• No results found

Studies  on  Islet  Amyloid  Polypeptide  Aggregation:  From  Model  Organism  to  Molecular  Mechanisms      Sebastian  W  Schultz

N/A
N/A
Protected

Academic year: 2021

Share "Studies  on  Islet  Amyloid  Polypeptide  Aggregation:  From  Model  Organism  to  Molecular  Mechanisms      Sebastian  W  Schultz"

Copied!
110
0
0

Loading.... (view fulltext now)

Full text

(1)

           

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    

(2)

                                                    ©  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  

(3)

                                 

Der  Weg  ist  das  Ziel  

                                                 

(4)

Department  of  Medical  Cell  Biology   Uppsala  University,   Sweden             Opponent    

Anne  Simonsen,  Associate  Professor   Department  of  Biochemistry  

University  of  Oslo,   Norway                                                          

(5)

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.  

                             

(6)

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.    

(7)

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.    

(8)

                     

(9)

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  

(10)

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            

(11)

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  

(12)

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  

(13)

                                     

Introduction

(14)

 

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.  

(15)

 

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].  

(16)

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  

(17)

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  

(18)

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.  

(19)

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].  

   

(20)

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).  

(21)

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  

(22)

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  

(23)

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-­‐

(24)

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.  

(25)

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].  

(26)

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].    

References

Related documents

Kinetic analysis of fibril formation using SPR In analogy to all molecular interactions, the stability of the binding between an Ab monomer and the fibril- lar end is dependent on

In paper II, we worked with neuroblastoma cells to investigate the cells viability after exposure to oligomeric species of TTR. We found that early oligomers were toxic to the

Department of Physics, Chemistry and Biology Linköping University, SE- 58183 Linköping, Sweden.

This ar ti cle pre sents the de sign, pa tient re cruit ment, and in clu sion process for a ran dom ized con trolled trial com par ing bariatric surgery and in ten sive non -

How- ever, this equivalence fails once the capacitor is integrated into more advanced iontronic circuitry, as (iontronic) components then would connect between the two

After the conclusions from Paper 1, that small alterations in the thiophene backbone have impact on the properties as amyloid probe, and since a method for asymmetric

Our results demonstrate reduced propagation properties of Aβ aggregated in the presence of Aβ-seeds formed together with p-FTAA compared to Aβ-seeds formed without

Linköping 2011 Studies on Islet Am yloid P ol ypeptide Aggr eg ation : F rom Model Or ganism t o Molecular Mechanisms Sebastian W Schultz. Studies on Islet Amyloid