   amyloid   formation.   IAPP   has   been   shown   to  induce   caspase-­‐3   activation   and   JNK  phosphorylation   in   a   time-­‐   and   concentration  dependant  manner  when   added   to   RINm5F   cells   [316].   Addition   of   amyloidogenic  IAPP  to  pancreatic  beta  cell   lines  did  trigger  apoptosis  via  caspase-­‐8  and  caspase-­‐3  activation,  whereas   (25,28,29triprolyl)-­‐IAPP   (not   capable  of   forming  amyloid)  did  not  have   such  effect   [58].  The   same  group  detected  an  up-­‐regulation  of   Fas-­‐associated  death  receptor  in  beta  cells  as  response  to  hIAPP  exposure  [57].  Zraika  et  al.   found  an  induction  of  oxidative  stress  and  following  production  of  reactive  oxygen  species  (ROS)   as   response   to   IAPP   presence   in   culturing   medium.   In   their   hands,   this  increase  in  oxidative  stress  did  not  mediate  beta  cell  apoptosis  in  the  short  term.  The  authors   speculated   that   ROS   did   accelerate   amyloid   formation   via   an   unidentified  feedback   mechanism   and   thereby   contribute   to   beta   cell   apoptosis   [317].   The  involvement  of  ROS  in  mediating  toxicity  is  of  interest  in  regard  to  a  study  recently  published   by   Li   et   al..   They   detected   mitochondrial   dysfunction   accompanied   by  cytochrome  c  release,  modulation  of  expression  levels  of  Bcl-­‐2  family  members,  and  caspase   activation   in   Ins-­‐1   cells   that   were   subjected   to   freshly   dissolved   hIAPP.  These  effects  were  not  seen  upon  mIAPP  exposure  [318].  Sensitivity  to  IAPP  amyloid  formation  depends  on  the  cell  type  though.  While  beta  cells  are  evidently  vulnerable  to  islet  amyloid  and  undergo  apoptosis,  were  alpha-­‐cells  much  more  resistant  to  islet  amyloid  induced  apoptosis  in  a  study  conducted  in  the  lab  of  L.  Marzban  [319].  Also  in  vivo   studies  have  shown   increased  beta  cell  apoptosis,  both   in   transgenic  mouse  models  and  HIP  rats   [142,320].  Experiments  conducted   in  baboons  even  showed  a  direct  correlation  between  pancreatic  islet  amyloidosis  severity  and  apoptosis  [160].  Taken   together,   islet   amyloid   formation   and/or   presence   are   likely   triggers   of   cell  death.  The  performed  studies  often  show  induction  of  apoptosis  upon  extracellular  hIAPP  presence.  An  intracellular  event  leading  to  apoptosis  could  be  the  activation  of  CHOP   upon   ER-­‐stress,   however,   results   showing   such   CHOP   involvement   are  contradictive   (see   ER-­‐stress   and   unfolded   protein   response   (UPR)).   It   therefore  remains  elusive  if  apoptosis  even  can  be  triggered  by  hIAPP  while  the  protein  still  is  intracellular.  

Introduction    

   

  41  

Autophagy  

The   term  autophagy,  which   literally  means   “self-­‐eating”   (coined  by  Nobel  Laureate  Christian  de  Duve  in  1963),  comprises  pathways  that  allow  cells   to  digest  cytosolic  components   via   lysosomal   degradation.   Proteasome   and   autophagy-­‐mediated  degradation  are  the  main  cellular  pathways  for  protein  and  organelle  turnover  and  all  eukaryotic  cells  can  employ  autophagy  [321,322].  Today,  three  different  classes  of  autophagy   are   distinguished:   microautophagy,   chaperone-­‐mediated   autophagy  (CMA),  and  macroautophagy.  Microautophagy   is   mainly   studied   in   yeast   (containing   vacuoles   instead   of  lysosomes)  and  consists  of  direct  invagination  of  the  vacuolar  boundary  membrane  and  budding  of  autophagic  bodies  into  the  vacuolar  lumen  [323].  Very  little  is  known  about   molecular   mechanisms   underlying   microautophagy   in   eukaryotic   cells.  However,   a   publication   from   Ana   Maria   Cuervos   lab   recently   described   a  microautophagy-­‐like   process   (named   endosomal   microautophagy,   e-­‐MI)   in  mammalian  cells  where  soluble  cytosolic  proteins  selectively  were  taken  up  by  late  endosomes/multivesicular   bodies   (MVBs).   Cargo   selection   was   dependent   on   the  chaperone   Hsc70   and   electrostatic   interactions   with   the   endosomal   membrane  [324].  In   CMA   is   cytosolic   cargo   selectively   recognized   by   a   complex   of   molecular  chaperones,   including   Hsc70,   bound   by   the   lysosome-­‐associated   membrane   type  protein  2A  (LAMP-­‐2A)  and  taken  up  by  the  lysosome  [325].  Some  of  the  differences  between  CMA  and  e-­‐MI  are  the  binding  to  LAMP-­‐2A  and  the  requirement  of  protein  unfolding  in  CMA  [324].  The  third  common  type  of  autophagy,  macroautophagy   (henceforth  referred  to  as  autophagy),   is   initiated  by  the   formation  of  a  cytosolic  double  membrane  structure  called   the  phagophore   (also   called   isolation  membrane).   Cytosolic   components   are  both   selectively   and   non-­‐selectively   engulfed   during   growth   of   phagophores.  Autophagosomes  originate   from   the   closure  of  phagophores   and  will   subsequently  fuse   with   lysosomes   and   thereby   enable   degradation   of   captured   cytosolic  constituents  [294,326,327].  It  is  matter  of  intensive  research  on  how  the  membrane  of   phagophores   initially   is   formed.   Results   from   several   independent   experiments  gave   rise   to   different   models   in   which   the   ER,   Golgi,   or   the   outer   membrane   of  mitochondria   respectively   is   source   of   the   initial   phagophore   double   membrane  [327,328].  Autophagy  has  early  been  identified  as  a  cellular  response  mechanism  to  starvation,  where  resulting  macromolecules  from  cytosolic  bulk  degradation  can  be  recycled  back   to   the  cytosol  and  be  reused.  But  autophagy  has  also  been  shown  to  turn   over   substrates   in   a   selective   manner   in   yeast,   a   pathway   known   as   the  cytoplasm-­‐to-­‐vacuole   targeting   (CVT)   pathway   [329].   In   analogy,   cargo   selective  degradation   of   aggregated   proteins   (aggrephagy   [330]),   mitochondria   (mitophagy  [331]),   ribosomes  (ribophagy  [332]),  peroxisomes  (pexophagy  [333]),  endoplasmic  reticulum  (reticulophagy  [293])  and  many  more  have  been  reported  for  mammalian  systems  [334].  The  role  of  aggrephagy  will  be  further  addressed  later  in  this  section.  

Introduction    

   

 42  

I   first   want   to   shed   some   light   on   several,   selected   key   events   of   autophagy   (see  Figure  8).    

Figure  8:   Summary   of  macroautophagy.  The  ULK  complex  and  PI3K  complex   I  are  both  required   for  nucleation  and   initial  expansion  of   the  phagophore.  Once  the  phagophore   is   formed   are   the   ubiquitin-­‐like   proteins   Atg12   and   LC3   (mammalian  homolog  of  Atg8)  with  their  respective  conjugation  systems  recruited  and  activated.  Further   membrane   expansion   and   maturation   leads   to   vesicle   closure   and  autophagosome   formation.   This   vesicle   can   fuse   with   different   endocytic  compartments   or   directly   with   lysosomes   forming   autolysosomes.   Phagophore-­‐sequestered   material   can   thereby   be   degraded.   Released   macromolecules   can   be  recycled  back  to  the  cytosol  (adopted  and  modified  from  [294,335]).  

Introduction    

   

  43  

Genetic  screens  in  S.  cerevisiae  have  led  to  the  identification  of  numerous  autophagy-­‐related  (ATG)  genes  and  many  homologs  have  been   identified  and  characterized   in  higher   eukaryotes   [294].   In   general,   autophagy   can  be  divided   into   three   steps:   1)  induction/nucleation;  2)  expansion;  and  3)  maturation  [335].    Basal   levels   of   autophagy   are   low   under   normal   conditions.   However,   intra-­‐   and  extracellular   stress   factors,   such   as   starvation,   ER-­‐stress,   hypoxia   and   pathogen  invasion   can   induce   autophagy   [294].   One   important   regulator   of   autophagy   is  mammalian   Target   of   Rapamycin   (mTOR),   a   serine/threonine   protein   kinase   that  inhibits   autophagy.   Amino   acid   starvation   negatively   regulates  mTOR   and   thereby  induces  autophagy  [336,337].  Substrates  of  mTOR  include  the  Unc-­‐51-­‐like  kinase  1  (ULK1)  and  -­‐2  (ULK2)  (mammalian  homologs  of  yeast  Atg1),  which  form  a  complex  with   mammalian   Atg13   (mATG13)   and   FIP200   (the   focal   adhesion   kinase   family-­‐interacting   protein   of   200   kDa).   Dephosphorylation   of   this   complex   (requiring  inactivation   of   mTOR)   induces   autophagy   [338].   As   mentioned   above,   much   is  unknown  about  the  site  and  exact  mechanisms  underlying  initial  phagophore  double  membrane   formation.   However,   the   class   III   phosphatidylinositol   3-­‐kinase   (PI3K)  complex   I,   consisting  of   the   class   III  PI3K  Vps34   (vacuolar  protein   sorting  34),   the  Vps15-­‐like  serine/threonine  kinase  p150,  Beclin1  (Atg6  in  yeast),  and  Atg14L  (also  named   Barkor)   and   its   generation   of   phosphatidylinositol   (3,4,5)-­‐trisphosphate  (PI3P)  has  been  shown  to  be  essential   in   the   initial   step  of  phagophore  membrane  formation.  Bcl-­‐2,  which  already  was  mentioned  in  the  context  of  UPR  and  ER-­‐stress,  inhibits   autophagy   by   binding   and   sequestering   Beclin1   under   nutrient-­‐rich  conditions.   Atg14L   has   been   assigned   a   pivotal   role   for   this   PI3K   complex   I   once  autophagy  is  activated  as  it  recruits  the  complex  to  the  site  of  phagophore  formation,  stimulates   it’s   activity   and   even   interacts  with   proteins,   such   as   LC3   (microtubule  associated  protein  1   light  chain  3,   the  mammalian  homolog  of  yeast  atg8),   that  are  crucial   in   later   steps   of   autophagy.   Elongation   of   the   phagophore   relies   on   the  ubiquitin-­‐like   proteins   Atg12   and   Atg8/LC3   and   their   respective   conjugation  systems.   Following   the   processing   by   a   cysteine   protease   (Atg4)   is   Atg8/LC3  covalently   linked  to  phosphatidylethanolamine  (PE)  through  the  action  of  Atg7  (E1  activating  enzyme),  Atg3  (E2  activating  enzyme)  and  the  E3  like  Atg5-­‐Atg12  complex  The   Atg5-­‐Atg12   complex   is   coupled   to   the   phagophore   via   Atg16L.   Lipidation   of  Atg8/LC3   is   a   central   step   in   autophagy.   Once   activated   and   lipid-­‐conjugated   is  Atg8/LC3   localized   to   both   sides   of   the   phagophore.   Atg4   removes   only   Atg8/LC3  residing   on   the   cytosolic   side   of   the   autophagosome   prior   to   autophagosome-­‐lysosome/endosome   fusion.   It   even  has  been   reported   that  Atg8   can  modulate   the  size  of  autophagosomes  by  influencing  membrane  curvature.  Taken  together,  it  is  of  no   surprise   that   activation   of   Atg8/LC3   is   widely   used   to   monitor   autophagy  [294,339].   It   has   been   proposed   that   stepwise   fusion   of   autophagosomes   with  different  endosomal  populations  account  for  maturation  and  culminates  in  the  fusion  with   lysosomes,   the  organelle   responsible   for  degradation  –  a  model   supported  by  several  findings.  Coat  protein  complex  I  (COPI),  that  is  involved  in  ER-­‐Golgi  transport  as  well  as  in  the  maintenance  of  endosomal/lysosomal  function,  has  been  identified  

Introduction    

   

 44  

to   be   necessary   for   autophagy.   Down-­‐regulation   of   COPI   coatomer   subunits   with  siRNA  in  GFP-­‐LC3  transfected  cells  resulted  in  the  accumulation  of  GFP-­‐LC3-­‐positive  autophagosomes.   These   autophagosomes   were   not   fused   with   lysosomes.   Loss   of  COPI  was  also  accompanied  by  an  accumulation  of  p62  and  ubiquitinated  proteins  [340].   Impairment   of   ESCRT  machinery  has   similar   effects   and   results   in   defective  autophagy  [341].  Other  known  protein  complexes  of  the  endocytic  pathway  that  also  have  been  demonstrated  to  be  involved  in  autophagosomes  maturation  include  class  III  PI3K  complex  II  (Vps34  together  with  p150  and  Beclin1)  with  UVRAG  as  positive  and  Rubicon  (RUN  domain  and  cysteine-­‐rich  domain  containing,  Beclin1-­‐interacting  protein)  as  negative   regulator,   and   the  Vps-­‐C  complexes  HOPS   (homotypic  vacuole  fusion  and  protein  sorting)  and  CORVET  (class  X  core  vacuole/endosome  tethering)  [342,343,344].   With   the   discovery   of   the   PI3P-­‐binding   protein   FYCO1   (FYVE   and  coiled-­‐coil   domain-­‐containing   protein   1)   binding   both   LC3   and   Rab7   (a   late  endosomal   marker)   there   even   is   a   molecular   link   between   autophagy   and   the  endocytic   machinery.   FYCO1  mediates   microtubule   plus   end-­‐directed   transport   of  autophagosomes   [345].   It   has   been   shown   before   that   the   transport   of  autophagosomes  along  microtubules  depends  on  dynein  [346].    Selective  autophagy  in  form  of  CVT  has  been  known  in  yeast  for  a  long  time  and  has  gained   major   attention   in   mammalian   systems   over   the   last   years.   Selectivity  requires   crucial,   additional   steps   to   the   above  described   autophagy  process:   cargo  has  to  be  recognized  by  a  specific  receptor  and  must  be  delivered  to  the  autophagic  machinery.  Here,  I  want  to  focus  on  selective  degradation  of  aggregated  proteins  via  the   autophagy   pathway   (aggrephagy)   and   introduce   essential   players   in   tagging  aggregating   proteins,   recognition   of   cargo   and   finally   feeding   aggregated   proteins  destined   for  degradation   into   the  autophagy  pathway   (see  Figure  9).  Ubiquitin  has  been   proven   to   be   crucial   when   it   comes   to   tag   proteins   that   are   determined   for  degradation  but  lately  also  in  endocytosis,  signal  transduction  and  DNA  repair  [347].  Conjugation   of   ubiquitin   depends   on   a   complex   reaction   cascade   requiring   the  enzymes   E1   (activating   ubiquitin),   E2   (ubiquitin   conjugating   enzyme),   and   E3  (ubiquitin   ligase).   As   result   ubiquitin   is   covalently   bound   via   an   isopeptide   bond  between   the   C-­‐terminal   glycine   of   ubiquitin   and   the   ε-­‐amino   group   of   a   lysine  residue   of   the   substrate   protein.   The   E3   ubiquitin   ligase   provides   substrate  specificity   by   recognizing   its   protein   substrate   and   bringing   it   to   the   E2   ubiquitin  conjugating  enzyme.  Cells  contain  several  of  E1,  E2,  and  E3  enzymes  providing  cells  with  a  great  tool  for  primary  selectivity  for  this  signalling  machinery  [348].  Ubiquitin  itself   contains   seven   lysine   residues   and   ubiquitin   attached   to   itself   forming   a  polyubiquitin   tag.   The   best-­‐characterised   linkages   occur   via   Lys48,   targeting   the  substrate   for   proteasomal   degradation,   and   via   Lys63,   which   is   preferred   by  ubiquitin-­‐binding  autophagy  receptors.  Furthermore,  Lys63  ubiquitination  has  been  reported   to   be   a   potent   enhancer   of   inclusion   formation   and   lead   to   substrate  degradation   via   the   autophagy/lysosome   degradation   pathway   [347,349,350,351].  Also  more   atypical   sites,   such   as   Lys6   or   Lys29,   for   polyubiquitination   have   been  reported  but  the  exact  role  of  these  ubiquitin  chains  is  still  poorly  understood  [352].    

Introduction    

   

  45  

 

Figure  9:   Schematic  drawing  of  aggrephagy.  Aggregate-­‐prone  proteins  become  ubiquitinated  and  are  then  recognized  by  the  adaptor  protein  p62,  which  in  its  turn  has   the   capacity   to   further   drive   aggregation.   This   inclusion   attracts   Alfy,   which  recruits  Atg5-­‐Atg12  and  in  addition  can  bind  PI3P  in  the  phagophore  membrane.  LC3  can  directly  interact  with  the  whole  complex.  The  spatial  proximity  of  LC3  and  Atg5-­‐Atg12  (binding  phagophore  bound  Atg16L)  allows  for  PE-­‐conjugation  of  LC3  by  the  E3-­‐like  ligase  properties  of  Atg5-­‐Atg12-­‐Atg16L  and  following  incorporation  of  LC3-­‐PE  to  the  phagophore  membrane.  In  this  way  a  autophagosome  can  be  formed  that  tightly   engulfs   the   protein   aggregate   and   is   devoid   of   other   cytosolic   components  (adopted  and  modified  from  [335]).  

Taken  together,  ubiquitin  conjugation  offers  several  possibilities  to  flag  proteins  and  organelles   in   different   ways   by   variation   of   chain   length   and   various   sites   for  ubiquitin  self-­‐attachment  and  thereby  act  as  a  signal  for  distinct  subsequent  cellular  

Introduction    

   

 46  

processes.   Molecular   links   between   ubiquitinated   proteins   and   autophagy   were  identified   in   form  of  the  sequestosome  marker  SQSTM1/p62  and  NBR1  (neighbour  of  BRCA1  gene).  The  conserved   functional  homolog   for  p62/NBR1   in  Drosophila   is  Ref(2)p.   Both,   p62   and   NBR1   contain   an   ubiquitin-­‐associated   (UBA)   domain,  allowing  for  interactions  with  ubiquitin,  a  LIR  (LC3  interacting  region)  domain,  and  a  N-­‐terminal   PB1   (Phox   and   Bem1p)   domain   that   facilitates   self-­‐   and   hetero-­‐oligomerisation   with   other   proteins   [347,353,354].   p62   is   necessary   for   the  formation   of   protein   aggregates   that   are   degraded   by   autophagy   [355,356,357].  Experiments  in  mice  further  proved  the  active  role  of  p62  in  protein  aggregation  and  that   these   protein   aggregates   are   targeted   for   autophagic   degradation.   Tissue  specific  autophagy  deficiency  (Atg7  knockout  mice)  led  to  p62  and  ubiquitin  positive  inclusions   in   neurons   and   hepatocytes.   However,   formation   of   ubiquitin-­‐positive  inclusions   was   suppressed   in   p62   knockout   mice,   underscoring   the   aggregate-­‐promoting  role  of  p62  [358].  Due  to  its  active  involvement  in  autophagy  serves  p62  as  useful  marker  for  autophagic  turnover  [359].  Several  diseases  are  associated  with  p62   containing   ubiquitinated   protein   inclusions,   such   as   Lewy   bodies   in   PD,  neurofibrillary   tangles   (AD),   Huntingtin   aggregates   (HD)   and   Mallory   bodies  (alcoholic   and   non-­‐alcoholic   steatohepatitis).   It   remains   to   be   clarified   if   these  inclusions  are  a  consequence  of  defective  autophagy  [330,360,361].  Very  recently,  Filimonenko  et  al.  were  able  to  identify  Alfy  (PI3P-­‐binding  Autophagy-­‐linked   FYVE   domain   protein)   to   be   actively   involved   in   autophagic   degradation   of  polyglutamine   expanded,   aggregated   proteins   [362].   The   400   kDa   protein   Alfy  usually   resides   in   the   nucleus   decorating   the   nuclear  membrane.   The   presence   of  ubiquitinated,   aggregated   proteins   in   the   cytosol   leads   to   relocalization   of   Alfy   to  these  aggregates  [363].  Alfy  can  directly  interact  with  p62  and  with  Atg5  [357,362].  In  vitro,  Alfy  is  necessary  to  recruit  atg5  to  polyQ  protein  aggregates.  In  addition,  Alfy  scaffolds   the   Atg5-­‐Atg12-­‐Atg16L   complex   to   p62-­‐   and   ubiquitin-­‐positive   polyQ  inclusions  [362].  The  Atg5-­‐Atg12-­‐Atg16L  complex  on  the  other  hand  is  important  for  LC3  lipidation  [364].  Taken  together,  all  these  interactions  allow  for  LC3  lipidation  in  close  spatial  proximity  to  ubiquitinated,  aggregated  proteins  and  explain  the  absence  of   other   cytosolic   components   in   aggregate   filled   autophagosomes   [362].   Primary  neurons  expressing  polyQ  Htt   (Huntingtin)  had   fewer  poly   inclusions  upon  ectopic  Alfy  expression.  These  results  were  confirmed  in  vivo  with  a  Drosophila  model  where  polyQ  production  provokes  a  phenotype  that  is  due  to  toxicity.  The  outcome  of  this  polyQ-­‐mediated   toxicity   was  much  milder   once   bchs   (blue   cheese,   the   Drosophila  homologue  of  Alfy)  was  co-­‐expressed   [362].  Reduced   levels  of  bchs   in  mutant   flies  had  opposite  effects  and  led  to  shortened  live  span  and  extensive  neurodegeneration  [365].  It  remains  to  be  elucidated  if  Alfy  directly  recognizes  ubiquitinated  aggregates  or  if  this  interaction  is  mediated  by  p62  [330].  Polyubiquitination   and   p62   are   not   only   required   for   selective   autophagy   of  aggregated  proteins.  Aged,  damaged  mitochondria  have  to  be  turned  over  and  the  E3  ubiquitin  ligase  Parkin  has  been  ascribed  an  important  function  in  mitophagy.  Parkin  induces   Lys63-­‐linked   polyubiquitination   of   mitochondrial   substrates   and   thereby  

Introduction    

   

  47  

recruits   p62,   which   in   its   turn   mediates   mitochondrial   clustering   [366].   There   is  contradicting  data  concerning  whether  mitophagy  depends  on  p62  or  not.  Narendra  et  al.  observed  no  decline  in  Parkin  mediated  mitophagy  in  MEFs  (mouse  embryonic  fibroblasts)  or  HeLa  cells  upon  p62  knockdown  or  siRNA  depletion  of  p62  whereas  Geisler   et   al   saw   a   complete   blockage   of   final   clearance   of   damaged  mitochondria  once  HeLa  cells  were  treated  with  siRNA  against  p62  [366,367].  There  also  has  been  a   report   of   p62   and   ubiquitin   dependent   autophagy   of   peroxisomes   (pexophagy)  [368].      Over  the  years  has  autophagy  been  implicated  in  many  neurodegenerative  diseases,  such   as   Huntington’s   disease   (HD),   Parkinson’s   disease   (PD),   Alzheimer’s   disease  (AD),   or   amyotrophic   lateral   sclerosis   (ALS).   All   four   are   also   associated   with  accumulation  of  protein  aggregates  [330].  Both,  HD  and  PD  have  been  shown  to  be  connected  to  elevated  autophagy  and  noteworthy,  autophagy  was  only  triggered  by  an  aggregate  prone  mutant  of  huntingtin  and  not  by  the  wild-­‐type  form.  Cytosolic  α-­‐synuclein   aggregates   can   be   degraded   by   macroautophagy   and   CMA  [369,370,371,372].   In   ALS   loss   of   motor   neurons   deprives   patients   of   voluntary  controlled   muscle   movements.   The   disease   is   associated   with   ubiquitinated,   p62  positive   protein   inclusions   of   TDP-­‐43   (TAR   DNA   binding   protein   43)   or   SOD1  (superoxide   dismutase   1)   or   rare   mutations   in   a   subunit   of   the   ESCRT   complex  [373,374].   A   defective   ESCRT   complex   in   its   turn   has   been   shown   to   result   in  autophagosome  accumulation  [341].  But  also  point  mutations  of  the  p150  subunit  of  dynactin   resulting   in   defects   in   the   transport   machinery   along  microtubules   have  been   implicated   in   ALS.   Transport   along   microtubules   is   necessary   for  autophagosome-­‐lysosome   fusion   and   therefore   crucial   for   functional   autophagy  [375,376].  Extensive  alterations  in  macroautophagy  can  also  be  found  in  patients  with  AD.  In  an  immuno-­‐electron  microscopy  study  on  neocortical  biopsies   from  AD  patients,  were  autophagosomes,   multivesicular   bodies,   multilamellar   bodies,   and   cathepsin-­‐containing   autophagolysosomes   the   predominant   organelles   and   occupied  most   of  the   cytosol   of   dystrophic   neurites.   Autophagy   was   seen   in   cell   bodies   with  neurofibrillary   pathology   and   was   associated   with   a   relative   depletion   of  mitochondria   and   other   organelles.   It   was   speculated   that   the   accumulations   of  immature   autophagic   vacuoles   result   from   impaired   transport   to   and   fusion   with  lysosomes   thereby   hampering   protective   effects   of   autophagy   [377].   Disruption   of  lysosomal  proteolysis  in  primary  mouse  cortical  neurons  by  inhibiting  cathepsins,  or  by   supressing   lysosomal   acidification,   impaired   transport   of   autolysosomes,  endosomes   and   lysosomes   and   led   to   accumulations   of   these   structures   within  dystrophic  axonal  swellings.  Such  a  phenotype  can  also  be  seen  in  numerous  mouse  models   of   AD.   The   phenotype  was   not   caused   by   general   disruption   of   the   axonal  transport   machinery,   as   mitochondria   and   cathepsin-­‐lacking   organelles   were   not  influenced   in   their   movements.   Once   lysosomal   function   was   restored   the   axonal  dystrophy  was  reversed  [378].  Enhanced  lysosomal  cathepsin  activity  (achieved  by  

Introduction    

   

 48  

genetically   deletion   of   cystatin   B,   an   endogenous   lysosomal   protease   inhibitor)  increased   autophagy   function   and   rescued   the   autophagy-­‐lysosomal   related  pathology   in   the   Alzheimer’s   disease   mouse   model   TgCRND8.   This   intervention  prevented   also   behaviour   defects,   measured   by   learning   and   memory   in   fear  conditioning   and   odour   habituation   tests   [379].   Aβ42   also   induced  neurodegeneration   mediated   by   age-­‐dependant   autophagy-­‐lysosomal   injury   in   a  Drosophila   model   of   Alzheimer’s   disease   [253].   Recently,   the   age   dependence  was  shown   to   be   of   high   importance   as   brain   ageing   is   accompanied   with   increasing  defects   of   the   autophagy-­‐lysosomal   system   and   accumulation   of   dysfunctional  autophagosomes  and  autolysosomes.  As  a  consequence  are  intracellular  membranes  and   organelles   damaged.   Such   changes   could   be   achieved   in   young   Drosophila  expressing   Aβ42   and   this   raised   the   question   if   chronic   deterioration   of   the  autophagy-­‐lysosomal   system   by   Aβ42   simply   accelerates   brain   ageing   [380].   This  concept   is   supported   by   a   work   done   three   years   earlier   with   Drosophila.   The  decreased   expression   of   autophagy   genes  was   shown   to   be   part   of   normal   ageing,  and   disruption   of   the   autophagy   pathway   reduced   lifespan   of   flies.   A   genetically  induced   enhancement   of   autophagy   extended   the   average   survival   and   promoted  resistance   to  oxidative  stress,  at   the  same   time  as  age  dependent  accumulations  of  ubiquitinated  proteins  was  reduced  [381].    We  are   in   the  beginning  of  understanding   the  role  of  autophagy   in  pancreatic  beta  cells  and   its   connections   to   type  2  diabetes.  Beta  cell   specific  disruption  of  Atg7   in  transgenic  mice   led   to  reduced  beta  cell  mass,  accumulation  of  ubiquitin-­‐  and  p62-­‐containing   inclusions,   swollen  mitochondria   and  distended  ER.  Mice  with  depleted  Atg7  in  beta  cells  developed  hypoinsulinaemia  and  hyperglycaemia.  Beta  cells  from  these  mice   were   defective   in   glucose-­‐induced   Ca2+   increase.   Taken   together   these  results   point   to   a   physiological   role   of   basal   autophagy   in   the   maintenance   of  structure,   mass   and   function   of   pancreatic   beta   cells   [382,383].   The   formation   of  autophagosomes  was  also  up-­‐regulated  in  pancreatic  beta  cells  of  diabetic  db/db  and  in  non-­‐diabetic  high-­‐fat-­‐fed  C57BL/6  mice  [382].  Very   little   is   known   about   the   influence   of   IAPP   aggregation   on   autophagy.   In  HIP  rats  an  increased  number  of  autophagosomes  and  presence  of  cytosolic  p62-­‐positive  inclusions  was  detected  when  compared  with  wild-­‐type  rats.  Already  obesity  per  se  up-­‐regulated  autophagy  and  might  be  an  effect  of  increased  protein  burden  as  such  animals   developed   insulin   resistance   that   is   compensated  with   higher   insulin   and  IAPP   production.   However,   hemizygous   HIP   rats   exhibited   a   more   prominent  autophagy   induction   and   it   is   tempting   to   ascribe   this   effect   to   the   amyloidogenic  character   of   hIAPP.   Furthermore,   when   hIAPP  was   expressed   in   INS   832/13   cells  was   autophagy   increased   compared   to   cells   expressing   the   non-­‐amyloidogenic  ratIAPP   (rIAPP).   The   induction   of   autophagy  had   a   cell   protective   effect   since   p62  RNA  silencing   in  hIAPP  cells   led  to   increased  cell  death,  while  p62  over-­‐expression  ameliorated  cell  survival  [384].  There  is  a  complex  entanglement  between  autophagy  and  cell  death.  The  underlying  mechanisms   of   this   cross-­‐regulation   are   largely   unknown.   It   has   been   reported  

Introduction    

   

  49  

numerous  times  that  cell  death  can  be  induced  via  autophagy  (named  autophagic,  or  type  II,  cell  death)  but  autophagy  has  also  been  shown  to  be  protective  against  cell  death   by   inhibiting   apoptosis   [385,386,387,388].   Autophagic   cell   death   is   distinct  from   apoptosis   and   necrosis   even   though   it   sometimes   shares   common   features  [388].  Lately  there  have  been  reports  from  independent  groups  describing  cell  death  in   glioma   cells   and   the   midgut   of   developing   Drosophila   respectively   that   was  dependant   on   autophagy   but   not   on   effector   caspases   or   lysosomal   damage  [389,390].   In   case   of   midgut   cells   in   Drosophila   this   apoptosis-­‐independent,  autophagy-­‐dependant,  programmed  cell  death  was  seen  despite  the  presence  of  high  caspase  activity  in  these  cells  [389].      

   

                                                                                     

   

 

 

                                           

Aims  of  the  Thesis

Aims  of  the  Thesis      

 

 52  

The  term  amyloid  was  introduced  to  the  scientific  field  more  than  150  years  ago,  and  already   in   the   beginning   of   the   20th   century   was   the   association   between   islet  amyloid   and   type   2   diabetes   described.   Since   then,   many   researchers   have   spent  time   and   efforts   to   understand   more   about   amyloid   and   its   relation   to   different  diseases.  In  the  past  5  years  I  have  been  a  part  of  this  research  in  order  to  increase  our   knowledge   on   islet   amyloid   polypeptide   (IAPP).   I   have   explored   IAPP   toxicity,  and   disclosed   some   of   the   cellular   mechanisms   that   are   activated   by   IAPP-­‐aggregates.   For   this  purpose   I   set  different   aims   that   I   tried   to   achieve  during   this  time:    

• To  set  up  a  new  tool  that  allows  to  monitor  the  activation  of  apoptosis  in  vitro  (paper  I).  

 • To   identify  which   species   of   proIAPP   or   its  metabolites   is  most   toxic  when  

added   to  cells:   solubilised  protein,  mature   fibrils,  or  a  mixture  of  preformed  fibrils  and  solubilised  protein  (paper  I).  

 • To   establish   Drosophila   melanogaster   as   model   system   to   study  

aggregation/amyloid  formation  of  proIAPP  and  IAPP  in  vivo  (paper  II  and  III).    

• To  determine  in  vivo  which  cellular  mechanisms  are  altered  upon  intracellular  presence  of  aggregation  prone  hproIAPP  and  hIAPP  and  to  investigate  if  these  mechanisms  are  involved  in  the  described  toxicity.  The  identification  of  such  pathways   allows   us   to   understand   more   about   initial   events   that   might   be  involved   in   islet   amyloid   formation   and   subsequently   lead   to   the  manifestation  of  type  2  diabetes  (paper  III).    

                                 

 

                                     

Material  and  Methods  

 54  

Detailed  description  of  materials  and  protocols  for  all  methods  used  can  be  found  in  the   papers   and  manuscript.   Here   I  want   highlight   some   of   the   pitfalls,   advantages  and  disadvantages  that  should  be  taken  into  account.  

Working  with  Drosophila  

P-­‐element  insertion  

At  the  time  we  generated  transgenic  Drosophila  strains  we  decided  for  an  approach  with  P-­‐element  mediated  transformation  [391].  Because  the  transgene  insertion  site  is  random  there  will  be  variations  in  the  efficiency  of  gene  transcription.  In  our  case  the  gene  of  interest  was  inserted  together  with  a  mini-­‐white  gene  that  is  responsible  for   red   eye   colour.  Micro-­‐injection  was   performed   in   a  white  mutant   background,  and   red   eyes   can   only   be   found   in   flies   with   successful   P-­‐element   insertion.   The  intensity  of  red  colour  depends  on  the  transcription  rate  of  the  mini-­‐white  gene  and  provides  a  hint  of  transcription  rates  for  the  adjacent  gene  of  interest.  An  additional  advantage  with  the  random  insertions  is  the  possibility  to  combine  transgenic  lines  with  different  insertion  loci.  It  should  be  mentioned  that  random  insertion  can  lead  to   disruption   of   Drosophila   genes   and   have   detrimental   effects.   When   designing  experiments   it   is   therefore   important   to   control   for   effects   caused   by   P-­‐element  insertion.   Hence,   in   paper   II,   we   included   transgenic   hproIAPP   flies   that  were   not  crossed  with  the  used  Gal4  driver  line  in  our  survival  experiment.  In   recent   years,   a   new   technique   has   emerged   that   allows   for   gene-­‐insertion   at  distinct,   predefined  gene   loci.  An  advantage  with   this   is   the  possibility   to   compare  the  effects  of  inserted  genes  that  encode  for  different  proteins.  On  the  downside  one  loses  the  opportunity  to  randomly  generate  flies  with  different  gene  expression  rates  and  recombine  them.    

Survival  assay  

We  conducted  numerous  survival  assays,  and  one  obvious  advantage  of  such  assay  is  the  fast  read-­‐out  in  form  of  dead  or  alive  flies.  For  each  survival  assay  I  used  100-­‐150  flies  with  25  flies  per  vial  kept  at  constant  temperature  and  humidity  with  a  12  hours  dark-­‐light   cycle.   Despite   these   high   demands   I   have   experienced   fluctuations  between  survival-­‐assays  performed  at  different  time  points.  E.g.  the  median  survival  for   hproIAPP   expressing   flies   varied   but   was   always   significantly   shorter   than  estimated   for   control   flies.   The   reason   for   the   pitfalls   in   survival-­‐assay   could   be  differences   in   food  quality  e.g.   the  surrounding  relative  humidity  will   influence   the  water  content  in  food  left  to  cool  down  after  preparation.  Flies  can  drown  in  wet  and  become  dehydrated  in  dry  food.  Therefore  we  used  mean  survival  data  from  a  single  

Material  and  Methods    

 

 

  55  

experiment   and   used   information   from  multiple   experiments   as   a   description   of   a  trend.  

Detection  methods  

Immunofluorescence  –  tissue  preparation  

Many   results   in   this   thesis   are   based   on   immunofluorescence.   Material   for   these  experiments   was   obtained   in   different   ways.   In   several   experiments   heads   from  decapitated   flies   were   embedded   in   Tissue-­‐Tek®   O.C.T.TM   and   sectioned.   Several  sections   from   different   transgenic   animals   were   placed   on   the   same   slide.   This  assures   that   antibody   incubation  was   equal   for   all   specimens   placed   on   the   same  slide  and  this  allows  a  better  comparison  of  the  results.  Sections  of  fly  heads  allowed  us  to  analyse  antibody  binding  not  only  in  the  CNS  but  also  in  the  surrounding  tissue,  such  as  the  detection  of  hIAPP  and  hproIAPP  in  the  head  fat  body.  A  drawback  with  the  10  μm  thick  head  sections  is  the  lack  of  three-­‐dimensional  information.  Analysis  of   consecutive   sections   can   partly   circumvent   this   problem.   However,   analysis   of  dissected   brains   from  Drosophila   offers   a  more   sophisticated  way   to   obtain   three-­‐dimensional  data.  Counting  LNvs  in  dissected  brains  is  a  more  exact  procedure  and  also  less  work–intensive  compared  to  sections.  A  disadvantage  with  dissected  brains  can  be  encountered  when  quantifying  absolute  fluorescence  signals.  LNvs  are  located  at  different  levels  in  the  brain  and  the  distance  the  fluorescent  light  has  to  penetrate  varies   and   subsequently   the   absolute   fluorescence.   In   addition,   dissected   whole  brains  exclude  the  opportunity  to  analyse  surrounding  tissue,  such  as  the  fat  body.  

Congo  Red  or  pFTAA  

The   classic   definition   of   amyloid   includes   affinity   for   the   dye   Congo   red   and   the  appearance  of  green  birefringence  when  viewed  in  cross-­‐polarized  light.  The  fly  head  capsule  has  strong  affinity  for  Congo  red  (the  same  applies  to  pFTAA)  and  results  in  green   birefringence,   which   makes   it   difficult   to   detect   small   amyloid   aggregates.  Another  drawback  is  that  Congo  red  staining  does  not  allow  for  parallel  staining  with  fluorophore-­‐conjugated   antibodies,   as   these   require   mounting   in   water   based  mounting  medium  (e.g.  1:1  PBS/glycerol)  that  dissolve  the  Congo  red  binding.  This  is  different  from  staining  amyloid  with  pFTAA  that  is  solubilised  in  PBS.  pFTAA  can  be  combined  with  immunofluorescence,  and  used  for  co-­‐localisation  studies.  However,  pFTAA  does  not  solely  define  amyloid  deposits  instead,  pFTAA  can  also  bind  to  other  highly  ordered  protein  structures.  

Material  and  Methods      

 

 56  

Image  processing  

After   image   acquisition   there   are   programs   available   for   image   processing   that  facilitates  presentation  of  data.  However,   it   is   important   to  keep   in  mind   that   such  powerful   tools   can   provide   misleading   results   when   used   frivolous.   Especially  quantification   of   reactivity   requires   a   careful   stringent   approach   and   only   images  that  are  acquired  under  identical  circumstances  can  be  used.                        

 

                                     

Results  and  Discussion                            

Results  and  Discussion      

 

 58  

 

Extracellular  amyloid  formation  induces  apoptosis  (paper  I)  

This  work   introduces   a   novel   detection   assay   for   caspase-­‐3-­‐like   activation   in   beta  cells.   The   system   was   used   to   investigate   the   effects   of   proIAPP   and   the   three  putative   processing   metabolites   N-­‐IAPP,   IAPP,   and   IAPP-­‐C   on   beta   cells.   We  compared   the   putative   toxic   effects   exerted   by  mature   amyloid   fibrils,   solubilised  protein,  and  the  process  of  fibril  formation.  The  system  is  based  on  the  measurement  of  fluorescence  resonance  energy  transfer  (FRET).   We   transfected   murine   Beta-­‐TC-­‐6   cells   with   a   vector   construct   (pFRET2-­‐DEVD)  that  codes  for  a  fusion  protein  made  up  by  enhanced  cyan  fluorescent  protein  (ECFP,  emission  wavelength  480nm)  linked  to  enhanced  yellow  fluorescent  protein  (EYFP,   emission   wavelength   535nm)   and   established   a   stable   cell   line.   The   two  fluorophores  are  linked  by  the  caspase-­‐3  specific  cleavage  site  residues  DEVD.  When  the   two   fluorophores   are   linked   to   each   other   is   energy   transferred   from  ECFP   to  EYFP  via  FRET.  If  caspase-­‐3  is  activated  it  will  cleave  the  fusion  protein  at  its  DEVD  site  and  the  two  fluorophores  will  disperse.  As  a  result  will  FRET  from  ECFP  to  EYFP  be  lost  in  such  cells  as  the  energy  transfer  only  occurs  when  the  fluorophores  are  in  very  close  spatial  proximity.  Loss  of  FRET  can  be  used  as  a  direct  measurement  of  caspase-­‐3   activity.   The   excitation  wavelength   in   our   experiments  was  440  nm  and  emission   was   measured   at   480   nm   (ECFP)   and   535   nm   (EYFP)   respectively.   The  more   FRET   occurs,   the   more   light   with   535   nm   is   emitted.   Loss   of   FRET   was  monitored   as   a   decrease   in   535   nm/480   nm   ratio.   The   535   nm/   480   nm   ratio  decreased   within   4   hours   from   initially   2.2   to   1.2   upon   addition   of   the   apoptotic  agent   staurosporine.   This   method   has   several   advantages   compared   to   other  apoptosis-­‐detection   assays:   since   apoptosis   is   a   transient   event   it   is   of   interest   to  follow   the   same   cell   population   over   time   with   several   measurements.   Staining  methods   such   as   TUNEL   (detecting   DNA   fragmentation),   the   vybrant   apoptosis  detection   kit   or   the   apoptosis   assay   using   Ac-­‐DEVD-­‐AMC   are   all   restricted   to   one  single  time  point  and  therefore  do  not  offer  time  studies  in  the  same  way.  The  same  is  true  for  the  frequently  used  MTT  assay,  which  in  fact  is  not  a  true  apoptosis  assay  since   it  measures   living   cells.  With   the  new  method  described  one  doesn’t  have   to  consider  factors  such  as  cell  density  or  rate  of  cell  division  since  the  number  of  cells  does   not   influence   the  measured   ratio   of   emitted   light.   This   is   not   true   for   assays  where  absolute  values,  such  as  number  of  stained  cells,  are  measured.  A  drawback  of  the  here  described  conditions  for  the  FRET  assay  is  the  relatively  poor  medium  that  limits   cell   survival   to   up   to   12   hours.   It   would   be   desirable   to   define   an   assay  condition  that  prolongs  cell  survival.  We   analysed   in   which   form   IAPP   was   capable   of   inducing   apoptosis.   Addition   of  solubilised   synthetic   IAPP   did   not   activate   caspase-­‐3,   but   a   combination   of  

Results  and  Discussion    

 

 

  59  

solubilized  IAPP  with  preformed  IAPP  fibrils  induced  caspase-­‐3  activity.  In  a  parallel  experiment   with   a   mixture   of   solubilized   IAPP   and   preformed   fibrils   was   fibril  formation   monitored   with   a   ThT   assay.   Samples   from   different   time   points   were  investigated   with   an   electron   microscope,   and   fibrillation   over   time   could   be  confirmed.  Since  extracellular  fibril  formation  of  hIAPP  induced  apoptosis,  we  asked  if  fibril  formation  of  hproIAPP  or  the  processing  intermediates  had  similar  effects  or  not.   We   therefore   expressed   recombinant   peptides   corresponding   to   proIAPP  (recproIAPP),  IAPP  with  the  N-­‐terminal  flanking  peptide  (recN+IAPP);  IAPP  with  the  C-­‐terminal   flanking   peptid   (recIAPP+C)   and   fully   processed   IAPP   (recIAPP).   All  peptides   were   expressed   in   E.   coli   as   fusion   protein   with   a   N-­‐terminal   GST  (glutathione  S-­‐transferase)-­‐tag.  The  GST-­‐tag  was  enzymatically  removed,  which   led  to   spontaneous   amyloid-­‐like   fibril   formation   of   all   recombinant   peptides   as  confirmed   with   Cong   red   staining   and   TEM.   Peptides   were   added   to   the   cells   as  preformed  amyloid-­‐like   fibrils,   in  monomeric   form  (50  μM),  or   in  monomeric   form  (50   μM)  mixed   together  with   preformed   synthetic   hIAPP   (30   nM)   fibrils.   Only   the  mixed  solution  led  to  induction  of  apoptosis  and  connected  reduction  in  cell  viability.  There  was  no  significant  difference  in  the  degree  of  cell  death  irrespective  of  which  hproIAPP  metabolite  was  added  together  with  hIAPP-­‐fibrills   to   the  Beta-­‐TC-­‐6  cells.  The   cell   viability   after   12   hours   compared   to   a   negative   control   consisting   of  solubilized   hIAPP   without   additional   seeds,   was   as   follows:   recproIAPP:   51%,  recN+IAPP:  54%,  recIAPP:  28%,  recIAPP+C:  41%.    In  general,  apoptosis  has  been  described  to  be  responsible  for  the  reduced  beta  cell  mass   found   in  patients  with   type  2  diabetes   [152].   Islet   amyloid   formation  and/or  presence  has   been   shown   to   induce   apoptosis   and   therefore   been   suggested   to   lie  behind  the  observed  beta  cell  loss  connected  to  type  2  diabetes  [57,58,142,316,392].  In   two  different  studies  were  apoptotic  elements  coupled   to   the  extrinsic  pathway,  namely  caspase-­‐8  and  FAS-­‐associated  death  receptor  signalling,  identified  [57,58].  It  has   previously   been   shown   by   our   group   that   intracellular   IAPP   amyloid   fibrils   in  beta   cells   of   mice   and   humans   contain   proIAPP   and/or   incomplete   processed  proIAPP  [142].  Several  in  vitro  experiments  have  further  strengthened  the  idea  that  aberrant  processing  of  proIAPP  might  play  a  crucial   role   in  amyloid   formation  and  subsequent  toxicity  [139,140,141,143,393].  All  together,  the  results  presented  in  this  study   show   that   proIAPP   and   the   processing   intermediates   contain   the   same   cell  toxic  capacity  as   IAPP,  and  underline   that   the   fibril  propagation  can  have  a  central  role  in  the  beta  cells  reduction  in  type  2  diabetes.      

Results  and  Discussion      

 

 60  

 

Characterisation  of  a  new  Drosophila  model  for  studies  of  IAPP  aggregation  (paper  II)  

This  paper  is  the  first  description  of  a  Drosophila  melanogaster  model  for  studies  of  the   effects   of   IAPP   and   proIAPP   aggregation.   IAPP   belongs   to   the   calcitonin   gene  peptide  family  together  with  calcitonin,  calcitonin  gene-­‐related  peptide,   intermedin  and   adrenomedullin   [88].   Drosophila   lacks   nucleotide/amino   acid   sequence  similarity  to  any  of  these  genes/proteins  making  it  plausible  that  observed  effects  of  (pro)IAPP  are  due   to   its  ectopic  expression.  When  setting  up   this  new   IAPP  model  system   we   wanted   to   study   the   consequences   of   protein   aggregation   and   we  therefore  included  the  non-­‐amyloidogenic  mouse  IAPP  (mIAPP)  as  control  to  all  our  experiments.   The   absence   of   propagation   propensity   for   mIAPP   and   its   sequence  homology   to   human   IAPP   (hIAPP)   makes   it   a   perfect   control   that   can   help   us   to  distinguish  between  effects  caused  mere  protein  over-­‐expression  and  effects  due  to  protein   aggregation.   Furthermore,  mIAPP-­‐expressing   flies   can   be   used   as   negative  controls  when  staining  with  amyloid-­‐specific  dyes.  Insertion  of   the  Gal4  dependant   transgenes   into   the  Drosophila   genome   is   random  and  it  was  important  to  make  sure  that  this  insertion  itself  does  not  introduce  new  phenotypes.  Multiple  transgenic  lines  were  established.  We  chose  three  independent  lines   that   were   unique   in   their   transgene   insertion   site   for   each   transgene  (hproIAPP,   hIAPP,   and   mIAPP)   and   determined   the   mRNA   levels   as   well   as   their  survival   when   transgene   expression   was   driven   to   the   CNS   (elavC155,Gal4).   MRNA  expression   levels   for  each   transgene  were  dependant  on   the   insertion  site  but   this  did  not  significantly  influence  the  survival.  Neither  hIAPP  nor  mIAPP  expressing  flies  had  a  significant  alteration  in  survival  compared  to  control  flies  that  only  produced  Gal4  in  the  CNS.  However,  hproIAPP  expression  had  toxic  effects  and  for  two  out  of  three   transgenic   lines,   reduction   of   lifespan   was   significant   when   compared   to  control   flies   (p<0.02).   Also   the   third   line   lived   shorter   than   control   flies   but   this  effect  did  not  reach  statistical  significance  (p  =  0.0577).  Interestingly,  the  transgenic  proIAPP  line  with  the  shortest  lifespan  had  the  lowest  mRNA  expression  levels.  The  estimated   mRNA   levels   in   this   hproIAPP   line   were   most   comparable   to   the  expression  levels  detected  for  hIAPP  and  mIAPP  transgenes  and  we  therefore  chose  this  hproIAPP  line  together  with  one  mIAPP  and  one  hIAPP  transgene  for  all  further  experiments.    However,   the   choice  of  Gal4  driver   line  and   connected   site  of   expression  mattered  for  hproIAPP  toxicity.  Only  when  expressed  in  the  CNS  by  elavC155,Gal4  led  proIAPP  to  accelerated   death   of   the   whole   organism.   With   this   Gal4   driver   was   hproIAPPs  relative  toxicity  (when  compared  to  suitable  control  flies)  persistent  though  and  not  influenced   by   increase   or   decrease   in   temperature.   The   observed   toxicity   for  

Results  and  Discussion    

 

 

  61  

hproIAPP   was   confirmed   to   be   dependant   on   hproIAPP   expression   as   transgenic  hproIAPP   flies   lacking   a   Gal4   driver   displayed   no   such   shorter   survival.   If   not  otherwise  stated  was  the  expression  of  the  respective  transgenes  for  all  subsequent  experiments  driven  to  CNS  by  elavC155,Gal4.  Minor   amounts   of   both   hIAPP   and   hproIAPP   could   be   detected   by  immunofluorescence  at  site  of  expression.  But  the  majority  of  the  respective  proteins  were  found  in  the  neighbouring  head  fat  body.  Unspecific  cross-­‐reactions  of  the  used  antibodies   could   be   excluded,   as   Gal4   driver   flies   devoid   of   hproIAPP   of   hIAPP  showed  no  such  staining  at  neurons  or  in  the  fat  body.  Gal4  dependant  expression  of  GFP  with  a  nuclear  localisation  sequence  did  not  lead  to  a  GFP  signal  in  the  fat  body  when  the  same  elavC155,Gal4-­‐driver  was  used.  This  additional  experiment  allowed  us  to  exclude   the   presence   of   Gal4   in   the   fat   body   elavC155,Gal4   driver   flies   (unpublished  data).  All  expressed  proteins  contained  a  signal  peptide  determining  the  protein  for  secretion   and  we   concluded   that   proteins   found   in   the   fat   body  were   successfully  secreted   from   neurons   and   finally   delivered   to   or   taken   up   by   the   fat   body   for  storage   and/or   degradation.   This   finding   is   not   very   surprising   taking   in  consideration  that  the  fly  fat  body  partly  corresponds  to  the  mammalian  liver  and  is  involved   in   immune   responses.   Flies   expressing   TTR   in   the   eye   also   show  accumulations  of  the  protein  in  the  fat  body  [259].  The  amount  of  hproIAPP  or  hIAPP  found  within   the   humoral/CNS   barrier   differed   depending   on   the   age   of   the   flies.  Amyloid   aggregation   in   man   is   related   to   age   and   we   expected   to   find   more  hIAPP/hproIAPP   in   neurons   from   old   flies.   Surprisingly,   this   was   not   the   case.  Instead  more  hIAPP  and  hproIAPP  was  found  in  5d  and  15d  young  flies  as  compared  to   40d   old   flies.   This   prompted   us   to   investigate   the   levels   of   Gal4   over   time.  Expression  of  Gal4  dependant  nlsGFP  allowed  us   to   follow   the  occurrence  of  Gal4.  When  comparing  GFP  intensity  in  1,  5,  15  and  30d  old  flies  it  was  obvious  that  green  fluorescence   was   strongest   at   day   5,   slight   reduction   occurred   at   day   15   and   the  signal  was  almost  absent  after  30  days.  This  decrease  of  Gal4  expression  levels  with  age  can  explain  the  low  amount  of  hproIAPP  and  hIAPP  in  neurons  of  old  flies  when  compared  to  younger  flies.    At   this   stage   we   had   shown   toxicity   due   to   pan-­‐neuronal   hproIAPP   expression  whereas  hIAPP  did  not  alter   longevity.  On  the  other  hand  we  were  not  able  to   find  any   differences   in   protein   abundance   or   tissue   distribution.   In   human,   proIAPP   is  processed   at   two   sites   by   prohormone   convertases   1/3   and   2   and   Drosophila  contains   a   homolog   to   PC2,   amontillado.   It   can   be   speculated   that   processing   of  hproIAPP  by  amontillado  could   lead   to   intracellular   imbalances   that  cause   toxicity.  However,   processing   of   hproIAPP   by   amontillado   is   unlikely   as   shown   by   the  presence  of  the  N-­‐terminal  processing  site  of  hproIAPP  and  the  C-­‐terminal  flanking  peptide   of   hproIAPP   in   the   fat   body.   The   presence   of   hproIAPP   was   verified   by  immunolabelling   with   antibodies   specific   for   these   two   processing   sites   [143].  HproIAPP   in   the   fat   body   has   been   produced   and   secreted   from   neurons   in   its  unprocessed  form.  

Results  and  Discussion      

 

 62  

The   presence   of   hIAPP   and   hproIAPP   within   the   humoral/CNS   barrier   raised   the  question  whether  the  detected  proteins  are  intra-­‐  or  extracellular.  We  designed  flies  that  co-­‐expressed  Gal4-­‐dependant  nlsGFP,  with  the  majority  of  GFP  expressed  in  the  nucleus  and  only  a  small  portion  of  GFP  leaking  to  the  cytosol.  With  GFP  as  a  cellular  marker  we  could  show  that  hproIAPP  and  hIAPP  was   found   intracellular   replacing  the  main   portion   of   the   cytosol.   However,   in   several   cases   the   proteins  were   also  detected   extracellular.   As   secreted   protein   can   be   expected   to   rapidly   diffuse   we  concluded   that   detected   extracellular   proteins   were   due   to   protein   aggregate  depositions  representing  areas  with  high  local  protein  concentrations.  Extracellular  deposits   were   often   found   at   areas   devoid   of   cell   nuclei.   Similar   findings   of  intracellular   IAPP   replacing   the   cytosol   can  be  done   in   hIAPP   transgenic  mice   and  there  it  associated  with  cell  death  caused  by  apoptosis  [142].  To   confirm   the   presence   of   amyloid   like   aggregates   in   flies   sections   of  Drosophila  heads   were   stained   with   amyloid   specific   dyes,   such   as   Congo   red   and   pFTAA  (pentameric  formic  thiophene  acetic  acid)  [394,395].  Both  dyes  exhibit  high  affinity  for   the   chitin   rich   exoskeleton   of   Drosophila   [396].   However,   flies   expressing  hproIAPP  and  hIAPP  both  contained  structures  in  the  neurons,  but  more  frequent  in  the   fat   body,   with   affinity   for   the   dyes.   This   presence   of   pFTAA   and   Congo   red  staining  besides   the  exoskeleton  was  absent   in  mIAPP  expressing   flies   and   control  flies   that  only  expressed  Gal4.  The  observed,  higher  prevalence  of  pFTAA   labelling  compared   to   Congo   red   is   probably   due   to   pFTAAs   capacity   to   bind   pre-­‐amyloid  aggregates   [397,398].   Co-­‐staining   of   pFTAA   with   an   IAPP-­‐specific   antibody  confirmed  that  pFTAA  positive  structures  were  made  up  of  hIAPP/hproIAPP.  Several  areas  were  only  stained  with  the  antibody  but  not  with  pFTAA.  But  all  structures  that  stained  with  pFTAA  also  bound  the  IAPP-­‐specific  antibody.  Staining  with  pFTAA  was  often  found  in  close  vicinity  to  nuclei.  With  transmission  electron  microscopy  (TEM)  we  detected  highly  ordered,  electron  dense  aggregates  surrounding  fat  body  nuclei.  This  ultra-­‐structural  analysis  of  the  fat  body  revealed  two  distinct  populations  of  aggregates;  an  ordered,  electron  dense  and  a  less  ordered,  lighter  structure.  Both  structures  were  found  in  hproIAPP  and  hIAPP  expressing  flies  and  were  completely  absent  in  flies  producing  mIAPP  or  Gal4  only.  Even  though  we  failed  to  stain  these  aggregates  with  an  antibody,  we  are  convinced  that  they  are  most  likely  represent  hproIAPP  or  hIAPP.  Absence  of  antibody  labelling  of  tissue  prepared  for  EM  most   likely   is  due  to  epitope  blockage  caused  by  fixation  and   embedding   in   epon.   The   detected   aggregates   were   structurally   distinct   form  aggregates  found  in  TTR  expressing  flies  [259].  HproIAPP  and  hIAPP  expressing  flies  showed  occasionally  morphological  changes  of  cell   nuclei.   This   was   manifested   as   an   evenly   dotted   pattern   that   had   replaced  heterochromatin   and   euchromatin   and   points   towards   cell   fragmentation   and   cell  death.  However,  this  was  not  accompanied  with  classic  apoptotic  hallmarks  such  as  apoptotic  bodies  or  nucleus  shrinkage.    Taken  together  this  paper  presents  the  characterisation  of  a  novel  Drosophila  model  that   can   be   used   for   studies   on   IAPP   aggregation.   Both   hproIAPP   and   hIAPP   form  

Results  and  Discussion    

 

 

  63  

aggregates   in   this  model   but   only   hproIAPP   exerts   toxicity.   This   finding   opens   for  further   studies   to   understand   more   about   mechanistic   links   between   protein  aggregation  and  toxicity.  Clearly,  protein  aggregation  per  se  is  not  always  toxic  and  it  remains  to  be  elucidated  which  additional  factors  are  important  in  determining  the  outcome  in  regards  to  cell/organism  survival.      

Results  and  Discussion      

 

 64  

 

HproIAPP  and  hIAPP  trigger  selective  autophagy  (paper  III)  

We   took   advantage   of   the   unique   genetic   toolbox   offered   by   the   Drosophila  melanogaster   system   and   investigated   the   effects   hproIAPP   and   hIAPP   have   on  several  molecular  pathways.  The  model  as  it  is  presented  in  paper  II  did  not  allow  us  to  distinguish  between  responses   triggered  by  extracellular  protein  deposition  and  reactions   due   to   intracellular   actions.   We   therefore   modified   our   system   in   two  ways:   (i)   expression   of   all   transgenes   was   reduced   to   8   ventral   lateral   neurons  (LNvs)  using  a  pdf-­‐Gal4  driver  line  [399]  and  (ii)  we  included  co-­‐expression  of  Gal4  dependant  GFP  with  a  nuclear  localization  sequence  (nlsGFP)  if  not  stated  different.  This  nlsGFP  enabled  us  to  monitor  all  8  cells  as  they  fluoresce  green.  Similar  to  the  results  with  a  pan-­‐neuronal  Gal4  driver  (paper  II)  were  hIAPP  and  hproIAPP  present  intra-­‐  and  extracellular.  Comparing  expression  in  brains  of  1  day,  15  day,  and  30  day  old   flies   we   detected   a   reduction   of   LNvs   over   time.   A   similar   decrease   in   LNv  number  with  age  was  observed  for  flies  that  additionally  expressed  mIAPP.  However,  the   loss   of   green   fluorescent   LNvs  was   significantly   higher   for   flies   that   expressed  hproIAPP  or  hIAPP.  There  was  no  significant  difference  in  toxicity  for  hproIAPP  and  hIAPP.  We  had  previously  shown  that  pan-­‐neuronal  expression  of  hproIAPP,  but  not  hIAPP,  shortens  lifespan.  However,  the  two  proteins  affect  viability  on  a  cellular  level  in   similar  ways.  The   loss  of  LNvs   in  hproIAPP  and  hIAPP  expressing   flies  does  not  occur  during  development  as  no  difference   in  LNv  number  was   found   in  1  day  old  flies   expressing   any   of   the   transgenes,   including   mIAPP   and   control   flies   (nlsGFP  only).  Even  though  extracellular  protein  depositions  were  found,  it  is  most  likely  that  loss  of  cell  number  is  due  to  intracellular  events.  Extracellular  deposition  can  occur  at   any  neuron   that   is   close   to  LNvs.  The   chance  of  deposition   surrounding  LNvs   is  low  compared  to  the  probability  of  deposition  next  to  any  other  neuron.  Monitoring   death   of   LNvs   allowed   us   to   distinguish   between   mIAPP   and  hproIAPP/hIAPP.   This   raised   the   questions   about   underlying  mechanisms   for   this  difference.  Apoptosis  has  been  reported  to  be  involved  in  beta  cell  death  of  patients  with  type  2  diabetes  [152].  Similar  results  are  obtained  from  mouse  models  and   in  vitro  studies  investigating  effects  of  IAPP  aggregation  [57,58,142,316,320].  However,  we  were  not   able   to  detect   any   sign  of   apoptosis   in  LNvs  due   to  hproIAPP,  hIAPP,  mIAPP,   or   sole   nlsGFP   expression.   Immunofluorescence   with   an   antibody  recognizing   cleaved   caspase-­‐3   and   TUNEL   staining   are   both   well-­‐established  markers   for  apoptosis  but   failed   to   stain  LNvs   in  15  day  and  30  day  old   flies.  This  was   irrespective   of   the   studied   transgene.   Since   apoptosis   is   a   transient   event  we  wanted   to   extend   the   window   where   caspase   activity   can   be   monitored   and  therefore  made  use  of  an   in  vivo  caspase  activity  sensor,  called  Apoliner  [400].  The  basic  idea  of  Apoliner  resembles  our  previously  used  detection  assay  for  caspase-­‐3-­‐

Results  and  Discussion    

 

 

  65  

like  activation  (paper  I).  Apoliner  terms  a  fusion  protein  of  N-­‐terminal  mRFP  and  C-­‐terminal   eGFP.   A   synthetic   caspase-­‐3   cleavage   site   links   both   fluorophores.   An   N-­‐terminal  mCD8  transmembrane  domain  destines   the   fusion  protein   to  membranes.  Upon   caspase-­‐3   activity   is   this   fusion   protein   cleaved   and   the   two   fluorophores  separated.  Separation  of  the  two  fluorophores  can  be  detected  as  eGFP  relocates  to  the  nucleus  due  to  the  nuclear  localization  sequence  preceding  eGFP.  Expression  of  Apoliner  alone  or   together  with  hproIAPP,  hIAPP,  or  mIAPP  never  resulted   in  such  eGFP   relocation.   A   genetic   approach   to   investigate   the   involvement   of   apoptosis  revealed  similar  results.  Inhibition  of  apoptosis  by  the  viral  caspase  inhibitor  protein  p35   did   not   revert   cell   death   mediated   by   hproIAPP.   Taken   together,   all   results  strongly  suggest  that  cell  death  of  LNvs  is  independent  of  apoptosis.    Reports  of  apoptotic  cell  death  of  beta  cells  do  not  necessarily  contradict  our  results  as   apoptosis   still   might   be   triggered   by   extracellular   protein   aggregates   and/or  amyloid   fibrils.   Actually,   several   reports   show  activity   of  members   of   the   extrinsic  pathway  suggesting  the  initiation  of  apoptosis  due  to  extracellular  events  [57,58].  Viral  over-­‐expression  of  hproIAPP  has  been  shown   to  cause  ER-­‐stress,   induce  UPR  and  lead  to  cell  death  [301].  As  our  proteins  are  secreted  and  travel  through  the  ER  we  considered  the  possibility  that  ER-­‐stress  caused  cell  death.  One  key  event  of  ER-­‐stress   is   splicing   of   Xbp1  mRNA,   thereby   activating   this   transcription   factor   [274].  We  made   use   of   a   Gal4   dependant   Xbp1-­‐GFP   reporter   construct.   In   this   construct,  GFP  mRNA   is   out   of   frame   and   not   transcribed   unless   Xbp1  mRNA   is   spliced.   Co-­‐expression  of  hproIAPP,  hIAPP  or  mIAPP  in  LNvs  did  not   lead  to  Xbp1  splicing  and  subsequent  GFP  signal  after  5,  15  or  30  days.  This  reporter  did  work  though,  as  ER-­‐stress   was   detected   with   Apoliner   once   RNAi   against   hsf   was   expressed.   Hsf   is   a  transcription   factor  known  to  play  a  crucial  role   in  regulating   the  response   to  heat  shock  and  misfolding  proteins  [401,402].  Therefore,  it  is  of  no  surprise  that  lower  hsf  levels   increased   sensitivity   for   ER-­‐stress.   Additional   experiments   underscored   our  finding   that   neither   hproIAPP   nor   hIAPP   trigger   ER-­‐stress.   The   ER   residing  chaperone   BiP   (hsc70)   is   important   in   sensing   ER-­‐stress   and   inducing   UPR   but  immunofluorescence  showed  no  hsc70  accumulation  in  the  ER.    As   low  hsf   levels   reduced   the   capacity  of  ER   to  handle   stress  we   tested   the  effects  caused  by  co-­‐expression  of  hproIAPP/hIAPP  with  RNAi  against  hsf.  Staining  with  an  IAPP  specific  antibody  in  such  cells  lead  to  accumulation  of  hIAPP/hproIAPP  next  to  the  nucleus  and  absence  of  extracellular  protein.  This  result  was  interpreted  to  be  a  direct   consequence   of   general   down-­‐regulation   of   protein   synthesis,   a   possible  outcome  of  mild  UPR.  Low  hsf  levels  are  responsible  for  the  increased  sensitivity  of  the   ER   to  misfolded   proteins.   In   conclusion,  we   can   rule   out   the   exhaustion   of   ER  folding  capacity  evoked  by  hIAPP  and/or  hproIAPP  expression  as  a  cause  for  chronic  ER-­‐stress  leading  to  UPR  and  subsequent  cell  death.  The  absence  of  ER-­‐stress  in  our  model   is   in  accordance  with   findings   from  human  pancreas  of  patients  with   type  2  diabetes  and  a  hIAPP  expressing  mouse  model  [306].    Since   neither   apoptosis   nor   ER-­‐stress   could   be   accounted   for   cell   death  we   asked  which   other   known   cell   death   pathways   could   be   triggered   by   the   presence   of  

Results  and  Discussion      

 

 66  

amyloid   prone   peptides.   Autophagy   has   emerged   as   a   potential   player   in   selective  degradation  of   aggregated  proteins   and  has  been   shown   to  have  protective   effects  against   neurodegeneration.  We   therefore   quantified  mRNA   levels   of   ATG4,   ATG8a,  ATG8b  and  Ref(2)P  (Drosophila  homolog  of  p62)  in  flies  expressing  hproIAPP,  hIAPP,  or  mIAPP  in  the  whole  CNS  (elavGal4,  C155-­‐driver  line).  Only  hproIAPP  expressing  flies  had   elevated   levels   of   all   of   the   investigated   mRNAs.   Noteworthy,   only   hproIAPP  shortens  lifespan  when  expressed  in  the  whole  CNS  (paper  II).  The  co-­‐localization  of  antibodies  recognizing  Atg8a/Atg8b  (Drosophila  homolog  of  LC3)  and  Ref(2)P  with  protein  aggregates  made  up  of  hIAPP  or  hproIAPP  further  strengthened  the  concept  of  autophagy  involvement.  An  autonomously  induction  of  autophagy  in  response  to  hIAPP  and  hproIAPP  expression  was  again  shown  in  LNvs.  Expression  of  the  fusion  protein   mCherry-­‐Atg8a   reports   autophagic   activity   as   cytoplasmic   mCherry-­‐Atg8a  gets   recruited   to   autophagosomal   membranes   and   concentrates   in   autolysosomes  upon   autophagosome-­‐lysosome   fusion.   Only   hIAPP   and   hproIAPP,   but   not  mIAPP,  led   to   an   accumulation   of   this   autophagy  marker   in   LNvs.   In   a   next   step,  we  were  able   to   show   that   mCherry-­‐Atg8a   accumulation   is   not   a   consequence   of   defective  autophagosome-­‐lysosome   fusion.  We   used   the   pH   sensitive  marker  mCherry-­‐GFP-­‐Atg8a.   Both   fluorophores,   GFP   and   mCherry,   emit   light   from   autophagosomes  whereas   only   red   signal   (mCherry)   can   be   seen   from   autolysosomes   since   GFP   is  unstable  in  the  acidic  environment  of  autolysosomes.  Co-­‐expression  of  this  reporter  with   hproIAPP   or   hIAPP   resulted   only   in   red   fluorescence   consistent   with   an  accumulation  of  autolysosomes.    HproIAPP   and   hIAPP   do   not   lead   to   defective   autophagy.   But   is   autophagy   still  involved   in   toxicity?  To  address   this  question  we   tested   if   genetic  up-­‐regulation  of  autophagy  can  hamper  the  increased  death  of  hproIAPP  and  hIAPP  expressing  LNvs.  We  made  use  of  a  TOR-­‐inhibitor  (TOR-­‐TED),  which  was  co-­‐expressed  together  with  hproIAPP,   hIAPP,   mIAPP   or   only   nlsGFP.   Autophagy   up-­‐regulation   completely  restored  survival  of  LNvs  that  expressed  hIAPP  and  hproIAPP.  Neuronal  survival  was  now  similar  to  control,  and  mIAPP  expressing  flies.  Previously,  autophagy  has  been  ascribed  a  neuroprotective  effect   in  aging  by  promoting  cell   survival  and   longevity  through  increased  resistance  to  accumulation  of  ubiquitinated  and  oxidized  proteins,  and   oxidative   stress   [381,403].   One   can   speculate   that   this   protective   effect   is  hampered   if   cells   are   challenged   with   aggregating   hIAPP   or   hproIAPP.   This  hypothesis   was   tested   by   down-­‐regulation   of   autophagy   with   RNAi   against   Atg8a  and   Atg8b.   If   hIAPP   and   hproIAPP   exert   toxicity   by   reducing   protective   effects   of  autophagy,  down-­‐regulation  of  autophagy  should  have  similar  effects  as  sole  hIAPP  or   hproIAPP   expression   has.   Counting   the   number   of   LNvs   over   time,   we   saw   a  significant   decrease   in   cell   viability   once   autophagy   was   down-­‐regulated   and  survival   rates   dropped   to   levels   that   are   comparable   to   hproIAPP   and   hIAPP  producing  cells.  In  addition,  down-­‐regulation  of  autophagy  did  not  potentiate  hIAPP  or   hproIAPPs   toxicity.   Taken   together   this   indicates   that   hIAPP   and   hproIAPP  expression  per  se  already  hampers  the  neuroprotective  effects  of  autophagy.  

Results  and  Discussion    

 

 

  67  

Previously,  it  has  been  shown  that  Ref(2)P  and  ubiquitin  positive  protein  aggregates  accumulate   in   brains   of   aged   Drosophila   –   this   has   been   suggested   to   be   a  consequence  of  decreased  autophagy  capacity  over  time  [356,404].  Ubiquitin  is  a  tag  known  to  direct  proteins  and  organelles  for  degradation  [405].  Ref(2)P  is  involved  in  promoting   protein   aggregation   and   the   capacity   of   Ref(2)P   to   bind   ubiquitin   and  Atg8   offers   a   molecular   link   between   ubiquitinated   substrates   and   the   autophagy  pathway   [354,356,406,407].  As  ubiquitin  and  Ref(2)P  can  deliver  substrates   to   the  autophagic  machinery  we  asked  if  these  two  molecules  are  involved  in  the  selective  recognition  of  hproIAPP  and  hIAPP.   Indeed,   frequent  ubiquitination  of   intracellular  substrates   was   detected   in   cells   expressing   hproIAPP   and   hIAPP.   No   such  ubiquitination   pattern  was   present   in   cells   producing   non-­‐aggregating  mIAPP.  We  were   able   to   identify  Ref(2)P   to  be   in   involved   in   recognition  of  ubiquitin  positive  accumulations   and   targeting   them   for   autophagic   degradation.   This   conclusion   is  based   on   the   finding   that   a   decrease   in   autophagic   capacity,   mediated   by   RNAi  against  Atg8a  and  Atg8b,  leads  to  accumulation  of  intracellular  vesicles  that  stain  for  ubiquitin  and  Ref(2)P.  Such  vesicles  were  found  in  all  cells  with  decreased  Atg8a  and  Atg8b  levels,  however  hIAPP  and  hproIAPP  expression  gave  rise  to  a  new  population  of   such   vesicles   that   were   unique   with   their   increased   surface   area.   The   smaller,  common  vesicles  probably  represent  normal  cellular  substrates  that  are  recognized  by  Ref(2)P   and   turned   over   by   autophagy.  We   concluded   that   the   new  population  consisting   of   enlarged   vesicles   is   related   to   accumulation   of   hIAPP   and   hproIAPP.  Lack   of   RNAi   against   Atg8a   and   Atg8b   was   enough   to   completely   abolish   the  presence   of   these   large   ubiquitin   and   Ref(2)P   containing   vesicles.   This   further  underlines  the  physiological  role  of  autophagy  in  degrading  these  vesicles.  We  have  shown  for  hIAPP  and  hproIAPP  that  autophagosomes  were  still  capable  to  fuse  with  lysosomes.  However,  autophagic  capacity  was  not  sufficient  to  protect  cells  from   accelerated   cell   death.   For  Aβ   and  Alzheimer’s   disease   it   has   been   suggested  that   neurodegeneration   is   related   to   defective   autophagosome-­‐lysosme   fusion,  which  can  be  observed  in  dystrophic  dendrites  and  axons  [408].  This  points  towards  separate  toxicity  mechanisms  for  Aβ  and  hIAPP/hproIAPP.  Expression  of  Aβ42  in  the  whole   CNS   shortens   life   span.   This   reduction   in   survival   is   even  more   prominent  when  Aβ42  contains  the  E22G  mutation.  However,  LNvs  that  produce  Aβ42,  with  and  without  E22G  mutation,  are  not  influenced  in  their  viability.  This  demonstrates  that  Aβ42  can  successfully  exit  the  cells  before  exerting  toxicity  and  further  support  our  findings   that   the   identity   of   the   amyloid   protein,   and   not   the   sole   propensity   to  aggregate,  determines  the  initiation  of  cell  death.  It  remains  to  be  elucidated  which  further  characteristics  of  the  amyloid  protein  are  important  in  cell  death  initiation.    Last   but   no   least,   our   finding   of   extracellular   ubiquitin   and   Ref(2)P   aggregates   at  sites  where  we  expected  missing  LNvs  resembles  a  phenotype  recently  described  in  the  pancreas  of  HIP  rats.  In  these  rats  deposits  contained  p62  and  ubiquitin  and  are  suspected  to  be  connected  to  beta  cell  death  [384].  This  phenotypic  mimicry  stresses  the  physiological  relevance  of  findings  from  our  Drosophila  model.  

 

 

 

     

 

 

 

                           

General  Discussion    

and    

Future  Perspectives        

General  Discussion  and  Future  Perspectives      

 

 70  

“But  you  know,   flies  are  not  humans!?”   is  a   reaction   I   faced  several   times  over   the  last  years  when  describing  my  research.  And  yes,  without  doubt  there  are  numerous  differences   between   fruit   flies   and   humans.   But   does   that   necessarily   mean   that  nature  needs   to   find  novel  solutions   for  common  problems  during  evolution?  Most  certainly  not!  As  matter  of  fact,  without  research  in  flies  we  would  not  know  as  much  about   human  biology   as  we  do   today.  As   history   tends   to   repeat   itself,   there   is   no  reason  to  doubt  that  research  conducted  in  Drosophila  melanogaster  will  continue  to  provide  us  with  fundamental  insights  in  human  biology.      So  what  can  we  learn  from  our  Drosophila  model?    It  is  well  described  in  the  literature  that  IAPP  is  one  of  most  amyloidogenic  proteins  in   vitro.   In   an   aqueous   in   vitro   environment,   synthetic   and   recombinant   IAPP  spontaneously  forms  amyloid-­‐like  fibrils.  One  would  predict  that  over-­‐expression  of  IAPP   in   the  brain  of  Drosophila  would  result   in   large  amounts  of  amyloid  deposits,  evoking   a   strong   phenotype.   However,   the   amounts   of   IAPP   deposition   and  associated   phenotype   was   less   pronounced   than   expected.   Reported   phenotypes,  such   as   shortened   survival   when   expressing   proIAPP   in   the   whole   brain,   protein  deposition  in  the  fat  body  and  loss  of  pdf-­‐neurons  that  expressed  hIAPP  or  hproIAPP  were   stable   and   reproducible,   but   not   as   distinct   as   the   results   from   in   vitro  experiments  suggest.    This   inconsistency   between   in   vitro   and   in   vivo   properties   also   exists   when  comparing  results  from  in  vitro  experiments  with  the  human  situation.  Patients  with  type  2  diabetes  develop   islet   amyloid  with   involved  beta   cell   reduction,  where   the  vast  majority  of  individuals  that  do  not  suffer  from  the  disease  produce  proIAPP  and  IAPP   every   day   without   deposition   of   islet   amyloid.   As   type   2   diabetes   is   closely  related  to  age,  it  is  very  likely  that  the  involved  amyloid  phenotype  also  occurs  later  in  life,  further  underscoring  the  fact  that  there  must  be  cellular  mechanisms  to  either  preclude   IAPP   aggregation   and/or   to   degrade   IAPP   aggregates.   In   order   to   fully  comprehend   the   impact   of   amyloid   formation   it   is   crucial   to   identify  mechanisms  that  are  affected  when  IAPP  aggregates,  and  to   learn  more  about   the  chronological  order  regarding  amyloid  formation  and  cellular  responses.  Drosophila  research  with  its  genetic  toolbox  offers  unequalled  opportunities  to  address  such  questions  in  vivo.      We   were   able   to   identify   autophagy   to   be   selectively   triggered   by   hIAPP   and  hproIAPP   over-­‐expression   and   our   results   demonstrate   that   this   activation  neutralizes  the  protective  effects  of  autophagy.  It  has  previously  been  shown  in  mice  that  impaired  autophagy  in  beta  cells  provokes  phenotypes  resembling  those  of  type  2  diabetes   [382,383].  First  experiments   in  hIAPP  transgenic  rats  confirm  a  role   for  IAPP   in   affecting   the   autophagy   pathway   [384].   These   results   emphasize   the  physiological  relevance  of  our  findings.  It  remains  to  be  answered  how  autophagy  is  exactly   triggered   by   hIAPP   and   hproIAPP.   We   took   advantage   of   the   non-­‐

General  Discussion  and  Future  Perspectives    

 

 

  71  

amyloidogenic  properties  of  mIAPP  and  compared  its  impact  with  those  of  hproIAPP  and  hIAPP.  We  managed  to  point  out  several  molecules,  such  as  ubiquitin  and  p62,  which   are   involved   in   the   selective   recognition   of   IAPP   aggregates.   IAPP   is   a  secretory   protein   and   it   is   of   great   interest   to   understand   how   cytosol   resident  ubiquitin   and   p62   can   recognize   protein   misfolding   that   occurs   in   the   secretory  pathway.  Can  ubiquitin  and  p62  target  complete  secretory  granules  for  degradation,  and  if  so,  what  is  the  initial  signal  for  such  event?  Or  does  protein  misfolding  in  the  secretory   pathway   lead   to   membrane   disruption   of   e.g.   ER,   Golgi   or   secretory  granules,   followed   by   leakage   of   aggregates   to   the   cytosol   and   subsequent  recognition   by   selective   autophagy?   Jiminez   et   al.   recently   presented   a   poster  demonstrating   selective   recruitment   of   the   autophagic  machinery   upon   damage   of  trafficking  organelles.  Damage  of  the  Golgi  apparatus,  early  and  recycling  endosomes  all   resulted   in  ubiquitin,  p62/NBR1  and  LC3   recruitment   followed  by  up-­‐regulated  autophagy.  As  for  hIAPP,  Engel  et  al.  described  a  model  in  which  hIAPP  is  cytotoxic  by  its  capacity  to  bind  membranes  resulting  in  fibril  growth  and  significant  changes  of  membrane  curvature,  which   finally   leads   to  physical  breakage  of   the  membrane  [63].   Our   results   from   paper   I   reveal   that   neither   mature   fibrils   nor   solubilised  proteins  were  capable   to   induce  apoptosis.  However,  caspase-­‐3  was  activated  once  we  accelerated  fibril  formation  by  addition  of  small  amounts  of  preformed  fibrils  to  solubilised   monomers   of   hproIAPP   or   any   of   its   metabolites.   This   finding  substantiates   that   the   process   of   fibrillization   is   important   in   toxicity   and   is   in  accordance  with  the  model  suggested  by  Engel  et  al.  [63].  From  these  results,  a  new  model  emerges  in  which  hIAPP  and  hproIAPP  can  form  intracellular  aggregates  that  disrupt   the   surrounding   organelle   membrane.   This   starts   a   reaction   cascade   that  involves  activation  of  selective  autophagy   in  order  to  degrade  the  hIAPP/hproIAPP  aggregate  and/or  the  dysfunctional  organelle.  Important  roles  for  autophagy  include  general   cytosol   turnover,   but   also   the   protection   of   cells   from   accumulation   of  damaged  organelles.  Our  experiments  clearly  showed  an  up-­‐regulation  of  autophagy  upon   hIAPP   and   hproIAPP   expression   and   that   extensive   autophagy   activation  hampers  its  normal  protective  effects.    Insulin  resistance   leading  to  elevated  insulin  demand  and  simultaneous   increase   in  IAPP   production   precedes   type   2   diabetes.   This   intra-­‐organelle   concentration-­‐increase   could   cause   intracellular   aggregation,   organelle   damage,   and   provoke  autophagic   breakdown.   In   short,   this   can   protect   cells   from   intracellular   protein  deposition,  however  if  this  state  becomes  chronic  the  autophagy  machinery  can  get  out  of  balance  and  not  be  sufficient  to  maintain  cellular  homeostasis,  and  eventually  lead  to  cell  death.  Thus  initiating  a  chain  reaction  where  IAPP  aggregates  from  dying  beta   cells   are   released   to   the   extracellular   space.  There,   preformed  aggregates   can  act  as  seed  for  IAPP  secreted  from  surrounding  beta  cells,  which  will  lead  to  growth  of  islet  amyloid  deposits,  thereby  causing  beta  cell  death  and  result  in  an  incapacity  to  produce  sufficient  amounts  of  insulin.  In  this  way  hIAPP  aggregation  would  play  a  major   role   in   causing   type   2   diabetes.   Such   a  model   leads   to   questions   about   the  current  treatment  of  type  2  diabetes  by  administrating  drugs,  e.g.  sulfonylurea  that  

General  Discussion  and  Future  Perspectives      

 

 72  

increase   insulin   production   in   order   to   meet   elevated   insulin   demand.   Instead,  insulin   treatment   at   an   early   stage   of   the   disease   together   with   arrangements   to  decrease  insulin  resistance,  e.g.  weight  reduction,  and  increased  amounts  of  exercise,  should  be  considered  as  better  option   to   reduce  beta  cell   stress  and   to  postpone  a  complete  dependence  for  patients  on  external  insulin  administration.    If  extensive  expression  of  hIAPP  and  hproIAPP  is  cytotoxic,  why  did  only  hproIAPP  but  not  hIAPP  shorten  the  survival  of  Drosophila  when  expressed  in  the  whole  brain?  I  think  the  answer  can  be  found  in  the  nature  of  the  chosen  Gal4  driver  responsible  for  ectopic  pan-­‐neuronal  expression.  Gal4  production  decreases  rapidly  with  age  and  this  decrease  of  hIAPP  and  hproIAPP  production  may  be  sufficient  to  avoid  a  collapse  of   the   autophagy   machinery.   In   this   way   neither   hproIAPP   nor   hIAPP   will   exert  intracellular   toxicity.   The   observed   shortening   of   lifespan   is   probably   due   to  extracellular   events.   Such   events   could   include   hproIAPP   to   be   secreted   in   a   form  that   allowed   further   aggregation   on   surrounding   cell   membranes   and   thereby  causing   cell   death.   It   also   is   possible   that   hproIAPP   interacts   with   surrounding  neurons  differently  than  hIAPP,  thus  allowing  for  longer  retention  of  hproIAPP  in  the  brain  before  being  cleared  to  the  fat  body.      In  the  past,  attempts  have  been  made  to  identify  general  toxicity  mechanisms  for  all  amyloid   proteins.   Comparing   the   outcomes   of   Aβ   expression   with   hIAPP   and  hproIAPP  production,  we  were  able  to  demonstrate  that  Aβ   toxicity   is  not   initiated  intracellular.   Cell   death   was   triggered   before   the   protein   was   secreted   in   cells  producing   hIAPP   or   hproIAPP.   It   will   be   interesting   to   see   if   Aβ   still   activates  autophagy  in  the  same  way  as  hIAPP  and  hproIAPP  do.  This  will  give  us  clues  about  the   relation   of   autophagy   initiation   and   cell   death.   Since  Aβ,   as  well   as   hIAPP   and  hproIAPP,  can  form  amyloid,  there  clearly  have  to  be  additional  factors  that  influence  toxicity.  Such  factors  might  include  the  property  to  interact  with  membranes  or  the  speed  of  aggregation.  Both  these  factors  can  affect  the  site  of  initial  fibril  formation  and   fibril   growth   on  membranes.   It   is   conceivable   that   Aβ   does   not   interact   with  membranes  of  organelles  of  the  secretory  pathway  but  with  the  extracellular  part  of  the   cell   membrane.   In   this   case   could   toxicity   be   initiated   by   the   extracellular  presence  of  Aβ.      Future   research   will   further   dissect   the   molecular   pathways   affected   by   different  amyloid   proteins   and   clarify   if   amyloid   deposits   are   a   cause   or   consequence   to   its  associated  disease.  This  remains  the  crucial  question  to  answer  in  the  light  of  disease  treatment.            

 

 

 

                                     

Acknowledgements  

Acknowledgements      

 

 74  

I  want  to  express  my  sincere  gratitude  to  everybody  who  contributed  to  this  thesis  in  one  way  or  another.  In  particular  I  want  to  thank  to:    Gunilla  Westermark,  my   supervisor.   Firstly,   thank   you   for   introducing  me   to   the  field  and  giving  me  the  opportunity  to  join  your  team.  During  my  time  in  your  lab  I  got   the  chance   to   learn  a   lot  and   I  am  grateful   for  all   the  help  you  provided  to  me.  Your  hard-­‐working  attitude  and  drive  to  fight  a  little  bit  harder  combined  with  great  generosity  was  a  crucial  support  for  my  thesis.    Mattias  Alenius,  after  one  evening  of  discussing  research  you  soon  became  not  only  a  mentor   for  me,  but   also  a  dear   friend.   I   am  glad   that  you  were  always  willing   to  share  you  immense  knowledge  of  research   in  general,  and  Drosophila   in  particular,  with  me.  Many  ideas  in  this  thesis  are  a  result  of  the  discussions  with  you  over  the  last  years.      Tor  Erik  Rusten,  for  your  help  in  discussing  results  and  designing  good  experiments  to  unravel  some  of  the  mysteries  of  autophagy.  I  really  hope  this  thesis  doesn’t  put  an  end  to  our  co-­‐operation.    Peter  Nilsson,  for  our  your  collaboration  and  sharing  your  knowledge  on  LCPs.    Stefan   Klintström,   for   your   unlimited   commitment   to   Forum   Scientium.   It   was   a  privilege   to   be   part   of   Forum   Scientium   and   to   meet   great   PhD   students   from  completely  different  fields.      Per  Westermark,  you  had  potentially  the  biggest  influence  on  the  title  of  my  thesis,  as  you  were  the  person  who  named  IAPP.  It  was  always  a  great  pleasure  to  meet  you  and  to  take  part  of  your  vast  experience  on  amyloid  research.    Stefan  Thor  and  his  research  group,  for  helping  us  getting  started  with  our  own  fly  research  and  later  helping  us  to  find  solutions  to  get  fly  flood.    Xiahong   Gu,   for   joining   our   group   and   showing   so  much   interest   in  Drosophila.   I  really  enjoyed  working  together  and  wish  you  all   the  best  with   the  continuation  of  this  project!  I  can’t  wait  to  read  your  thesis  one  day!    Johan  Paulsson,  for  taking  me  under  your  wing  when  I  started  in  the  group.  I  often  think  back  to  our  work-­‐intensive  but  great  days  and  two  fantastic  conferences  with  you  as  a  roommate.  From  you   I   learned  that  4am  might  not  be   the  best   time  to  do  Victor-­‐measurements,  but  also  to  make  sure  not  to  burn  my  forehead  skin….    Marie  Oskarsson,   for  being  such  a  wonderful  co-­‐worker  and  always  being  able   to  make  me  smile.  You   left  a  huge  vacuum  behind  you  when  you  moved  to  Uppsala.   I  wish  you  all  the  best  for  your  future!  

Acknowledgements    

 

 

  75  

Sofia  Nyström,   for  great  discussions.  Maybe  not  all  of  our  discussions  were  worth  the  time  and  energy  we  put  in  to  them,  but  I  wouldn’t  want  to  miss  any  one  of  them.  I  miss  being  able  to  turn  around  and  get  good  advice  from  you.    Jana   Sponarova,   for  a  great   time   together   in  our  group.  Even   though   I   still  prefer  German  to  Czech  beer,  I  am  glad  for  all  your  effort  to  teach  us  more  about  the  Czech  Republic.  Prague  still  is  very  high  on  my  list  and  hope  for  your  help  the  day  I  finally  make  it  there…    Marie-­‐Louise  Eskilsson,  for  your  immense  patience  when  sectioning  fly  heads.  It  is  easy   to   say   what   you   want   to   have,   but   sometimes   extremely   difficult   to   make   it  happen.  You  always  succeeded…    Aida,  for  being  so  extremely  kind  and  always  willing  to  help.  All  the  small  things  you  helped  with  when  I  was  not  in  the  lab  made  a  huge  difference  for  this  work!    Mildred,  for  contributing  to  the  research  presented  in  this  thesis.  It  was  fun  working  with  you.    Bengt-­‐Arne  Fredriksson,  for  your  help  with  the  microscope.    I  also  want  to   thank  all   the  other  PhD  students/postdocs  that  were  not  part  of  our  group  but  still  made  a  difference  for  me:    Tobias,   it   was   a   great   privilege   for   me   to   sit   next   to   a   person   I   had   so   much   in  common   with   –   starting   from   being   the   only   PhD   student   in   your   group   left   in  Linköping,   to  having  realized  the  huge  advantages  of  Mac  over  Windows  J.   I   think  after  so  many  years  in  the  same  office,  water-­‐skiing  at  your  summerhouse  and  many  parties  we  maybe  should  go  for  a  camping  trip…  Åsa,  Siri,  and  Cissi,  unfortunately  not  all   three  of  you  were  around  during   the   last  year.  But   I  will  miss  being  able   to  come   to   your   office   and   discuss   science   or   any   of   the   other   important   issues   that  matter   in   life.   A   coffee   break  without   you  doesn’t   feel   complete.  Without   you  Åsa,  there  would   be   so  much   good  music   I   wouldn´t   have   had   a   clue   exists.  Cissi,   you  helped  me   to  understand   that  Germany  does  not  export  as  much  music  as  Sweden  does,  and  Siri,   I  can’t  tell  you  how  much  I  appreciate  your  sense  of  humour  –  even  though  it  was  not  always  an  ego  boost  J.  Ia,  I  always  was  fascinated  by  the  energy  you  radiate.      Gosia,  for  all  our  conversations  on  Drosophila  and  life  in  Sweden.  You  taught  me  a  lot  and  I  keep  my  fingers  crossed  that  everything  works  out  the  way  you  wish  for!    Björn,  for  great  training  company.  Go  sub22!  We  both  know  that  you  can  do  it!    Anita  for  sharing  your  wisdom  with  me.  Good  luck  in  Freiburg!  

Acknowledgements      

 

 76  

Peter   Strålfors,   Sven   Hammarström,  Mats   Söderström,   and   all   other  members  from  level  12  (former  and  actual)  for  creating  a  nice  working  atmosphere.    Anders,  Calle,  Jocke,  Johan  J,  Johan  O,  Jonas,  Lisa,  Patiyan,  Ryan,  Ullis,  I  probably  could   fill   a   separate   book   listening   all   the   great   adventures,   evenings,   parties,  conferences,  and  experiments  (I  never  looked  at  15ml  the  same  way  again)  over  the  last  years.  And  I  think  such  a  book  couldn’t  come  close  to  describe  how  grateful  I  am  for  having  been  a  PhD  student  the  same  time  you  were  in  Linköping.    Maria,  Pelle,  Magdalena,   Jason,   Jenny,   and  Mattias,   thank  you  so  much  for  your  friendship.   I  really  hope  we  manage  to   increase  the  number  of  evenings  discussing  the  big  questions  of   life!  Pelle  and  Maria,   it   is  great   to  get   those  phone  calls  when  you  try  to  get  me  out  climbing  or  invite  me  over  for  dinner.  I  am  looking  forward  to  be   able   to   spend  more   time  with   you   again.  Magda   and   Jason,   for   all   those   great  vacations,   from   ski   mountaineering   in   Canada   to   paddling   kayak   in   the   Swedish  archipelago.  Even  though  there  are  thousands  of  kilometres  between  us  now,  so  are  you  always  close  to  me!  Jenny  and  Mattias,  for  just  being  there  –  no  matter  when!      Pernilla,   for  all   the  time  we  shared  over  the   last  years.   I  am  deeply  grateful   for  all  your  support  and  wish  you  all  the  best  for  your  future!      Christine,  Ulrike,  and  Katharina,   for  being  the  best  sisters  a  younger  brother  can  ask  for  (at  least  after  you  all  had  left  the  house  J).  Katharina,  you  made  sure  to  fill  in  for  our  parents  when  you  thought  life  was  getting  a  bit  to  easy  for  me  J.  But  you  also   always   took,   and   still   take,   time   to   listen   to  my  problems.  Ulrike,   you   always  were,  and  still  are,  my  role-­‐model  of  how  to  interact  with  other  people  and  you  are  an  amazing  mentor.  Christine,  you  had  a  huge  impact  on  me,  and  still  have.  To  know  that  all  of  you  always  are  there  for  me  and  support  me  whenever  it   is  needed  is  an  invaluable  secureness  that  often  allowed  me  to  dare  just  a  bit  more.        Meinen   Eltern  möchte   ganz  herzlich   für   all   die  Unterstützung  danken.  Ohne  Euch  und   all   die   Werte   die   Ihr   mir   versucht   habt   zu   übermitteln,   hätte   ich   all   das   nie  gewagt  und  erst  recht  nicht  geschafft!      Hanna,   for   all   your   support   over   the   last  months.   All   the   laughs,   adventures,   and  experiences  we  shared  during  this  time  gave  me  all  the  energy  needed  to  finish  my  work.   And   now   I   can’t   wait   to   discover   the   world   together   with   you   –   and   who  knows,  one  day,  after  a  great  climb  we  might  stand  on  top  of  “Presten”…  

 

                                     

References    

References      

 

 78  

1. Anfinsen CB (1973) Principles that govern the folding of protein chains. Science 181: 223-230.

2. Levinthal C (1969) How to fold graciously. Mossbauer Spectroscopy in Biological Systems, Proceedings of a Meeting held at Allerton House, Monticello, Illinois: 22.

3. Karplus M (1997) The Levinthal paradox: yesterday and today. Folding & design 2: S69-75.

4. Dobson CM (2003) Protein folding and misfolding. Nature 426: 884-890. 5. Ellis RJ, Minton AP (2003) Cell biology: join the crowd. Nature 425: 27-28. 6. Gething MJ, Sambrook J (1992) Protein folding in the cell. Nature 355: 33-

45. 7. Richter K, Haslbeck M, Buchner J (2010) The heat shock response: life on

the verge of death. Molecular cell 40: 253-266. 8. Virchow R (1854) Zur Cellulose-Frage. Virchows Archiv A, Pathological

Anatomy and Histopathology 6: 416-426. 9. Aterman K (1976) A historical note on the iodine-sulphuric acid reaction of

amyloid. Histochemistry 49: 131-143. 10. Friedreich N, Kekulé A (1859) Zur Amyloidfrage. Virchows Archiv A,

Pathological Anatomy and Histopathology 16. 11. Bennhold H (1922) Eine spezifische Amyloidfärbung mit Kongorot. Münch

Med Wochenschr 69: 1537-1538. 12. Divry P, Florkin M (1927) Sur les propriétées optique de l'amyloide. CR Soc

Biol (Paris) 97: 1808-1810. 13. Puchtler H, Sweat F, Levine M (1962) On the binding of Congo red by

amyloid. J Histochem Cytochem 10: 355-364. 14. Westermark P (2005) Part I: Overview of Amyloidosis and Amyloid Proteins.

Amyloid proteins The beta sheet confirmation and disease, ed J D Sipe. 15. Cohen AS, Calkins E (1959) Electron microscopic observations on a fibrous

component in amyloid of diverse origins. Nature 183: 1202-1203. 16. Shirahama T, Cohen AS (1976) A brief review of the ultrastructure of

amyloid. Amyloidosis, eds J Marrink and M H van Rijswijk: 51-57. 17. Eanes ED, Glenner GG (1968) X-ray diffraction studies on amyloid

filaments. The journal of histochemistry and cytochemistry : official journal of the Histochemistry Society 16: 673-677.

18. Westermark P, Benson MD, Buxbaum JN, Cohen AS, Frangione B, et al. (2007) A primer of amyloid nomenclature. Amyloid 14: 179-183.

19. Sipe JD, Benson MD, Buxbaum JN, Ikeda S, Merlini G, et al. (2010) Amyloid fibril protein nomenclature: 2010 recommendations from the nomenclature committee of the International Society of Amyloidosis. Amyloid : the international journal of experimental and clinical investigation : the official journal of the International Society of Amyloidosis 17: 101-104.

20. Kelly JW (1997) Amyloid fibril formation and protein misassembly: a structural quest for insights into amyloid and prion diseases. Structure 5: 595-600.

21. Lansbury PT, Jr. (1992) In pursuit of the molecular structure of amyloid plaque: new technology provides unexpected and critical information. Biochemistry 31: 6865-6870.

22. Sunde M, Blake C (1997) The structure of amyloid fibrils by electron microscopy and X-ray diffraction. Advances in protein chemistry 50: 123-159.

23. Serpell LC, Sunde M, Benson MD, Tennent GA, Pepys MB, et al. (2000) The protofilament substructure of amyloid fibrils. Journal of molecular biology 300: 1033-1039.

24. Makin OS, Serpell LC (2005) Structures for amyloid fibrils. The FEBS journal 272: 5950-5961.

References    

 

 

  79  

25. Fandrich M, Meinhardt J, Grigorieff N (2009) Structural polymorphism of Alzheimer Abeta and other amyloid fibrils. Prion 3: 89-93.

26. Greenwald J, Riek R (2010) Biology of amyloid: structure, function, and regulation. Structure 18: 1244-1260.

27. Wang L, Schubert D, Sawaya MR, Eisenberg D, Riek R (2010) Multidimensional structure-activity relationship of a protein in its aggregated states. Angewandte Chemie 49: 3904-3908.

28. Leveugle B, Ding W, Durkin JT, Mistretta S, Eisle J, et al. (1997) Heparin promotes beta-secretase cleavage of the Alzheimer's amyloid precursor protein. Neurochemistry international 30: 543-548.

29. Scholefield Z, Yates EA, Wayne G, Amour A, McDowell W, et al. (2003) Heparan sulfate regulates amyloid precursor protein processing by BACE1, the Alzheimer's beta-secretase. The Journal of cell biology 163: 97-107.

30. Li JP, Galvis ML, Gong F, Zhang X, Zcharia E, et al. (2005) In vivo fragmentation of heparan sulfate by heparanase overexpression renders mice resistant to amyloid protein A amyloidosis. Proceedings of the National Academy of Sciences of the United States of America 102: 6473-6477.

31. Zhang X, Li JP (2010) Heparan sulfate proteoglycans in amyloidosis. Progress in molecular biology and translational science 93: 309-334.

32. Kisilevsky R, Szarek WA (2002) Novel glycosaminoglycan precursors as anti-amyloid agents part II. Journal of molecular neuroscience : MN 19: 45-50.

33. Hirschfield GM, Hawkins PN (2003) Amyloidosis: new strategies for treatment. The international journal of biochemistry & cell biology 35: 1608-1613.

34. Ariga T, Miyatake T, Yu RK (2010) Role of proteoglycans and glycosaminoglycans in the pathogenesis of Alzheimer's disease and related disorders: amyloidogenesis and therapeutic strategies--a review. Journal of neuroscience research 88: 2303-2315.

35. Pepys MB, Dyck RF, de Beer FC, Skinner M, Cohen AS (1979) Binding of serum amyloid P-component (SAP) by amyloid fibrils. Clinical and experimental immunology 38: 284-293.

36. Tennent GA, Lovat LB, Pepys MB (1995) Serum amyloid P component prevents proteolysis of the amyloid fibrils of Alzheimer disease and systemic amyloidosis. Proceedings of the National Academy of Sciences of the United States of America 92: 4299-4303.

37. Hawkins PN (2002) Serum amyloid P component scintigraphy for diagnosis and monitoring amyloidosis. Current opinion in nephrology and hypertension 11: 649-655.

38. Charge SB, Esiri MM, Bethune CA, Hansen BC, Clark A (1996) Apolipoprotein E is associated with islet amyloid and other amyloidoses: implications for Alzheimer's disease. The Journal of pathology 179: 443-447.

39. Verghese PB, Castellano JM, Holtzman DM (2011) Apolipoprotein E in Alzheimer's disease and other neurological disorders. Lancet neurology 10: 241-252.

40. Stefani M, Dobson CM (2003) Protein aggregation and aggregate toxicity: new insights into protein folding, misfolding diseases and biological evolution. Journal of molecular medicine 81: 678-699.

41. Chiti F, Dobson CM (2009) Amyloid formation by globular proteins under native conditions. Nature chemical biology 5: 15-22.

42. Jarrett JT, Lansbury PT, Jr. (1992) Amyloid fibril formation requires a chemically discriminating nucleation event: studies of an amyloidogenic sequence from the bacterial protein OsmB. Biochemistry 31: 12345-12352.

43. Jarrett JT, Lansbury PT, Jr. (1993) Seeding "one-dimensional crystallization" of amyloid: a pathogenic mechanism in Alzheimer's disease and scrapie? Cell 73: 1055-1058.

References      

 

 80  

44. Kelly JW (1998) The alternative conformations of amyloidogenic proteins and their multi-step assembly pathways. Current opinion in structural biology 8: 101-106.

45. Rochet JC, Lansbury PT, Jr. (2000) Amyloid fibrillogenesis: themes and variations. Current opinion in structural biology 10: 60-68.

46. Lundmark K, Westermark GT, Olsen A, Westermark P (2005) Protein fibrils

in nature can enhance amyloid protein A amyloidosis in mice: Cross-seeding as a disease mechanism. Proceedings of the National Academy of Sciences of the United States of America 102: 6098-6102.

47. Ranlov P (1967) The adoptive transfer of experimental mouse amyloidosis by intravenous injections of spleen cell extracts from casein-treated syngeneic donor mice. Acta pathologica et microbiologica Scandinavica 70: 321-335.

48. Ridley RM, Baker HF, Windle CP, Cummings RM (2006) Very long term studies of the seeding of beta-amyloidosis in primates. Journal of neural transmission 113: 1243-1251.

49. Westermark GT, Westermark P (2010) Prion-like aggregates: infectious agents in human disease. Trends in molecular medicine 16: 501-507.

50. Stefani M (2010) Biochemical and biophysical features of both oligomer/fibril and cell membrane in amyloid cytotoxicity. The FEBS journal 277: 4602-4613.

51. Glabe CG (2006) Common mechanisms of amyloid oligomer pathogenesis in degenerative disease. Neurobiology of aging 27: 570-575.

52. Glabe CG, Kayed R (2006) Common structure and toxic function of amyloid oligomers implies a common mechanism of pathogenesis. Neurology 66: S74-78.

53. Quist A, Doudevski I, Lin H, Azimova R, Ng D, et al. (2005) Amyloid ion channels: a common structural link for protein-misfolding disease. Proceedings of the National Academy of Sciences of the United States of America 102: 10427-10432.

54. Kayed R, Head E, Thompson JL, McIntire TM, Milton SC, et al. (2003) Common structure of soluble amyloid oligomers implies common mechanism of pathogenesis. Science 300: 486-489.

55. Mirzabekov TA, Lin MC, Kagan BL (1996) Pore formation by the cytotoxic islet amyloid peptide amylin. The Journal of biological chemistry 271: 1988-1992.

56. Cohen E, Bieschke J, Perciavalle RM, Kelly JW, Dillin A (2006) Opposing activities protect against age-onset proteotoxicity. Science 313: 1604-1610.

57. Zhang S, Liu H, Yu H, Cooper GJ (2008) Fas-associated death receptor signaling evoked by human amylin in islet beta-cells. Diabetes 57: 348-356.

58. Zhang S, Liu J, Dragunow M, Cooper GJ (2003) Fibrillogenic amylin evokes islet beta-cell apoptosis through linked activation of a caspase cascade and JNK1. The Journal of biological chemistry 278: 52810-52819.

59. Tomiyama T, Kaneko H, Kataoka K, Asano S, Endo N (1997) Rifampicin inhibits the toxicity of pre-aggregated amyloid peptides by binding to peptide fibrils and preventing amyloid-cell interaction. The Biochemical journal 322 ( Pt 3): 859-865.

60. Bhattacharya S, Latha JN, Kumresan R, Singh S (2007) Cloning and expression of human islet amyloid polypeptide in cultured cells. Biochemical and biophysical research communications 356: 622-628.

61. Jurgens CA, Toukatly MN, Fligner CL, Udayasankar J, Subramanian SL, et al. (2011) beta-cell loss and beta-cell apoptosis in human type 2 diabetes are related to islet amyloid deposition. The American journal of pathology 178: 2632-2640.

62. Zraika S, Hull RL, Verchere CB, Clark A, Potter KJ, et al. (2010) Toxic oligomers and islet beta cell death: guilty by association or convicted by circumstantial evidence? Diabetologia 53: 1046-1056.

References    

 

 

  81  

63. Engel MF, Khemtemourian L, Kleijer CC, Meeldijk HJ, Jacobs J, et al. (2008) Membrane damage by human islet amyloid polypeptide through fibril growth at the membrane. Proc Natl Acad Sci U S A 105: 6033-6038.

64. Knight JD, Miranker AD (2004) Phospholipid catalysis of diabetic amyloid assembly. Journal of molecular biology 341: 1175-1187.

65. Aisenbrey C, Borowik T, Bystrom R, Bokvist M, Lindstrom F, et al. (2008) How is protein aggregation in amyloidogenic diseases modulated by biological membranes? European biophysics journal : EBJ 37: 247-255.

66. Engel MF, Visser AJ, van Mierlo CP (2004) Conformation and orientation of a protein folding intermediate trapped by adsorption. Proceedings of the National Academy of Sciences of the United States of America 101: 11316-11321.

67. Norde W, Lyklema J (1991) Why proteins prefer interfaces. Journal of biomaterials science Polymer edition 2: 183-202.

68. Chernoff YO (2004) Amyloidogenic domains, prions and structural inheritance: rudiments of early life or recent acquisition? Current opinion in chemical biology 8: 665-671.

69. Fowler DM, Koulov AV, Balch WE, Kelly JW (2007) Functional amyloid--from bacteria to humans. Trends in biochemical sciences 32: 217-224.

70. Chapman MR, Robinson LS, Pinkner JS, Roth R, Heuser J, et al. (2002) Role of Escherichia coli curli operons in directing amyloid fiber formation. Science 295: 851-855.

71. Claessen D, Rink R, de Jong W, Siebring J, de Vreugd P, et al. (2003) A novel class of secreted hydrophobic proteins is involved in aerial hyphae formation in Streptomyces coelicolor by forming amyloid-like fibrils. Genes & development 17: 1714-1726.

72. True HL, Lindquist SL (2000) A yeast prion provides a mechanism for genetic variation and phenotypic diversity. Nature 407: 477-483.

73. Baxa U, Cheng N, Winkler DC, Chiu TK, Davies DR, et al. (2005) Filaments of the Ure2p prion protein have a cross-beta core structure. Journal of structural biology 150: 170-179.

74. Iconomidou VA, Vriend G, Hamodrakas SJ (2000) Amyloids protect the silkmoth oocyte and embryo. FEBS letters 479: 141-145.

75. Si K, Lindquist S, Kandel ER (2003) A neuronal isoform of the aplysia CPEB has prion-like properties. Cell 115: 879-891.

76. Fowler DM, Koulov AV, Alory-Jost C, Marks MS, Balch WE, et al. (2006) Functional amyloid formation within mammalian tissue. PLoS biology 4: e6.

77. Maji SK, Perrin MH, Sawaya MR, Jessberger S, Vadodaria K, et al. (2009) Functional amyloids as natural storage of peptide hormones in pituitary secretory granules. Science 325: 328-332.

78. Clark A, Cooper GJ, Lewis CE, Morris JF, Willis AC, et al. (1987) Islet amyloid formed from diabetes-associated peptide may be pathogenic in type-2 diabetes. Lancet 2: 231-234.

79. Opie EL (1901) The Relation Of Diabetes Mellitus to Lesions of the Pancreas. Hyaline Degeneration of the Islands Oe Langerhans. The Journal of experimental medicine 5: 527-540.

80. Westermark P, Grimelius L (1973) The pancreatic islet cells in insular amyloidosis in human diabetic and non-diabetic adults. Acta pathologica et microbiologica Scandinavica Section A, Pathology 81: 291-300.

81. Westermark P, Wernstedt C, Wilander E, Hayden DW, O'Brien TD, et al. (1987) Amyloid fibrils in human insulinoma and islets of Langerhans of the diabetic cat are derived from a neuropeptide-like protein also present in normal islet cells. Proc Natl Acad Sci U S A 84: 3881-3885.

References      

 

 82  

82. Westermark P, Wernstedt C, Wilander E, Sletten K (1986) A novel peptide in the calcitonin gene related peptide family as an amyloid fibril protein in the endocrine pancreas. Biochemical and biophysical research communications 140: 827-831.

83. Cooper GJ, Willis AC, Clark A, Turner RC, Sim RB, et al. (1987) Purification and characterization of a peptide from amyloid-rich pancreases of type 2 diabetic patients. Proceedings of the National Academy of Sciences of the United States of America 84: 8628-8632.

84. Christmanson L, Rorsman F, Stenman G, Westermark P, Betsholtz C (1990) The human islet amyloid polypeptide (IAPP) gene. Organization, chromosomal localization and functional identification of a promoter region. FEBS letters 267: 160-166.

85. Macfarlane WM, Campbell SC, Elrick LJ, Oates V, Bermano G, et al. (2000) Glucose regulates islet amyloid polypeptide gene transcription in a PDX1- and calcium-dependent manner. The Journal of biological chemistry 275: 15330-15335.

86. Macfarlane WM, Shepherd RM, Cosgrove KE, James RF, Dunne MJ, et al. (2000) Glucose modulation of insulin mRNA levels is dependent on transcription factor PDX-1 and occurs independently of changes in intracellular Ca2+. Diabetes 49: 418-423.

87. Mosselman S, Hoppener JW, Zandberg J, van Mansfeld AD, Geurts van Kessel AH, et al. (1988) Islet amyloid polypeptide: identification and chromosomal localization of the human gene. FEBS letters 239: 227-232.

88. Muff R, Born W, Lutz TA, Fischer JA (2004) Biological importance of the peptides of the calcitonin family as revealed by disruption and transfer of corresponding genes. Peptides 25: 2027-2038.

89. Wimalawansa SJ (1997) Amylin, calcitonin gene-related peptide, calcitonin, and adrenomedullin: a peptide superfamily. Critical reviews in neurobiology 11: 167-239.

90. Betsholtz C, Svensson V, Rorsman F, Engstrom U, Westermark GT, et al. (1989) Islet amyloid polypeptide (IAPP):cDNA cloning and identification of an amyloidogenic region associated with the species-specific occurrence of age-related diabetes mellitus. Experimental cell research 183: 484-493.

91. Lukinius A, Wilander E, Westermark GT, Engstrom U, Westermark P (1989) Co-localization of islet amyloid polypeptide and insulin in the B cell secretory granules of the human pancreatic islets. Diabetologia 32: 240-244.

92. Westermark P, Wilander E, Westermark GT, Johnson KH (1987) Islet amyloid polypeptide-like immunoreactivity in the islet B cells of type 2 (non-insulin-dependent) diabetic and non-diabetic individuals. Diabetologia 30: 887-892.

93. Leffert JD, Newgard CB, Okamoto H, Milburn JL, Luskey KL (1989) Rat amylin: cloning and tissue-specific expression in pancreatic islets. Proceedings of the National Academy of Sciences of the United States of America 86: 3127-3130.

94. Nishi M, Sanke T, Nagamatsu S, Bell GI, Steiner DF (1990) Islet amyloid polypeptide. A new beta cell secretory product related to islet amyloid deposits. The Journal of biological chemistry 265: 4173-4176.

95. Hutton JC (1984) Secretory granules. Experientia 40: 1091-1098. 96. Percy AJ, Trainor DA, Rittenhouse J, Phelps J, Koda JE (1996) Development

of sensitive immunoassays to detect amylin and amylin-like peptides in unextracted plasma. Clinical chemistry 42: 576-585.

97. Betsholtz C, Christmanson L, Engstrom U, Rorsman F, Jordan K, et al. (1990) Structure of cat islet amyloid polypeptide and identification of amino acid residues of potential significance for islet amyloid formation. Diabetes 39: 118-122.

References    

 

 

  83  

98. Fan L, Westermark G, Chan SJ, Steiner DF (1994) Altered gene structure and tissue expression of islet amyloid polypeptide in the chicken. Molecular endocrinology 8: 713-721.

99. Jordan K, Murtaugh MP, O'Brien TD, Westermark P, Betsholtz C, et al. (1990) Canine IAPP cDNA sequence provides important clues regarding diabetogenesis and amyloidogenesis in type 2 diabetes. Biochemical and biophysical research communications 169: 502-508.

100. Nishi M, Bell GI, Steiner DF (1990) Sequence of a cDNA encoding Syrian hamster islet amyloid polypeptide precursor. Nucleic acids research 18: 6726.

101. Nishi M, Chan SJ, Nagamatsu S, Bell GI, Steiner DF (1989) Conservation of the sequence of islet amyloid polypeptide in five mammals is consistent with its putative role as an islet hormone. Proceedings of the National Academy of Sciences of the United States of America 86: 5738-5742.

102. Westermark GT, Falkmer S, Steiner DF, Chan SJ, Engstrom U, et al. (2002) Islet amyloid polypeptide is expressed in the pancreatic islet parenchyma of the teleostean fish, Myoxocephalus (cottus) scorpius. Comparative biochemistry and physiology Part B, Biochemistry & molecular biology 133: 119-125.

103. Mulder H, Leckstrom A, Uddman R, Ekblad E, Westermark P, et al. (1995) Islet amyloid polypeptide (amylin) is expressed in sensory neurons. The Journal of neuroscience : the official journal of the Society for Neuroscience 15: 7625-7632.

104. Skofitsch G, Wimalawansa SJ, Jacobowitz DM, Gubisch W (1995) Comparative immunohistochemical distribution of amylin-like and calcitonin gene related peptide like immunoreactivity in the rat central nervous system. Canadian journal of physiology and pharmacology 73: 945-956.

105. Mulder H, Ekelund M, Ekblad E, Sundler F (1997) Islet amyloid polypeptide in the gut and pancreas: localization, ontogeny and gut motility effects. Peptides 18: 771-783.

106. Gebre-Medhin S, Mulder H, Pekny M, Westermark G, Tornell J, et al. (1998) Increased insulin secretion and glucose tolerance in mice lacking islet amyloid polypeptide (amylin). Biochemical and biophysical research communications 250: 271-277.

107. Akesson B, Panagiotidis G, Westermark P, Lundquist I (2003) Islet amyloid polypeptide inhibits glucagon release and exerts a dual action on insulin release from isolated islets. Regulatory peptides 111: 55-60.

108. Wang F, Adrian TE, Westermark GT, Ding X, Gasslander T, et al. (1999) Islet amyloid polypeptide tonally inhibits beta-, alpha-, and delta-cell secretion in isolated rat pancreatic islets. The American journal of physiology 276: E19-24.

109. Christopoulos G, Perry KJ, Morfis M, Tilakaratne N, Gao Y, et al. (1999) Multiple amylin receptors arise from receptor activity-modifying protein interaction with the calcitonin receptor gene product. Molecular pharmacology 56: 235-242.

110. McLatchie LM, Fraser NJ, Main MJ, Wise A, Brown J, et al. (1998) RAMPs regulate the transport and ligand specificity of the calcitonin-receptor-like receptor. Nature 393: 333-339.

111. Hay DL, Christopoulos G, Christopoulos A, Sexton PM (2004) Amylin receptors: molecular composition and pharmacology. Biochemical Society transactions 32: 865-867.

112. Muff R, Buhlmann N, Fischer JA, Born W (1999) An amylin receptor is revealed following co-transfection of a calcitonin receptor with receptor activity modifying proteins-1 or -3. Endocrinology 140: 2924-2927.

References      

 

 84  

113. Tilakaratne N, Christopoulos G, Zumpe ET, Foord SM, Sexton PM (2000) Amylin receptor phenotypes derived from human calcitonin receptor/RAMP coexpression exhibit pharmacological differences dependent on receptor isoform and host cell environment. The Journal of pharmacology and experimental therapeutics 294: 61-72.

114. Banks WA, Kastin AJ (1998) Differential permeability of the blood-brain barrier to two pancreatic peptides: insulin and amylin. Peptides 19: 883-889.

115. Dunican KC, Adams NM, Desilets AR (2010) The role of pramlintide for weight loss. The Annals of pharmacotherapy 44: 538-545.

116. Riddle MC, Drucker DJ (2006) Emerging therapies mimicking the effects of amylin and glucagon-like peptide 1. Diabetes care 29: 435-449.

117. (2003) Pramlintide: (AC 137, AC 0137, Symlin, Tripro-Amylin). BioDrugs : clinical immunotherapeutics, biopharmaceuticals and gene therapy 17: 73-79.

118. Dobolyi A (2009) Central amylin expression and its induction in rat dams. Journal of neurochemistry 111: 1490-1500.

119. Huang X, Yang J, Chang JK, Dun NJ (2010) Amylin suppresses acetic acid-induced visceral pain and spinal c-fos expression in the mouse. Neuroscience 165: 1429-1438.

120. Gebre-Medhin S, Mulder H, Zhang Y, Sundler F, Betsholtz C (1998) Reduced nociceptive behavior in islet amyloid polypeptide (amylin) knockout mice. Brain research Molecular brain research 63: 180-183.

121. Deems RO, Deacon RW, Young DA (1991) Amylin activates glycogen phosphorylase and inactivates glycogen synthase via a cAMP-independent mechanism. Biochemical and biophysical research communications 174: 716-720.

122. Young AA, Mott DM, Stone K, Cooper GJ (1991) Amylin activates glycogen phosphorylase in the isolated soleus muscle of the rat. FEBS letters 281: 149-151.

123. Abaffy T, Cooper GJ (2004) GSK3 involvement in amylin signaling in isolated rat soleus muscle. Peptides 25: 2119-2125.

124. Gilbey SG, Ghatei MA, Bretherton-Watt D, Zaidi M, Jones PM, et al. (1991) Islet amyloid polypeptide: production by an osteoblast cell line and possible role as a paracrine regulator of osteoclast function in man. Clinical science 81: 803-808.

125. Dacquin R, Davey RA, Laplace C, Levasseur R, Morris HA, et al. (2004) Amylin inhibits bone resorption while the calcitonin receptor controls bone formation in vivo. The Journal of cell biology 164: 509-514.

126. Sanke T, Bell GI, Sample C, Rubenstein AH, Steiner DF (1988) An islet amyloid peptide is derived from an 89-amino acid precursor by proteolytic processing. J Biol Chem 263: 17243-17246.

127. Halban PA, Irminger JC (1994) Sorting and processing of secretory proteins. The Biochemical journal 299 ( Pt 1): 1-18.

128. Marzban L, Trigo-Gonzalez G, Zhu X, Rhodes CJ, Halban PA, et al. (2004) Role of beta-cell prohormone convertase (PC)1/3 in processing of pro-islet amyloid polypeptide. Diabetes 53: 141-148.

129. Wang J, Xu J, Finnerty J, Furuta M, Steiner DF, et al. (2001) The prohormone convertase enzyme 2 (PC2) is essential for processing pro-islet amyloid polypeptide at the NH2-terminal cleavage site. Diabetes 50: 534-539.

130. Marzban L, Trigo-Gonzalez G, Verchere CB (2005) Processing of pro-islet amyloid polypeptide in the constitutive and regulated secretory pathways of beta cells. Molecular endocrinology 19: 2154-2163.

131. Marzban L, Soukhatcheva G, Verchere CB (2005) Role of carboxypeptidase E in processing of pro-islet amyloid polypeptide in {beta}-cells. Endocrinology 146: 1808-1817.

References    

 

 

  85  

132. Jutras I, Seidah NG, Reudelhuber TL, Brechler V (1997) Two activation states of the prohormone convertase PC1 in the secretory pathway. The Journal of biological chemistry 272: 15184-15188.

133. Zhou A, Mains RE (1994) Endoproteolytic processing of proopiomelanocortin and prohormone convertases 1 and 2 in neuroendocrine cells overexpressing prohormone convertases 1 or 2. The Journal of biological chemistry 269: 17440-17447.

134. Shennan KI, Taylor NA, Jermany JL, Matthews G, Docherty K (1995) Differences in pH optima and calcium requirements for maturation of the prohormone convertases PC2 and PC3 indicates different intracellular locations for these events. The Journal of biological chemistry 270: 1402-1407.

135. Muller L, Zhu X, Lindberg I (1997) Mechanism of the facilitation of PC2 maturation by 7B2: involvement in ProPC2 transport and activation but not folding. The Journal of cell biology 139: 625-638.

136. Zhu X, Lindberg I (1995) 7B2 facilitates the maturation of proPC2 in neuroendocrine cells and is required for the expression of enzymatic activity. The Journal of cell biology 129: 1641-1650.

137. Muller L, Cameron A, Fortenberry Y, Apletalina EV, Lindberg I (2000) Processing and sorting of the prohormone convertase 2 propeptide. The Journal of biological chemistry 275: 39213-39222.

138. Balasubramaniam A, Renugopalakrishnan V, Stein M, Fischer JE, Chance WT (1991) Syntheses, structures and anorectic effects of human and rat amylin. Peptides 12: 919-924.

139. Abedini A, Tracz SM, Cho JH, Raleigh DP (2006) Characterization of the heparin binding site in the N-terminus of human pro-islet amyloid polypeptide: implications for amyloid formation. Biochemistry 45: 9228-9237.

140. Meng F, Abedini A, Song B, Raleigh DP (2007) Amyloid formation by pro-islet amyloid polypeptide processing intermediates: examination of the role of protein heparan sulfate interactions and implications for islet amyloid formation in type 2 diabetes. Biochemistry 46: 12091-12099.

141. Park K, Verchere CB (2001) Identification of a heparin binding domain in the N-terminal cleavage site of pro-islet amyloid polypeptide. Implications for islet amyloid formation. The Journal of biological chemistry 276: 16611-16616.

142. Paulsson JF, Andersson A, Westermark P, Westermark GT (2006) Intracellular amyloid-like deposits contain unprocessed pro-islet amyloid polypeptide (proIAPP) in beta cells of transgenic mice overexpressing the gene for human IAPP and transplanted human islets. Diabetologia 49: 1237-1246.

143. Paulsson JF, Westermark GT (2005) Aberrant processing of human proislet amyloid polypeptide results in increased amyloid formation. Diabetes 54: 2117-2125.

144. Smeekens SP, Montag AG, Thomas G, Albiges-Rizo C, Carroll R, et al. (1992) Proinsulin processing by the subtilisin-related proprotein convertases furin, PC2, and PC3. Proceedings of the National Academy of Sciences of the United States of America 89: 8822-8826.

145. Zhu X, Orci L, Carroll R, Norrbom C, Ravazzola M, et al. (2002) Severe block in processing of proinsulin to insulin accompanied by elevation of des-64,65 proinsulin intermediates in islets of mice lacking prohormone convertase 1/3. Proceedings of the National Academy of Sciences of the United States of America 99: 10299-10304.

146. Furuta M, Carroll R, Martin S, Swift HH, Ravazzola M, et al. (1998) Incomplete processing of proinsulin to insulin accompanied by elevation of Des-31,32 proinsulin intermediates in islets of mice lacking active PC2. The Journal of biological chemistry 273: 3431-3437.

References      

 

 86  

147. Docherty K, Hutton JC (1983) Carboxypeptidase activity in the insulin secretory granule. FEBS letters 162: 137-141.

148. Leslie RD, Kolb H, Schloot NC, Buzzetti R, Mauricio D, et al. (2008) Diabetes classification: grey zones, sound and smoke: Action LADA 1. Diabetes/metabolism research and reviews 24: 511-519.

149. Cnop M, Welsh N, Jonas JC, Jorns A, Lenzen S, et al. (2005) Mechanisms of pancreatic beta-cell death in type 1 and type 2 diabetes: many differences, few similarities. Diabetes 54 Suppl 2: S97-107.

150. Clark A, Wells CA, Buley ID, Cruickshank JK, Vanhegan RI, et al. (1988) Islet amyloid, increased A-cells, reduced B-cells and exocrine fibrosis: quantitative changes in the pancreas in type 2 diabetes. Diabetes research 9: 151-159.

151. Westermark P (1972) Quantitative studies on amyloid in the islets of Langerhans. Ups J Med Sci 77: 91-94.

152. Butler AE, Janson J, Bonner-Weir S, Ritzel R, Rizza RA, et al. (2003) Beta-cell deficit and increased beta-cell apoptosis in humans with type 2 diabetes. Diabetes 52: 102-110.

153. Clark A, Charge SB, Badman MK, MacArthur DA, de Koning EJ (1996) Islet amyloid polypeptide: actions and role in the pathogenesis of diabetes. Biochem Soc Trans 24: 594-599.

154. Hoppener JW, Lips CJ (2006) Role of islet amyloid in type 2 diabetes mellitus. The international journal of biochemistry & cell biology 38: 726-736.

155. Stumvoll M, Goldstein BJ, van Haeften TW (2005) Type 2 diabetes: principles of pathogenesis and therapy. Lancet 365: 1333-1346.

156. Kahn BB (1998) Type 2 diabetes: when insulin secretion fails to compensate for insulin resistance. Cell 92: 593-596.

157. Larsson H, Ahren B (1996) Failure to adequately adapt reduced insulin sensitivity with increased insulin secretion in women with impaired glucose tolerance. Diabetologia 39: 1099-1107.

158. Rhodes CJ (2005) Type 2 diabetes-a matter of beta-cell life and death? Science 307: 380-384.

159. Westermark P (2011) Amyloid in the islets of Langerhans: Thoughts and some historical aspects. Ups J Med Sci 116: 81-89.

160. Guardado-Mendoza R, Davalli AM, Chavez AO, Hubbard GB, Dick EJ, et al. (2009) Pancreatic islet amyloidosis, beta-cell apoptosis, and alpha-cell proliferation are determinants of islet remodeling in type-2 diabetic baboons. Proceedings of the National Academy of Sciences of the United States of America 106: 13992-13997.

161. Westermark P, Andersson A, Westermark GT (2011) Islet amyloid polypeptide, islet amyloid, and diabetes mellitus. Physiological reviews 91: 795-826.

162. Westermark P, Engstrom U, Johnson KH, Westermark GT, Betsholtz C (1990) Islet amyloid polypeptide: pinpointing amino acid residues linked to amyloid fibril formation. Proceedings of the National Academy of Sciences of the United States of America 87: 5036-5040.

163. Hoppener JW, Oosterwijk C, Nieuwenhuis MG, Posthuma G, Thijssen JH, et al. (1999) Extensive islet amyloid formation is induced by development of Type II diabetes mellitus and contributes to its progression: pathogenesis of diabetes in a mouse model. Diabetologia 42: 427-434.

164. Soeller WC, Janson J, Hart SE, Parker JC, Carty MD, et al. (1998) Islet amyloid-associated diabetes in obese A(vy)/a mice expressing human islet amyloid polypeptide. Diabetes 47: 743-750.

165. Janson J, Ashley RH, Harrison D, McIntyre S, Butler PC (1999) The mechanism of islet amyloid polypeptide toxicity is membrane disruption by intermediate-sized toxic amyloid particles. Diabetes 48: 491-498.

References    

 

 

  87  

166. Lorenzo A, Razzaboni B, Weir GC, Yankner BA (1994) Pancreatic islet cell toxicity of amylin associated with type-2 diabetes mellitus. Nature 368: 756-760.

167. Engel MF (2009) Membrane permeabilization by Islet Amyloid Polypeptide. Chem Phys Lipids 160: 1-10.

168. Porte D, Jr. (1991) Banting lecture 1990. Beta-cells in type II diabetes mellitus. Diabetes 40: 166-180.

169. Sakagashira S, Sanke T, Hanabusa T, Shimomura H, Ohagi S, et al. (1996) Missense mutation of amylin gene (S20G) in Japanese NIDDM patients. Diabetes 45: 1279-1281.

170. Seino S (2001) S20G mutation of the amylin gene is associated with Type II diabetes in Japanese. Study Group of Comprehensive Analysis of Genetic Factors in Diabetes Mellitus. Diabetologia 44: 906-909.

171. Ma Z, Westermark GT, Sakagashira S, Sanke T, Gustavsson A, et al. (2001) Enhanced in vitro production of amyloid-like fibrils from mutant (S20G) islet amyloid polypeptide. Amyloid : the international journal of experimental and clinical investigation : the official journal of the International Society of Amyloidosis 8: 242-249.

172. Sakagashira S, Hiddinga HJ, Tateishi K, Sanke T, Hanabusa T, et al. (2000) S20G mutant amylin exhibits increased in vitro amyloidogenicity and increased intracellular cytotoxicity compared to wild-type amylin. The American journal of pathology 157: 2101-2109.

173. Nanga RP, Brender JR, Xu J, Hartman K, Subramanian V, et al. (2009) Three-dimensional structure and orientation of rat islet amyloid polypeptide protein in a membrane environment by solution NMR spectroscopy. Journal of the American Chemical Society 131: 8252-8261.

174. Wei L, Jiang P, Yau YH, Summer H, Shochat SG, et al. (2009) Residual structure in islet amyloid polypeptide mediates its interactions with soluble insulin. Biochemistry 48: 2368-2376.

175. Reddy AS, Wang L, Singh S, Ling YL, Buchanan L, et al. (2010) Stable and metastable states of human amylin in solution. Biophysical journal 99: 2208-2216.

176. Dupuis NF, Wu C, Shea JE, Bowers MT (2009) Human islet amyloid polypeptide monomers form ordered beta-hairpins: a possible direct amyloidogenic precursor. Journal of the American Chemical Society 131: 18283-18292.

177. Luca S, Yau WM, Leapman R, Tycko R (2007) Peptide conformation and supramolecular organization in amylin fibrils: constraints from solid-state NMR. Biochemistry 46: 13505-13522.

178. Westermark P, Li ZC, Westermark GT, Leckstrom A, Steiner DF (1996) Effects of beta cell granule components on human islet amyloid polypeptide fibril formation. FEBS letters 379: 203-206.

179. Jaikaran ET, Nilsson MR, Clark A (2004) Pancreatic beta-cell granule peptides form heteromolecular complexes which inhibit islet amyloid polypeptide fibril formation. The Biochemical journal 377: 709-716.

180. Janciauskiene S, Eriksson S, Carlemalm E, Ahren B (1997) B cell granule peptides affect human islet amyloid polypeptide (IAPP) fibril formation in vitro. Biochemical and biophysical research communications 236: 580-585.

181. Andersson A, Bohman S, Borg LA, Paulsson JF, Schultz SW, et al. (2008) Amyloid deposition in transplanted human pancreatic islets: a conceivable cause of their long-term failure. Experimental diabetes research 2008: 562985.

182. Brender JR, Lee EL, Hartman K, Wong PT, Ramamoorthy A, et al. (2011) Biphasic effects of insulin on islet amyloid polypeptide membrane disruption. Biophysical journal 100: 685-692.

References      

 

 88  

183. Stevens FJ, Kisilevsky R (2000) Immunoglobulin light chains, glycosaminoglycans, and amyloid. Cellular and molecular life sciences : CMLS 57: 441-449.

184. Furth AJ (1997) Glycated proteins in diabetes. British journal of biomedical science 54: 192-200.

185. Clodi M, Thomaseth K, Pacini G, Hermann K, Kautzky-Willer A, et al. (1998) Distribution and kinetics of amylin in humans. The American journal of physiology 274: E903-908.

186. Kapurniotu A, Bernhagen J, Greenfield N, Al-Abed Y, Teichberg S, et al. (1998) Contribution of advanced glycosylation to the amyloidogenicity of islet amyloid polypeptide. European journal of biochemistry / FEBS 251: 208-216.

187. Reaven GM, Chen YD (1988) Role of abnormal free fatty acid metabolism in the development of non-insulin-dependent diabetes mellitus. The American journal of medicine 85: 106-112.

188. Wyne KL (2003) Free fatty acids and type 2 diabetes mellitus. The American journal of medicine 115 Suppl 8A: 29S-36S.

189. Westermark GT, Gebre-Medhin S, Steiner DF, Westermark P (2000) Islet amyloid development in a mouse strain lacking endogenous islet amyloid polypeptide (IAPP) but expressing human IAPP. Molecular medicine 6: 998-1007.

190. Ma Z, Westermark GT (2002) Effects of free fatty acid on polymerization of islet amyloid polypeptide (IAPP) in vitro and on amyloid fibril formation in cultivated isolated islets of transgenic mice overexpressing human IAPP. Molecular medicine 8: 863-868.

191. Qi D, Cai K, Wang O, Li Z, Chen J, et al. (2010) Fatty acids induce amylin expression and secretion by pancreatic beta-cells. American journal of physiology Endocrinology and metabolism 298: E99-E107.

192. de Koning EJ, Hoppener JW, Verbeek JS, Oosterwijk C, van Hulst KL, et al. (1994) Human islet amyloid polypeptide accumulates at similar sites in islets of transgenic mice and humans. Diabetes 43: 640-644.

193. Narita R, Toshimori H, Nakazato M, Kuribayashi T, Toshimori T, et al. (1992) Islet amyloid polypeptide (IAPP) and pancreatic islet amyloid deposition in diabetic and non-diabetic patients. Diabetes research and clinical practice 15: 3-14.

194. Couce M, Kane LA, O'Brien TD, Charlesworth J, Soeller W, et al. (1996) Treatment with growth hormone and dexamethasone in mice transgenic for human islet amyloid polypeptide causes islet amyloidosis and beta-cell dysfunction. Diabetes 45: 1094-1101.

195. D'Alessio DA, Verchere CB, Kahn SE, Hoagland V, Baskin DG, et al. (1994) Pancreatic expression and secretion of human islet amyloid polypeptide in a transgenic mouse. Diabetes 43: 1457-1461.

196. Fox N, Schrementi J, Nishi M, Ohagi S, Chan SJ, et al. (1993) Human islet amyloid polypeptide transgenic mice as a model of non-insulin-dependent diabetes mellitus (NIDDM). FEBS letters 323: 40-44.

197. Yagui K, Yamaguchi T, Kanatsuka A, Shimada F, Huang CI, et al. (1995) Formation of islet amyloid fibrils in beta-secretory granules of transgenic mice expressing human islet amyloid polypeptide/amylin. European journal of endocrinology / European Federation of Endocrine Societies 132: 487-496.

198. Janson J, Soeller WC, Roche PC, Nelson RT, Torchia AJ, et al. (1996) Spontaneous diabetes mellitus in transgenic mice expressing human islet amyloid polypeptide. Proceedings of the National Academy of Sciences of the United States of America 93: 7283-7288.

References    

 

 

  89  

199. Verchere CB, D'Alessio DA, Palmiter RD, Weir GC, Bonner-Weir S, et al. (1996) Islet amyloid formation associated with hyperglycemia in transgenic mice with pancreatic beta cell expression of human islet amyloid polypeptide. Proceedings of the National Academy of Sciences of the United States of America 93: 3492-3496.

200. Kahn SE, Andrikopoulos S, Verchere CB, Wang F, Hull RL, et al. (2000) Oophorectomy promotes islet amyloid formation in a transgenic mouse model of Type II diabetes. Diabetologia 43: 1309-1312.

201. Butler AE, Jang J, Gurlo T, Carty MD, Soeller WC, et al. (2004) Diabetes due to a progressive defect in beta-cell mass in rats transgenic for human islet amyloid polypeptide (HIP Rat): a new model for type 2 diabetes. Diabetes 53: 1509-1516.

202. Crews L, Masliah E (2010) Molecular mechanisms of neurodegeneration in Alzheimer's disease. Human molecular genetics 19: R12-20.

203. Ittner LM, Gotz J (2011) Amyloid-beta and tau--a toxic pas de deux in Alzheimer's disease. Nature reviews Neuroscience 12: 65-72.

204. Association As (2010) 2010 Alzheimer's disease facts and figures. Alzheimer's & dementia : the journal of the Alzheimer's Association 6: 158-194.

205. Ballard C, Gauthier S, Corbett A, Brayne C, Aarsland D, et al. (2011) Alzheimer's disease. Lancet 377: 1019-1031.

206. Perrin RJ, Fagan AM, Holtzman DM (2009) Multimodal techniques for diagnosis and prognosis of Alzheimer's disease. Nature 461: 916-922.

207. Spies PE, Slats D, Sjogren JM, Kremer BP, Verhey FR, et al. (2010) The cerebrospinal fluid amyloid beta42/40 ratio in the differentiation of Alzheimer's disease from non-Alzheimer's dementia. Current Alzheimer research 7: 470-476.

208. Shoji M, Kanai M, Matsubara E, Tomidokoro Y, Shizuka M, et al. (2001) The levels of cerebrospinal fluid Abeta40 and Abeta42(43) are regulated age-dependently. Neurobiology of aging 22: 209-215.

209. Walsh DM, Selkoe DJ (2004) Oligomers on the brain: the emerging role of soluble protein aggregates in neurodegeneration. Protein and peptide letters 11: 213-228.

210. Rabinovici GD, Jagust WJ (2009) Amyloid imaging in aging and dementia: testing the amyloid hypothesis in vivo. Behavioural neurology 21: 117-128.

211. Nilsberth C, Westlind-Danielsson A, Eckman CB, Condron MM, Axelman K, et al. (2001) The 'Arctic' APP mutation (E693G) causes Alzheimer's disease by enhanced Abeta protofibril formation. Nature neuroscience 4: 887-893.

212. Philipson O, Lord A, Gumucio A, O'Callaghan P, Lannfelt L, et al. (2010) Animal models of amyloid-beta-related pathologies in Alzheimer's disease. The FEBS journal 277: 1389-1409.

213. Goedert M, Spillantini MG (2006) A century of Alzheimer's disease. Science 314: 777-781.

214. Ott A, Stolk RP, van Harskamp F, Pols HA, Hofman A, et al. (1999) Diabetes mellitus and the risk of dementia: The Rotterdam Study. Neurology 53: 1937-1942.

215. Gotz J, Ittner LM, Lim YA (2009) Common features between diabetes mellitus and Alzheimer's disease. Cellular and molecular life sciences : CMLS 66: 1321-1325.

216. Janson J, Laedtke T, Parisi JE, O'Brien P, Petersen RC, et al. (2004) Increased risk of type 2 diabetes in Alzheimer disease. Diabetes 53: 474-481.

217. Bennett RG, Hamel FG, Duckworth WC (2003) An insulin-degrading enzyme inhibitor decreases amylin degradation, increases amylin-induced cytotoxicity, and increases amyloid formation in insulinoma cell cultures. Diabetes 52: 2315-2320.

References      

 

 90  

218. Farris W, Mansourian S, Leissring MA, Eckman EA, Bertram L, et al. (2004) Partial loss-of-function mutations in insulin-degrading enzyme that induce diabetes also impair degradation of amyloid beta-protein. The American journal of pathology 164: 1425-1434.

219. Qiu WQ, Folstein MF (2006) Insulin, insulin-degrading enzyme and amyloid-beta peptide in Alzheimer's disease: review and hypothesis. Neurobiology of aging 27: 190-198.

220. Shen Y, Joachimiak A, Rosner MR, Tang WJ (2006) Structures of human insulin-degrading enzyme reveal a new substrate recognition mechanism. Nature 443: 870-874.

221. O'Nuallain B, Williams AD, Westermark P, Wetzel R (2004) Seeding specificity in amyloid growth induced by heterologous fibrils. The Journal of biological chemistry 279: 17490-17499.

222. Jhamandas JH, Li Z, Westaway D, Yang J, Jassar S, et al. (2011) Actions of beta-amyloid protein on human neurons are expressed through the amylin receptor. The American journal of pathology 178: 140-149.

223. Sturtevant AH (1913) The Linnear Arrangement of Six Sex-linked Factors in Drosophila, asshown by Their Mode of Association. Journal of Experimental Zoology: 43-59.

224. Rubin GM, Lewis EB (2000) A brief history of Drosophila's contributions to genome research. Science 287: 2216-2218.

225. Bellen HJ, Tong C, Tsuda H (2010) 100 years of Drosophila research and its impact on vertebrate neuroscience: a history lesson for the future. Nature reviews Neuroscience 11: 514-522.

226. Muller HJ (1927) Artificial Transmutation of the Gene. Science 66: 84-87. 227. Lewis EB, Bacher F (1968) Methods of feeding ethyl methane sulfonate

(EMS) to Drosophila males. Drosophila Information Service 43: 193. 228. McGuire SE, Deshazer M, Davis RL (2005) Thirty years of olfactory learning

and memory research in Drosophila melanogaster. Progress in neurobiology 76: 328-347.

229. Yan N, Shi Y (2005) Mechanisms of apoptosis through structural biology. Annual review of cell and developmental biology 21: 35-56.

230. Reiter LT, Potocki L, Chien S, Gribskov M, Bier E (2001) A systematic analysis of human disease-associated gene sequences in Drosophila melanogaster. Genome research 11: 1114-1125.

231. Rasheva VI, Domingos PM (2009) Cellular responses to endoplasmic reticulum stress and apoptosis. Apoptosis : an international journal on programmed cell death 14: 996-1007.

232. McPhee CK, Baehrecke EH (2009) Autophagy in Drosophila melanogaster. Biochimica et biophysica acta 1793: 1452-1460.

233. Zirin J, Perrimon N (2010) Drosophila as a model system to study autophagy. Seminars in immunopathology 32: 363-372.

234. Adams MD, Celniker SE, Holt RA, Evans CA, Gocayne JD, et al. (2000) The genome sequence of Drosophila melanogaster. Science 287: 2185-2195.

235. Brand AH, Perrimon N (1993) Targeted gene expression as a means of altering cell fates and generating dominant phenotypes. Development 118: 401-415.

236. Petersen NS (1990) Effects of heat and chemical stress on development. Advances in genetics 28: 275-296.

237. Petersen NS, Mitchell HK (1987) The induction of a multiple wing hair phenocopy by heat shock in mutant heterozygotes. Developmental biology 121: 335-341.

238. Yost HJ, Petersen RB, Lindquist S (1990) RNA metabolism: strategies for regulation in the heat shock response. Trends in genetics : TIG 6: 223-227.

References    

 

 

  91  

239. Pfeiffer BD, Jenett A, Hammonds AS, Ngo TT, Misra S, et al. (2008) Tools for neuroanatomy and neurogenetics in Drosophila. Proceedings of the National Academy of Sciences of the United States of America 105: 9715-9720.

240. Duffy JB (2002) GAL4 system in Drosophila: a fly geneticist's Swiss army knife. Genesis 34: 1-15.

241. Jones WD (2009) The expanding reach of the GAL4/UAS system into the behavioral neurobiology of Drosophila. BMB reports 42: 705-712.

242. Perrimon N, Ni JQ, Perkins L (2010) In vivo RNAi: today and tomorrow. Cold Spring Harbor perspectives in biology 2: a003640.

243. Lee T, Luo L (2001) Mosaic analysis with a repressible cell marker (MARCM) for Drosophila neural development. Trends in neurosciences 24: 251-254.

244. Greeve I, Kretzschmar D, Tschape JA, Beyn A, Brellinger C, et al. (2004) Age-dependent neurodegeneration and Alzheimer-amyloid plaque formation in transgenic Drosophila. The Journal of neuroscience : the official journal of the Society for Neuroscience 24: 3899-3906.

245. Crowther DC, Kinghorn KJ, Miranda E, Page R, Curry JA, et al. (2005) Intraneuronal Abeta, non-amyloid aggregates and neurodegeneration in a Drosophila model of Alzheimer's disease. Neuroscience 132: 123-135.

246. Finelli A, Kelkar A, Song HJ, Yang H, Konsolaki M (2004) A model for studying Alzheimer's Abeta42-induced toxicity in Drosophila melanogaster. Molecular and cellular neurosciences 26: 365-375.

247. Iijima K, Liu HP, Chiang AS, Hearn SA, Konsolaki M, et al. (2004) Dissecting the pathological effects of human Abeta40 and Abeta42 in Drosophila: a potential model for Alzheimer's disease. Proceedings of the National Academy of Sciences of the United States of America 101: 6623-6628.

248. Iijima-Ando K, Iijima K (2010) Transgenic Drosophila models of Alzheimer's disease and tauopathies. Brain structure & function 214: 245-262.

249. Nerelius C, Sandegren A, Sargsyan H, Raunak R, Leijonmarck H, et al. (2009) Alpha-helix targeting reduces amyloid-beta peptide toxicity. Proceedings of the National Academy of Sciences of the United States of America 106: 9191-9196.

250. Luheshi LM, Tartaglia GG, Brorsson AC, Pawar AP, Watson IE, et al. (2007) Systematic in vivo analysis of the intrinsic determinants of amyloid Beta pathogenicity. PLoS biology 5: e290.

251. Cao W, Song HJ, Gangi T, Kelkar A, Antani I, et al. (2008) Identification of novel genes that modify phenotypes induced by Alzheimer's beta-amyloid overexpression in Drosophila. Genetics 178: 1457-1471.

252. Rival T, Page RM, Chandraratna DS, Sendall TJ, Ryder E, et al. (2009)

Fenton chemistry and oxidative stress mediate the toxicity of the beta-amyloid peptide in a Drosophila model of Alzheimer's disease. The European journal of neuroscience 29: 1335-1347.

253. Ling D, Song HJ, Garza D, Neufeld TP, Salvaterra PM (2009) Abeta42-induced neurodegeneration via an age-dependent autophagic-lysosomal injury in Drosophila. PLoS One 4: e4201.

254. Jackson GR, Wiedau-Pazos M, Sang TK, Wagle N, Brown CA, et al. (2002) Human wild-type tau interacts with wingless pathway components and produces neurofibrillary pathology in Drosophila. Neuron 34: 509-519.

255. Wittmann CW, Wszolek MF, Shulman JM, Salvaterra PM, Lewis J, et al. (2001) Tauopathy in Drosophila: neurodegeneration without neurofibrillary tangles. Science 293: 711-714.

256. Nishimura I, Yang Y, Lu B (2004) PAR-1 kinase plays an initiator role in a temporally ordered phosphorylation process that confers tau toxicity in Drosophila. Cell 116: 671-682.

References      

 

 92  

257. Berg I, Thor S, Hammarstrom P (2009) Modeling familial amyloidotic polyneuropathy (Transthyretin V30M) in Drosophila melanogaster. Neuro-degenerative diseases 6: 127-138.

258. Pokrzywa M, Dacklin I, Hultmark D, Lundgren E (2007) Misfolded transthyretin causes behavioral changes in a Drosophila model for transthyretin-associated amyloidosis. Eur J Neurosci 26: 913-924.

259. Pokrzywa M, Dacklin I, Vestling M, Hultmark D, Lundgren E, et al. (2010) Uptake of aggregating transthyretin by fat body in a Drosophila model for TTR-associated amyloidosis. PLoS One 5: e14343.

260. Botella JA, Bayersdorfer F, Gmeiner F, Schneuwly S (2009) Modelling Parkinson's disease in Drosophila. Neuromolecular medicine 11: 268-280.

261. Feany MB, Bender WW (2000) A Drosophila model of Parkinson's disease. Nature 404: 394-398.

262. Auluck PK, Chan HY, Trojanowski JQ, Lee VM, Bonini NM (2002) Chaperone suppression of alpha-synuclein toxicity in a Drosophila model for Parkinson's disease. Science 295: 865-868.

263. Jackson GR, Salecker I, Dong X, Yao X, Arnheim N, et al. (1998) Polyglutamine-expanded human huntingtin transgenes induce degeneration of Drosophila photoreceptor neurons. Neuron 21: 633-642.

264. Ambegaokar SS, Roy B, Jackson GR (2010) Neurodegenerative models in Drosophila: polyglutamine disorders, Parkinson disease, and amyotrophic lateral sclerosis. Neurobiology of disease 40: 29-39.

265. Pandey UB, Nie Z, Batlevi Y, McCray BA, Ritson GP, et al. (2007) HDAC6 rescues neurodegeneration and provides an essential link between autophagy and the UPS. Nature 447: 859-863.

266. Nelson B, Nishimura S, Kanuka H, Kuranaga E, Inoue M, et al. (2005) Isolation of gene sets affected specifically by polyglutamine expression: implication of the TOR signaling pathway in neurodegeneration. Cell death and differentiation 12: 1115-1123.

267. Rusten TE, Filimonenko M, Rodahl LM, Stenmark H, Simonsen A (2007) ESCRTing autophagic clearance of aggregating proteins. Autophagy 4.

268. Rincon-Limas DE, Casas-Tinto S, Fernandez-Funez P (2010) Exploring prion protein biology in flies: genetics and beyond. Prion 4: 1-8.

269. Goloubinoff P, De Los Rios P (2007) The mechanism of Hsp70 chaperones: (entropic) pulling the models together. Trends in biochemical sciences 32: 372-380.

270. Muchowski PJ, Wacker JL (2005) Modulation of neurodegeneration by molecular chaperones. Nature reviews Neuroscience 6: 11-22.

271. Buchberger A, Bukau B, Sommer T (2010) Protein quality control in the cytosol and the endoplasmic reticulum: brothers in arms. Molecular cell 40: 238-252.

272. Bukau B, Weissman J, Horwich A (2006) Molecular chaperones and protein quality control. Cell 125: 443-451.

273. Ellgaard L, Helenius A (2003) Quality control in the endoplasmic reticulum. Nature reviews Molecular cell biology 4: 181-191.

274. Ron D, Walter P (2007) Signal integration in the endoplasmic reticulum unfolded protein response. Nature reviews Molecular cell biology 8: 519-529.

275. Aebi M, Bernasconi R, Clerc S, Molinari M (2010) N-glycan structures: recognition and processing in the ER. Trends in biochemical sciences 35: 74-82.

276. Caramelo JJ, Castro OA, Alonso LG, De Prat-Gay G, Parodi AJ (2003) UDP-Glc:glycoprotein glucosyltransferase recognizes structured and solvent accessible hydrophobic patches in molten globule-like folding intermediates. Proceedings of the National Academy of Sciences of the United States of America 100: 86-91.

References    

 

 

  93  

277. Caramelo JJ, Castro OA, de Prat-Gay G, Parodi AJ (2004) The endoplasmic reticulum glucosyltransferase recognizes nearly native glycoprotein folding intermediates. The Journal of biological chemistry 279: 46280-46285.

278. Maattanen P, Gehring K, Bergeron JJ, Thomas DY (2010) Protein quality control in the ER: the recognition of misfolded proteins. Seminars in cell & developmental biology 21: 500-511.

279. Zhou J, Liu CY, Back SH, Clark RL, Peisach D, et al. (2006) The crystal structure of human IRE1 luminal domain reveals a conserved dimerization interface required for activation of the unfolded protein response. Proceedings of the National Academy of Sciences of the United States of America 103: 14343-14348.

280. Credle JJ, Finer-Moore JS, Papa FR, Stroud RM, Walter P (2005) On the mechanism of sensing unfolded protein in the endoplasmic reticulum. Proceedings of the National Academy of Sciences of the United States of America 102: 18773-18784.

281. Kimata Y, Ishiwata-Kimata Y, Ito T, Hirata A, Suzuki T, et al. (2007) Two regulatory steps of ER-stress sensor Ire1 involving its cluster formation and interaction with unfolded proteins. The Journal of cell biology 179: 75-86.

282. Kimata Y, Oikawa D, Shimizu Y, Ishiwata-Kimata Y, Kohno K (2004) A role for BiP as an adjustor for the endoplasmic reticulum stress-sensing protein Ire1. The Journal of cell biology 167: 445-456.

283. Papa FR, Zhang C, Shokat K, Walter P (2003) Bypassing a kinase activity with an ATP-competitive drug. Science 302: 1533-1537.

284. Calfon M, Zeng H, Urano F, Till JH, Hubbard SR, et al. (2002) IRE1 couples endoplasmic reticulum load to secretory capacity by processing the XBP-1 mRNA. Nature 415: 92-96.

285. Yoshida H, Matsui T, Yamamoto A, Okada T, Mori K (2001) XBP1 mRNA is induced by ATF6 and spliced by IRE1 in response to ER stress to produce a highly active transcription factor. Cell 107: 881-891.

286. Yoshida H, Matsui T, Hosokawa N, Kaufman RJ, Nagata K, et al. (2003) A time-dependent phase shift in the mammalian unfolded protein response. Developmental cell 4: 265-271.

287. Yoneda T, Imaizumi K, Oono K, Yui D, Gomi F, et al. (2001) Activation of caspase-12, an endoplastic reticulum (ER) resident caspase, through tumor necrosis factor receptor-associated factor 2-dependent mechanism in response to the ER stress. The Journal of biological chemistry 276: 13935-13940.

288. Hetz C, Bernasconi P, Fisher J, Lee AH, Bassik MC, et al. (2006) Proapoptotic BAX and BAK modulate the unfolded protein response by a direct interaction with IRE1alpha. Science 312: 572-576.

289. Liu Y, Chang A (2008) Heat shock response relieves ER stress. The EMBO journal 27: 1049-1059.

290. Meusser B, Hirsch C, Jarosch E, Sommer T (2005) ERAD: the long road to destruction. Nature cell biology 7: 766-772.

291. Fu L, Sztul E (2009) ER-associated complexes (ERACs) containing aggregated cystic fibrosis transmembrane conductance regulator (CFTR) are degraded by autophagy. European journal of cell biology 88: 215-226.

292. Huyer G, Longsworth GL, Mason DL, Mallampalli MP, McCaffery JM, et al. (2004) A striking quality control subcompartment in Saccharomyces cerevisiae: the endoplasmic reticulum-associated compartment. Molecular biology of the cell 15: 908-921.

293. Bernales S, McDonald KL, Walter P (2006) Autophagy counterbalances endoplasmic reticulum expansion during the unfolded protein response. PLoS biology 4: e423.

294. He C, Klionsky DJ (2009) Regulation mechanisms and signaling pathways of autophagy. Annual review of genetics 43: 67-93.

References      

 

 94  

295. Cheng EH, Wei MC, Weiler S, Flavell RA, Mak TW, et al. (2001) BCL-2, BCL-X(L) sequester BH3 domain-only molecules preventing BAX- and BAK-mediated mitochondrial apoptosis. Molecular cell 8: 705-711.

296. Urano F, Wang X, Bertolotti A, Zhang Y, Chung P, et al. (2000) Coupling of stress in the ER to activation of JNK protein kinases by transmembrane protein kinase IRE1. Science 287: 664-666.

297. Lin JH, Li H, Yasumura D, Cohen HR, Zhang C, et al. (2007) IRE1 signaling affects cell fate during the unfolded protein response. Science 318: 944-949.

298. Nakagawa T, Yuan J (2000) Cross-talk between two cysteine protease families. Activation of caspase-12 by calpain in apoptosis. The Journal of cell biology 150: 887-894.

299. Nakagawa T, Zhu H, Morishima N, Li E, Xu J, et al. (2000) Caspase-12 mediates endoplasmic-reticulum-specific apoptosis and cytotoxicity by amyloid-beta. Nature 403: 98-103.

300. Rutkowski DT, Arnold SM, Miller CN, Wu J, Li J, et al. (2006) Adaptation to ER stress is mediated by differential stabilities of pro-survival and pro-apoptotic mRNAs and proteins. PLoS biology 4: e374.

301. Huang CJ, Haataja L, Gurlo T, Butler AE, Wu X, et al. (2007) Induction of endoplasmic reticulum stress-induced beta-cell apoptosis and accumulation of polyubiquitinated proteins by human islet amyloid polypeptide. American journal of physiology Endocrinology and metabolism 293: E1656-1662.

302. Huang CJ, Lin CY, Haataja L, Gurlo T, Butler AE, et al. (2007) High expression rates of human islet amyloid polypeptide induce endoplasmic reticulum stress mediated beta-cell apoptosis, a characteristic of humans with type 2 but not type 1 diabetes. Diabetes 56: 2016-2027.

303. Marchetti P, Bugliani M, Lupi R, Marselli L, Masini M, et al. (2007) The endoplasmic reticulum in pancreatic beta cells of type 2 diabetes patients. Diabetologia 50: 2486-2494.

304. Laybutt DR, Preston AM, Akerfeldt MC, Kench JG, Busch AK, et al. (2007) Endoplasmic reticulum stress contributes to beta cell apoptosis in type 2 diabetes. Diabetologia 50: 752-763.

305. Casas S, Gomis R, Gribble FM, Altirriba J, Knuutila S, et al. (2007) Impairment of the ubiquitin-proteasome pathway is a downstream endoplasmic reticulum stress response induced by extracellular human islet amyloid polypeptide and contributes to pancreatic beta-cell apoptosis. Diabetes 56: 2284-2294.

306. Hull RL, Zraika S, Udayasankar J, Aston-Mourney K, Subramanian SL, et al. (2009) Amyloid formation in human IAPP transgenic mouse islets and pancreas, and human pancreas, is not associated with endoplasmic reticulum stress. Diabetologia 52: 1102-1111.

307. Gurlo T, Ryazantsev S, Huang CJ, Yeh MW, Reber HA, et al. (2010) Evidence for proteotoxicity in beta cells in type 2 diabetes: toxic islet amyloid polypeptide oligomers form intracellularly in the secretory pathway. Am J Pathol 176: 861-869.

308. Richardson H, Kumar S (2002) Death to flies: Drosophila as a model system to study programmed cell death. Journal of immunological methods 265: 21-38.

309. Cory S, Adams JM (2002) The Bcl2 family: regulators of the cellular life-or-death switch. Nature reviews Cancer 2: 647-656.

310. Wang X (2001) The expanding role of mitochondria in apoptosis. Genes & development 15: 2922-2933.

311. Zou H, Li Y, Liu X, Wang X (1999) An APAF-1.cytochrome c multimeric complex is a functional apoptosome that activates procaspase-9. The Journal of biological chemistry 274: 11549-11556.

References    

 

 

  95  

312. Li H, Zhu H, Xu CJ, Yuan J (1998) Cleavage of BID by caspase 8 mediates the mitochondrial damage in the Fas pathway of apoptosis. Cell 94: 491-501.

313. Luo X, Budihardjo I, Zou H, Slaughter C, Wang X (1998) Bid, a Bcl2 interacting protein, mediates cytochrome c release from mitochondria in response to activation of cell surface death receptors. Cell 94: 481-490.

314. Ditzel M, Broemer M, Tenev T, Bolduc C, Lee TV, et al. (2008) Inactivation of effector caspases through nondegradative polyubiquitylation. Molecular cell 32: 540-553.

315. Butler AE, Janson J, Soeller WC, Butler PC (2003) Increased beta-cell apoptosis prevents adaptive increase in beta-cell mass in mouse model of type 2 diabetes: evidence for role of islet amyloid formation rather than direct action of amyloid. Diabetes 52: 2304-2314.

316. Rumora L, Hadzija M, Barisic K, Maysinger D, Grubiic TZ (2002) Amylin-induced cytotoxicity is associated with activation of caspase-3 and MAP kinases. Biological chemistry 383: 1751-1758.

317. Zraika S, Hull RL, Udayasankar J, Aston-Mourney K, Subramanian SL, et al. (2009) Oxidative stress is induced by islet amyloid formation and time-dependently mediates amyloid-induced beta cell apoptosis. Diabetologia 52: 626-635.

318. Li XL, Chen T, Wong YS, Xu G, Fan RR, et al. (2011) Involvement of mitochondrial dysfunction in human islet amyloid polypeptide-induced apoptosis in INS-1E pancreatic beta cells: An effect attenuated by phycocyanin. The international journal of biochemistry & cell biology 43: 525-534.

319. Law E, Lu S, Kieffer TJ, Warnock GL, Ao Z, et al. (2010) Differences between amyloid toxicity in alpha and beta cells in human and mouse islets and the role of caspase-3. Diabetologia 53: 1415-1427.

320. Matveyenko AV, Butler PC (2006) Beta-cell deficit due to increased apoptosis in the human islet amyloid polypeptide transgenic (HIP) rat recapitulates the metabolic defects present in type 2 diabetes. Diabetes 55: 2106-2114.

321. Reggiori F, Klionsky DJ (2002) Autophagy in the eukaryotic cell. Eukaryotic cell 1: 11-21.

322. De Duve C (1963) The lysosome. Scientific American 208: 64-72. 323. Uttenweiler A, Schwarz H, Neumann H, Mayer A (2007) The vacuolar

transporter chaperone (VTC) complex is required for microautophagy. Molecular biology of the cell 18: 166-175.

324. Sahu R, Kaushik S, Clement CC, Cannizzo ES, Scharf B, et al. (2011) Microautophagy of cytosolic proteins by late endosomes. Developmental cell 20: 131-139.

325. Cuervo AM (2010) Chaperone-mediated autophagy: selectivity pays off. Trends in endocrinology and metabolism: TEM 21: 142-150.

326. Seglen PO, Gordon PB, Holen I (1990) Non-selective autophagy. Seminars in cell biology 1: 441-448.

327. Tooze SA, Yoshimori T (2010) The origin of the autophagosomal membrane. Nature cell biology 12: 831-835.

328. McEwan DG, Dikic I (2010) Not all autophagy membranes are created equal. Cell 141: 564-566.

329. Yorimitsu T, Klionsky DJ (2005) Autophagy: molecular machinery for self-eating. Cell death and differentiation 12 Suppl 2: 1542-1552.

330. Knaevelsrud H, Simonsen A (2010) Fighting disease by selective autophagy of aggregate-prone proteins. FEBS Lett 584: 2635-2645.

331. Youle RJ, Narendra DP (2011) Mechanisms of mitophagy. Nature reviews Molecular cell biology 12: 9-14.

References      

 

 96  

332. Kraft C, Deplazes A, Sohrmann M, Peter M (2008) Mature ribosomes are selectively degraded upon starvation by an autophagy pathway requiring the Ubp3p/Bre5p ubiquitin protease. Nature cell biology 10: 602-610.

333. Dunn WA, Jr., Cregg JM, Kiel JA, van der Klei IJ, Oku M, et al. (2005) Pexophagy: the selective autophagy of peroxisomes. Autophagy 1: 75-83.

334. Klionsky DJ, Cuervo AM, Dunn WA, Jr., Levine B, van der Klei I, et al. (2007) How shall I eat thee? Autophagy 3: 413-416.

335. Yamamoto A, Simonsen A (2010) The elimination of accumulated and aggregated proteins: A role for aggrephagy in neurodegeneration. Neurobiol Dis.

336. Noda T, Ohsumi Y (1998) Tor, a phosphatidylinositol kinase homologue, controls autophagy in yeast. The Journal of biological chemistry 273: 3963-3966.

337. Scott RC, Schuldiner O, Neufeld TP (2004) Role and regulation of starvation-induced autophagy in the Drosophila fat body. Developmental cell 7: 167-178.

338. Hara T, Takamura A, Kishi C, Iemura S, Natsume T, et al. (2008) FIP200, a ULK-interacting protein, is required for autophagosome formation in mammalian cells. The Journal of cell biology 181: 497-510.

339. Simonsen A, Tooze SA (2009) Coordination of membrane events during autophagy by multiple class III PI3-kinase complexes. The Journal of cell biology 186: 773-782.

340. Razi M, Chan EY, Tooze SA (2009) Early endosomes and endosomal coatomer are required for autophagy. The Journal of cell biology 185: 305-321.

341. Rusten TE, Stenmark H (2009) How do ESCRT proteins control autophagy? Journal of cell science 122: 2179-2183.

342. Nickerson DP, Brett CL, Merz AJ (2009) Vps-C complexes: gatekeepers of endolysosomal traffic. Current opinion in cell biology 21: 543-551.

343. Lindmo K, Simonsen A, Brech A, Finley K, Rusten TE, et al. (2006) A dual function for Deep orange in programmed autophagy in the Drosophila melanogaster fat body. Experimental cell research 312: 2018-2027.

344. Zhong Y, Wang QJ, Li X, Yan Y, Backer JM, et al. (2009) Distinct regulation of autophagic activity by Atg14L and Rubicon associated with Beclin 1-phosphatidylinositol-3-kinase complex. Nature cell biology 11: 468-476.

345. Pankiv S, Alemu EA, Brech A, Bruun JA, Lamark T, et al. (2010) FYCO1 is a Rab7 effector that binds to LC3 and PI3P to mediate microtubule plus end-directed vesicle transport. The Journal of cell biology 188: 253-269.

346. Kimura S, Noda T, Yoshimori T (2008) Dynein-dependent movement of autophagosomes mediates efficient encounters with lysosomes. Cell structure and function 33: 109-122.

347. Kirkin V, McEwan DG, Novak I, Dikic I (2009) A role for ubiquitin in selective autophagy. Molecular cell 34: 259-269.

348. Hershko A, Ciechanover A (1998) The ubiquitin system. Annual review of biochemistry 67: 425-479.

349. Welchman RL, Gordon C, Mayer RJ (2005) Ubiquitin and ubiquitin-like proteins as multifunctional signals. Nature reviews Molecular cell biology 6: 599-609.

350. Tan JM, Wong ES, Kirkpatrick DS, Pletnikova O, Ko HS, et al. (2008) Lysine 63-linked ubiquitination promotes the formation and autophagic clearance of protein inclusions associated with neurodegenerative diseases. Human molecular genetics 17: 431-439.

351. Wooten MW, Geetha T, Babu JR, Seibenhener ML, Peng J, et al. (2008) Essential role of sequestosome 1/p62 in regulating accumulation of Lys63-ubiquitinated proteins. The Journal of biological chemistry 283: 6783-6789.

References    

 

 

  97  

352. Ikeda F, Dikic I (2008) Atypical ubiquitin chains: new molecular signals. 'Protein Modifications: Beyond the Usual Suspects' review series. EMBO reports 9: 536-542.

353. Kirkin V, Lamark T, Sou YS, Bjorkoy G, Nunn JL, et al. (2009) A role for NBR1 in autophagosomal degradation of ubiquitinated substrates. Mol Cell 33: 505-516.

354. Pankiv S, Clausen TH, Lamark T, Brech A, Bruun JA, et al. (2007) p62/SQSTM1 binds directly to Atg8/LC3 to facilitate degradation of ubiquitinated protein aggregates by autophagy. J Biol Chem 282: 24131-24145.

355. Bjorkoy G, Lamark T, Brech A, Outzen H, Perander M, 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: 603-614.

356. Nezis IP, Simonsen A, Sagona AP, Finley K, Gaumer S, et al. (2008) Ref(2)P, the Drosophila melanogaster homologue of mammalian p62, is required for the formation of protein aggregates in adult brain. J Cell Biol 180: 1065-1071.

357. Clausen TH, Lamark T, Isakson P, Finley K, Larsen KB, et al. (2010) p62/SQSTM1 and ALFY interact to facilitate the formation of p62 bodies/ALIS and their degradation by autophagy. Autophagy 6: 330-344.

358. Komatsu M, Waguri S, Koike M, Sou YS, Ueno T, et al. (2007) Homeostatic levels of p62 control cytoplasmic inclusion body formation in autophagy-deficient mice. Cell 131: 1149-1163.

359. Klionsky DJ, Abeliovich H, Agostinis P, Agrawal DK, Aliev G, et al. (2008) Guidelines for the use and interpretation of assays for monitoring autophagy in higher eukaryotes. Autophagy 4: 151-175.

360. Kuusisto E, Salminen A, Alafuzoff I (2001) Ubiquitin-binding protein p62 is present in neuronal and glial inclusions in human tauopathies and synucleinopathies. Neuroreport 12: 2085-2090.

361. Zatloukal K, Stumptner C, Fuchsbichler A, Heid H, Schnoelzer M, et al. (2002) p62 Is a common component of cytoplasmic inclusions in protein aggregation diseases. The American journal of pathology 160: 255-263.

362. Filimonenko M, Isakson P, Finley KD, Anderson M, Jeong H, et al. (2010) The selective macroautophagic degradation of aggregated proteins requires the PI3P-binding protein Alfy. Mol Cell 38: 265-279.

363. Simonsen A, Birkeland HC, Gillooly DJ, Mizushima N, Kuma A, et al. (2004) Alfy, a novel FYVE-domain-containing protein associated with protein granules and autophagic membranes. Journal of cell science 117: 4239-4251.

364. Fujita N, Itoh T, Omori H, Fukuda M, Noda T, et al. (2008) The Atg16L complex specifies the site of LC3 lipidation for membrane biogenesis in autophagy. Molecular biology of the cell 19: 2092-2100.

365. Finley KD, Edeen PT, Cumming RC, Mardahl-Dumesnil MD, Taylor BJ, et al. (2003) blue cheese mutations define a novel, conserved gene involved in progressive neural degeneration. The Journal of neuroscience : the official journal of the Society for Neuroscience 23: 1254-1264.

366. Narendra D, Kane LA, Hauser DN, Fearnley IM, Youle RJ (2010) p62/SQSTM1 is required for Parkin-induced mitochondrial clustering but not mitophagy; VDAC1 is dispensable for both. Autophagy 6: 1090-1106.

367. Geisler S, Holmstrom KM, Skujat D, Fiesel FC, Rothfuss OC, et al. (2010) PINK1/Parkin-mediated mitophagy is dependent on VDAC1 and p62/SQSTM1. Nature cell biology 12: 119-131.

368. Kim PK, Hailey DW, Mullen RT, Lippincott-Schwartz J (2008) Ubiquitin signals autophagic degradation of cytosolic proteins and peroxisomes. Proceedings of the National Academy of Sciences of the United States of America 105: 20567-20574.

References      

 

 98  

369. Nagata E, Sawa A, Ross CA, Snyder SH (2004) Autophagosome-like vacuole formation in Huntington's disease lymphoblasts. Neuroreport 15: 1325-1328.

370. Ravikumar B, Vacher C, Berger Z, Davies JE, Luo S, et al. (2004) Inhibition of mTOR induces autophagy and reduces toxicity of polyglutamine expansions in fly and mouse models of Huntington disease. Nat Genet 36: 585-595.

371. Webb JL, Ravikumar B, Atkins J, Skepper JN, Rubinsztein DC (2003) Alpha-Synuclein is degraded by both autophagy and the proteasome. J Biol Chem 278: 25009-25013.

372. Cuervo AM, Stefanis L, Fredenburg R, Lansbury PT, Sulzer D (2004) Impaired degradation of mutant alpha-synuclein by chaperone-mediated autophagy. Science 305: 1292-1295.

373. Parkinson N, Ince PG, Smith MO, Highley R, Skibinski G, et al. (2006) ALS phenotypes with mutations in CHMP2B (charged multivesicular body protein 2B). Neurology 67: 1074-1077.

374. Talbot K, Ansorge O (2006) Recent advances in the genetics of amyotrophic lateral sclerosis and frontotemporal dementia: common pathways in neurodegenerative disease. Human molecular genetics 15 Spec No 2: R182-187.

375. Munch C, Sedlmeier R, Meyer T, Homberg V, Sperfeld AD, et al. (2004) Point mutations of the p150 subunit of dynactin (DCTN1) gene in ALS. Neurology 63: 724-726.

376. Rubinsztein DC (2006) The roles of intracellular protein-degradation pathways in neurodegeneration. Nature 443: 780-786.

377. Nixon RA, Wegiel J, Kumar A, Yu WH, Peterhoff C, et al. (2005) Extensive involvement of autophagy in Alzheimer disease: an immuno-electron microscopy study. Journal of neuropathology and experimental neurology 64: 113-122.

378. Lee S, Sato Y, Nixon RA (2011) Lysosomal proteolysis inhibition selectively disrupts axonal transport of degradative organelles and causes an Alzheimer's-like axonal dystrophy. The Journal of neuroscience : the official journal of the Society for Neuroscience 31: 7817-7830.

379. Yang DS, Stavrides P, Mohan PS, Kaushik S, Kumar A, et al. (2011) Reversal of autophagy dysfunction in the TgCRND8 mouse model of Alzheimer's disease ameliorates amyloid pathologies and memory deficits. Brain : a journal of neurology 134: 258-277.

380. Ling D, Salvaterra PM (2011) Brain aging and Abeta neurotoxicity converge via deterioration in autophagy-lysosomal system: a conditional Drosophila model linking Alzheimer's neurodegeneration with aging. Acta neuropathologica 121: 183-191.

381. Simonsen A, Cumming RC, Brech A, Isakson P, Schubert DR, et al. (2008) Promoting basal levels of autophagy in the nervous system enhances longevity and oxidant resistance in adult Drosophila. Autophagy 4: 176-184.

382. Ebato C, Uchida T, Arakawa M, Komatsu M, Ueno T, et al. (2008) Autophagy is important in islet homeostasis and compensatory increase of beta cell mass in response to high-fat diet. Cell metabolism 8: 325-332.

383. Jung HS, Chung KW, Won Kim J, Kim J, Komatsu M, et al. (2008) Loss of autophagy diminishes pancreatic beta cell mass and function with resultant hyperglycemia. Cell metabolism 8: 318-324.

384. Rivera JF, Gurlo T, Daval M, Huang CJ, Matveyenko AV, et al. (2011) Human-IAPP disrupts the autophagy/lysosomal pathway in pancreatic beta-cells: protective role of p62-positive cytoplasmic inclusions. Cell death and differentiation 18: 415-426.

References    

 

 

  99  

385. Boya P, Gonzalez-Polo RA, Casares N, Perfettini JL, Dessen P, et al. (2005) Inhibition of macroautophagy triggers apoptosis. Molecular and cellular biology 25: 1025-1040.

386. Galluzzi L, Vicencio JM, Kepp O, Tasdemir E, Maiuri MC, et al. (2008) To die or not to die: that is the autophagic question. Current molecular medicine 8: 78-91.

387. Tsujimoto Y, Shimizu S (2005) Another way to die: autophagic programmed cell death. Cell death and differentiation 12 Suppl 2: 1528-1534.

388. Chen Y, Klionsky DJ (2011) The regulation of autophagy - unanswered questions. Journal of cell science 124: 161-170.

389. Denton D, Shravage B, Simin R, Mills K, Berry DL, et al. (2009) Autophagy, not apoptosis, is essential for midgut cell death in Drosophila. Current biology : CB 19: 1741-1746.

390. Voss V, Senft C, Lang V, Ronellenfitsch MW, Steinbach JP, et al. (2010) The pan-Bcl-2 inhibitor (-)-gossypol triggers autophagic cell death in malignant glioma. Molecular cancer research : MCR 8: 1002-1016.

391. Spradling AC, Rubin GM (1982) Transposition of cloned P elements into Drosophila germ line chromosomes. Science 218: 341-347.

392. Matveyenko AV, Butler PC (2006) Islet amyloid polypeptide (IAPP) transgenic rodents as models for type 2 diabetes. ILAR journal / National Research Council, Institute of Laboratory Animal Resources 47: 225-233.

393. Marzban L, Rhodes CJ, Steiner DF, Haataja L, Halban PA, et al. (2006) Impaired NH2-terminal processing of human proislet amyloid polypeptide by the prohormone convertase PC2 leads to amyloid formation and cell death. Diabetes 55: 2192-2201.

394. Aslund A, Sigurdson CJ, Klingstedt T, Grathwohl S, Bolmont T, et al. (2009) Novel pentameric thiophene derivatives for in vitro and in vivo optical imaging of a plethora of protein aggregates in cerebral amyloidoses. ACS chemical biology 4: 673-684.

395. Berg I, Nilsson KP, Thor S, Hammarstrom P (2010) Efficient imaging of amyloid deposits in Drosophila models of human amyloidoses. Nature protocols 5: 935-944.

396. Cohen E (1993) Chitin synthesis and degradation as targets for pesticide action. Archives of insect biochemistry and physiology 22: 245-261.

397. Hammarstrom P, Simon R, Nystrom S, Konradsson P, Aslund A, et al. (2010) A fluorescent pentameric thiophene derivative detects in vitro-formed prefibrillar protein aggregates. Biochemistry 49: 6838-6845.

398. Nilsson KP, Ikenberg K, Aslund A, Fransson S, Konradsson P, et al. (2010) Structural typing of systemic amyloidoses by luminescent-conjugated polymer spectroscopy. The American journal of pathology 176: 563-574.

399. Carolina R, Santiago W, María Fernanda C (2007) Neuronal death in <i>Drosophila</i> triggered by GAL4 accumulation. European Journal of Neuroscience 25: 683-694.

400. Bardet PL, Kolahgar G, Mynett A, Miguel-Aliaga I, Briscoe J, et al. (2008) A fluorescent reporter of caspase activity for live imaging. Proc Natl Acad Sci U S A 105: 13901-13905.

401. Hsu AL, Murphy CT, Kenyon C (2003) Regulation of aging and age-related disease by DAF-16 and heat-shock factor. Science 300: 1142-1145.

402. Morley JF, Morimoto RI (2004) Regulation of longevity in Caenorhabditis elegans by heat shock factor and molecular chaperones. Mol Biol Cell 15: 657-664.

403. Hara T, Nakamura K, Matsui M, Yamamoto A, Nakahara Y, et al. (2006) Suppression of basal autophagy in neural cells causes neurodegenerative disease in mice. Nature 441: 885-889.

References      

 

 100  

404. Bartlett BJ, Isakson P, Lewerenz J, Sanchez H, Kotzebue RW, et al. (2011) p62, Ref(2)P and ubiquitinated proteins are conserved markers of neuronal aging, aggregate formation and progressive autophagic defects. Autophagy 7: 572-583.

405. Clague MJ, Urbe S (2010) Ubiquitin: same molecule, different degradation pathways. Cell 143: 682-685.

406. Kirkin V, Lamark T, Johansen T, Dikic I (2009) NBR1 cooperates with p62 in selective autophagy of ubiquitinated targets. Autophagy 5: 732-733.

407. Lamark T, Kirkin V, Dikic I, Johansen T (2009) NBR1 and p62 as cargo receptors for selective autophagy of ubiquitinated targets. Cell Cycle 8: 1986-1990.

408. Nixon RA, Yang DS (2011) Autophagy failure in Alzheimer's disease-locating the primary defect. Neurobiology of disease 43: 38-45.