• Keine Ergebnisse gefunden

Evaluation of two molecular-based therapies in a mouse model of hypertrophic cardiomyopathy

N/A
N/A
Protected

Academic year: 2021

Aktie "Evaluation of two molecular-based therapies in a mouse model of hypertrophic cardiomyopathy"

Copied!
177
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Center  for  Experimental  Medicine  

University  Medical  Center  Hamburg-­‐Eppendorf  

 

 

 

 

 

Evaluation  of  two  molecular-­‐based  therapies  in  a  mouse  model  of  

hypertrophic  cardiomyopathy  

   

by  

Doreen  Stimpel  

      Dissertation  

Department  of  Chemistry    

Faculty  of  Mathematics,  Informatics  and  Natural  Sciences   University  of  Hamburg  

for  the  degree  of   Doctor  of  Natural  Sciences  

     

(2)

1.  Referee:  Prof.  Dr.  rer.  nat.  Peter  Heisig              2.  Referee:  Prof.  Dr.  med.  Thomas  Eschenhagen  

  Disputation:  19th  December  2013                                              

(3)

                     

For  my  beloved  family                                

(4)

 

T

ABLE  OF  CONTENTS

 

1   INTRODUCTION  ...  1  

1.1   Hypertrophic  cardiomyopathy  ...  3  

1.2   The  sarcomere  –  fundamental  contractile  unit  ...  6  

1.3   Cardiac  myosin-­‐binding  protein  C  ...  8  

1.3.1   Structural  role  ...  8  

1.3.2   Functional  role  ...  10  

1.3.3   MYBPC3  mutations  and  pathophysiological  mechanisms  ...  11  

1.4   Current  status  of  treatment  and  potential  novel  therapies  for  HCM  ...  13  

1.4.1   Drug  and  surgical  treatments  ...  14  

1.4.2   Molecular-­‐based  treatments  ...  16  

1.4.2.1   Spliceosome-­‐mediated  RNA  trans-­‐splicing  ...  16  

1.4.2.2   Exon  skipping/-­‐inclusion  ...  21  

1.4.2.3   Conventional  gene  therapy  ...  22  

1.5   Gene  delivery  tools:  AAV  as  a  suitable  vector  for  gene  transfer  ...  24  

1.6   Aim  of  the  work  ...  28  

2   MATERIALS  AND  METHODS  ...  30  

2.1   Materials  ...  30  

2.1.1   Mybpc3-­‐targeted  knock-­‐in  mouse  model  ...  30  

2.1.2   Cell  lines  and  isolated  neonatal  mouse  cardiomyocytes  ...  30  

2.2   Methods  ...  31  

2.2.1   Design  and  cloning  of  PTMs  ...  31  

2.2.2   Design  and  cloning  of  ‘toxic’  molecules  ...  35  

2.2.3   Design  and  cloning  of  full-­‐length  WT-­‐Mybpc3  ...  36  

2.2.4   Restriction  digestion  ...  36  

2.2.5   Ligation  ...  37  

2.2.6   Transformation  of  One  Shot®  TOP10  chemically  competent  E.  coli  ...  37  

2.2.7   Preparation  of  plasmid  DNA  ...  37  

2.2.8   Culture  and  transient  transfection  of  HEK293  cells  ...  38  

2.2.9   Production  of  recombinant  adeno-­‐associated  virus  serotypes  6  and  9  ...  38  

2.2.10   Production  of  recombinant  adenovirus  ...  39  

2.2.11   Isolation  and  culture  of  cardiomyocytes  ...  40  

2.2.12   AAV6-­‐mediated  gene  transfer  in  cardiomyocytes  ...  40  

2.2.13   Adenovirus-­‐mediated  gene  transfer  in  cardiomyocytes  ...  41  

2.2.14   DNA  and  RNA  analysis  ...  41  

2.2.14.1   RNA  isolation  ...  41  

2.2.14.2   Determination  of  the  RNA  and  DNA  concentration  ...  41  

2.2.14.3   Reverse  transcription  ...  41  

2.2.14.4   Polymerase  chain  reaction  ...  42  

2.2.14.5   Quantitative  PCR  ...  44  

2.2.14.6   Agarose  gel  electrophoresis  ...  46  

2.2.14.7   Preparative  agarose  gels  ...  46  

(5)

 

2.2.15.2   Determination  of  protein  concentration  ...  47  

2.2.15.3   Immunoprecipitation  of  proteins  ...  47  

2.2.15.4   Western  blot  ...  48  

2.2.15.5   Immunofluorescence  analysis  of  cardiomyocytes  ...  49  

2.2.16   In  vivo  studies  ...  49  

2.2.16.1   AAV9-­‐mediated  gene  transfer  in  Mybpc3-­‐targeted  KI  mice  ...  49  

2.2.16.2   In  vivo  bioluminescence  imaging  ...  49  

2.2.16.3   Echocardiography  ...  50  

2.2.16.4   Organ  extraction  ...  50  

2.2.16.5   Immunofluorescence  analysis  of  cardiac  sections  ...  51  

2.2.17   Statistical  analysis  ...  51  

3   RESULTS  ...  52  

3.1   Mybpc3-­‐targeted  knock-­‐in:  a  mouse  model  of  HCM  ...  52  

3.2   The  5’-­‐trans-­‐splicing  approach  in  Mybpc3-­‐targeted  KI  mice  ...  54  

3.3   Characterization  of  the  engineered  constructs  ...  56  

3.3.1   Validation  of  PTMs  ...  56  

3.3.2   Validation  of  ‘toxic’  molecules  ...  59  

3.3.3   Validation  of  full-­‐length  WT-­‐Mybpc3  ...  60  

3.4   Evaluation  of  5’-­‐trans-­‐splicing  ex  vivo  ...  61  

3.4.1   Validation  of  AAV6-­‐mediated  GFP  expression  in  isolated  cardiomyoctes  ...  61  

3.4.2   Characterization  of  5’-­‐trans-­‐splicing  at  mRNA  level  ...  62  

3.4.2.1   Detection  of  repaired  Mybpc3  mRNA  ...  63  

3.4.2.2   Evaluation  of  5’-­‐trans-­‐splicing  in  absence  of  the  polyadenylation  signal  in  the  PTM  ...  66  

3.4.2.3   Validation  of  repaired  Mybpc3  mRNA  by  sequencing  ...  68  

3.4.2.4   Semi-­‐quantification  of  repaired  Mybpc3  mRNA  ...  69  

3.4.3   Characterization  of  5’-­‐trans-­‐splicing  at  protein  level  ...  71  

3.4.3.1   Validation  of  potential  translation  of  PTM  and  PTM∆pA  in  HEK293  cells  ...  71  

3.4.3.2   Detection  of  repaired  cMyBP-­‐C  protein  ...  72  

3.4.4   Strategy  to  increase  the  efficiency  of  5’-­‐trans-­‐splicing  ...  77  

3.4.5   Summary:  5’-­‐trans-­‐splicing  partially  corrected  defective  Mybpc3  mRNA  ex  vivo  ...  80  

3.5   Evaluation  of  5’-­‐trans-­‐splicing  in  vivo  ...  82  

3.5.1   Validation  of  AAV9-­‐mediated  GFP  expression  in  the  heart  ...  82  

3.5.2   Evaluation  of  5’-­‐trans-­‐splicing  in  adult  Mybpc3-­‐targeted  KI  mouse  ...  83  

3.5.2.1   Characterization  of  cardiac  function  ...  83  

3.5.2.2   In  vivo  bioluminescence  imaging  ...  85  

3.5.2.3   Characterization  of  5’-­‐trans-­‐splicing  at  mRNA  level  ...  86  

3.5.2.4   Characterization  of  5’-­‐trans-­‐splicing  at  protein  level  ...  88  

3.5.3   Evaluation  of  5’-­‐trans-­‐splicing  in  neonatal  Mybpc3-­‐targeted  KI  mouse  ...  89  

3.5.3.1   Characterization  of  the  cardiac  function  ...  90  

3.5.3.2   Characterization  of  5’-­‐trans-­‐splicing  at  the  mRNA  level  ...  91  

3.5.3.3   Characterization  of  5’-­‐trans-­‐splicing  at  the  protein  level  ...  93  

3.5.4   Summary:  5’-­‐trans-­‐splicing  partially  corrected  defective  Mybpc3  mRNA  in  vivo  ...  94  

3.6   Conventional  gene-­‐based  therapy  in  Mybpc3-­‐targeted  KI  mice  ...  95  

3.6.1   Overexpression  of  WT-­‐Mybpc3  in  isolated  cardiomyocytes  ...  95  

3.6.1.1   Characterization  at  mRNA  level  ...  95  

3.6.1.2   Characterization  at  protein  level  ...  98  

(6)

 

3.6.3   Summary:  WT-­‐Mybpc3  gene  transfer  rescued  protein  haploinsufficiency  and  prevented  

accumulation  of  poison  peptides  in  a  mouse  model  of  HCM  ...  106  

4   DISCUSSION  ...  107  

4.1   Mybpc3-­‐targeted  knock-­‐in  mice  as  a  suitable  model  for  molecular  therapy  ...  108  

4.2   The  combination  of  virus  plus  promoter  plus  delivery  route  for  efficient  cardiac  gene  transfer  ...  110  

4.2.1   Adeno-­‐associated  virus  versus  adenovirus  ...  110  

4.2.2   Choice  of  a  suitable  promoter  ...  113  

4.2.3   Vector  delivery  techniques  in  vivo  ...  114  

4.3   Molecular-­‐based  approaches  to  target  the  cause  of  HCM  ...  115  

4.3.1   RNA-­‐based  therapy  to  prevent  the  accumulation  of  poison  polypeptides?  ...  115  

4.3.2   Conventional  gene  therapy  to  treat  haploinsufficiency  in  HCM?  ...  121  

4.4   Conclusion  -­‐  future  directions:  from  mice  to  men  ...  124  

5   SUMMARY  ...  127  

6   BIBLIOGRAPHY  ...  I  

7   APPENDIX  ...  XV  

7.1   Materials  ...  XV  

7.1.1   Recombinant  adeno-­‐associated  virus  ...  XV  

7.1.2   Recombinant  adenovirus  ...  XV  

7.1.3   Antibodies  ...  XVI  

7.1.3.1   Antibodies  used  for  Western  blot  ...  XVI  

7.1.3.2   Antibodies  used  for  immunofluorescence  staining  ...  XVI  

7.1.4   Bacterial  strains  ...  XVII  

7.1.5   Chemicals  ...  XVII  

7.1.6   Consumable  materials  ...  XIX  

7.1.7   Kits  ...  XX  

7.1.8   Laboratory  equipment  ...  XX  

7.1.9   Restriction  enzymes  ...  XXI  

7.1.10   Oligonucleotides  ...  XXII  

7.1.10.1   Mybpc3  primer  sequences  for  WT-­‐Myppc3  ...  XXII  

7.1.10.2   Mybpc3  primer  sequences  used  for  PTM  coding  domain  ...  XXII  

7.1.10.3   Mybpc3  primer  sequences  used  for  PTM  binding  domain  ...  XXII  

7.1.10.4   Mybpc3  primer  sequences  for  PCR  and  sequencing  ...  XXII  

7.1.10.5   Mybpc3  primer  and  probe  sequences  for  qPCR  ...  XXIII  

7.1.10.6   Sequences  of  the  binding  domains  ...  XXIII  

7.1.10.7   Sequences  of  the  promoters  ...  XXIV  

7.1.10.8   Vectors  ...  XXIV  

7.2   H-­‐  &  P-­‐Phrases  ...  XXV  

7.3   Protein  and  DNA  markers  ...  XXIX  

(7)

 

9   DECLARATION  ...  XXXIV  

10   CURRICULUM  VITAE  ...  XXXVI              

(8)

 

1 Introduction  

"Cardiomyopathies   are   a   heterogeneous   group   of   diseases   of   the   myocardium   associated   with   mechanical   and/or   electrical   dysfunction   that   usually   (but   not   invariably)   exhibit   inappropriate  ventricular  hypertrophy  or  dilation  and  are  due  to  a  variety  of  causes  that  are   frequently   genetic.   Cardiomyopathies   either   are   confined   to   the   heart   or   are   part   of   generalized   systemic   disorders,   often   leading   to   cardiovascular   death   or   progressive   heart   failure-­‐related  disability"  (Maron  et  al.,  2006).  

Worldwide   the   estimated   prevalence   of   all   types   of   cardiomyopathies   is   about   3%   in   the   general  population  (Cecchi  et  al.,  2012).  Cardiomyopathies,  frequently  with  a  genetic  cause,   include   a   variety   of   myocardial   disorders   in   which   the   heart   muscle   is   structurally   and   functionally   abnormal   (Elliott   et   al.,   2008).   The   degree   of   cardiac   dysfunction   ranges   from   lifelong   symptomless   forms   to   major   health   problems,   such   as   progressive   heart   failure,   arrhythmia,  thromboembolism  or  sudden  death  (Franz  et  al.,  2001).  A  majority  of  patients   remains  undiagnosed  or  misdiagnosed  with  more  prevalent  cardiac  conditions  (Cecchi  et  al.,   2012).   Whereas   major   progress   has   been   made   in   improving   the   prognosis   of   affected   patients,  cardiomyopathies  still  remain  a  considerable  challenge  in  the  health  care  system   and  an  economic  burden  across  Europe  and  the  rest  of  the  world.  

Since   the   1950’s   several   definitions,   nomenclatures   and   classification   schemes   have   been   acquired   by   experts.   Cardiomyopathies   are   currently   grouped   by   the   ‘European   Society   of   Cardiology   Working   Group   on   Myocardial   and   Pericardial   Diseases’   into   specific   morphological   and   functional   characteristics   with   sub-­‐classifications   into   familial   and   non-­‐ familial  subset  (Figure  1;  Elliott  et  al.,  2008).  The  four  major  subtypes  of  cardiomyopathies   are   hypertrophic   cardiomyopathy   (HCM),   dilated   cardiomyopathy   (DCM),   arrhythmogenic   right   ventricular   cardiomyopathy   (ARVC)   and   restrictive   cardiomyopathy   (RCM).   The   classification   also   includes   an   unclassified   group   with   no   typical   phenotype.   The   main   characteristics  of  the  different  subtypes  can  be  summarized  as  follows:  

(9)

 

Hypertrophic  cardiomyopathy  is  characterized  by  non-­‐dilated  ventricular  chambers,  but  an  

unexplained  thickening  of  the  left  ventricle,  due  to  an  enlargement  of  cardiomyocytes  and   therefore  the  ventricle  is  less  able  to  relax  and  to  fill  with  blood.  

Dilated  cardiomyopathy  is  a  weakness  in  the  walls  of  the  heart  that  causes  dilation  of  the  

left  ventricle,  compromising  the  heart’s  efficiency  and  increasing  the  risk  of  congestive  heart   failure,  arrhythmias  and  the  formation  of  blood  clots.  

Arrhythmogenic   right   ventricular   cardiomyopathy   occurs   if   the   muscle   tissue   in   the   right  

ventricle  is  replaced  with  fibrofatty  tissue,  which  disrupts  the  heart’s  electrical  signaling  and   causes  ventricular  arrhythmias.  

Restrictive  cardiomyopathy  involves  loss  of  elasticity  of  the  ventricles  due  to  stiff  tissue  that  

prevents   the   ventricles   from   normal   relaxation   and   from   adequately   blood   filling   prior   contraction.  

                 

Figure   1:   Classification   scheme   of   cardiomyopathies  -­‐  European   Society   of   Cardiology.   HCM,   hypertrophic  

cardiomyopathy;  DCM,  dilated  cardiomyopathy;  ARVC,  arrhythmogenic  right  ventricular  cardiomyopathy;  RCM,   restrictive  cardiomyopathy;  idiopathic,  no  identifiable  cause  (Elliott  et  al.,  2008).  

This   long-­‐standing   classification   of   cardiomyopathies   is   a   multi-­‐disciplinary   approach   to   predict   major   complications,   improve   risk   stratification   and   optimize   treatment   in   each   subtype.   However,   many   cardiomyopathies   are   caused   by   a   variety   of   gene   abnormalities   and   are   the   consequence   of   interactions   between   multiple   disease   genes,   unidentified  

(10)

 

For  example,  different  mutations  within  the  same  gene  can  result  in  different  subtypes,  e.g.   mutations  in  MYH7  gene,  encoding  beta-­‐myosin  heavy  chain,  can  cause  either  hypertrophic   or   dilated   cardiomyopathy   (Kamisago   et   al.,   2000).   In   contrast,   the   same   mutation   in   one   distinct   gene   can   arise   with   diverse   cardiac   phenotypes   at   different   ages   within   the   same   family.   The   classification   of   cardiomyopathies   based   on   morphological   and   functional   phenotypes  is  an  essential  tool  for  clinicians  to  manage  these  complex  heart  diseases.  The   list  of  putative  causative  genes  or  non-­‐genetic  causes  responsible  for  the  distinct  phenotype   is  provided  for  each  subtype  in  the  literature  and  is  required  to  be  constantly  updated  by   diverse  expert  committees  (Cecchi  et  al.,  2012).  

This   work   focuses   on   hypertrophic   cardiomyopathy   (HCM),   the   most   common   inherited   cardiovascular   disorder.   It   provides   state   of   the   art   on   HCM,   including   consequences   of   disease-­‐causing   mutations   and   describes   how   targeting   the   molecular   defects   have   given   early  promise  for  potential  new  therapies.  

1.1 Hypertrophic  cardiomyopathy  

Hypertrophic  cardiomyopathy  (HCM)  is  a  common,  but  unexplained  structural  abnormality   of  the  cardiac  muscle  and  the  most  prevalent  cause  of  heart-­‐related  sudden  death  in  young   people.   The   prominent   morphologic   feature   of   HCM   is   a   massive   asymmetrical   left   ventricular  hypertrophy  (LVH),  which  mainly  involves  the  interventricular  septum  (Figure  2).   Clinically   diagnosed   HCM   patients   show   increased   left   ventricular   wall   thicknesses   ranging   from   mild   (15  mm)   to   massive   (˃30  mm)   magnitudes   based   on   echocardiographic   measurements   (Maron,   2002).   The   abnormal   thickening   of   the   left   ventricle   may   result   in   left   ventricular   outflow   tract   obstruction   and   is   associated   with   an   initially   normal   systolic   function,   whereas   the   diastolic   function   is   impaired   (Maron,   2002).   In   fact,   disease   progression  may  lead  to  a  thinning  of  the  left  ventricle  wall  and  an  enlargement  of  the  left   cavity,  which  is  associated  with  reduced  ejection  fraction.  This  significantly  increases  the  risk   of   irreversible   heart   failure   and   unexpected   sudden   cardiac   death,   especially   in   young   individuals  during  exercise  (Maron  et  al.,  2000).  HCM  occurs  mostly  in  the  absence  of  any   other  cardiac  or  systemic  disorders  that  themselves  would  be  able  to  trigger  hypertrophy  in  

(11)

 

remarkable   variability   in   disease   development,   age   of   onset   and   severity   of   symptoms,   showing   both   benign   and   malignant   manifestations   (Richard   et   al.,   2003).   Indeed,   many   patients   display   a   clinical   history   completely   asymptomatic,   whereas   in   other   cases   HCM   leads  to  symptoms  such  as  vertigo,  chest  pain,  syncope,  dyspnea  and  can  turn  into  malignant   arrhythmias  and  progressive  heart  failure  (Gersh  et  al.,  2011).  The  main  histological  features   include   chaotically   oriented   cardiomyocytes   (=myocardial   disarray),   as   well   as   myocyte   hypertrophy  and  increased  interstitial  fibrosis  (Figure  2).    

                                                                       

Figure  2:  Histological  characteristics  of  hypertrophic  cardiomyopathy.  Upper  scheme  shows  hypertrophied  and  

normal   hearts.   The   lower   part   represents   corresponding   histological   section   stained   with   hematoxylin   and   eosin.  Hypertrophied  cardiomyocytes  display  disarray  and  increased  myocardial  fibrosis.  Figure  adapted  from   the  Mayo  clinic  website  (upper  panel)  and  from  Ho  et  al.,  2010  (lower  panel).  

HCM  has  a  prevalence  of  1:500  in  the  general  population  and  occurs  equally  in  both  sexes   (Maron   et   al.,   1995).   It   accounts   for   36%   of   sudden   cardiac   death   among   competitive   athletes   (Maron   et   al.,   1996).   In   the   majority   of   cases   HCM   is   inherited   in   an   autosomal-­‐ dominant  pattern  (Richard  et  al.,  2003).  The  variability  of  the  HCM  phenotype  is  attributed   to  over  1000  mutations  in  at  least  19  different  genes  (Table  1),  which  have  been  identified  as   a  potential  cause  of  the  disease  (Schlossarek  et  al.,  2011,  Friedrich  and  Carrier,  2012).  Most   known   genes   encode   components   of   the   contractile   apparatus,   the   sarcomere,   which   elucidates  HCM  as  a  ‘sarcomeropathy’.  Most  of  the  HCM  patients  are  heterozygous  for  the   mutation  and  3-­‐5%  of  them  carry  two  independent  mutations  at  once  resulting  in  a  more   severe   phenotype   than   patients   with   a   single   mutation   (Richard   et   al.,   2003,   Ingles   et   al.,  

(12)

 

mutation  (Ho  et  al.,  2000).  Mutations  in  the  MYH7  gene  encoding  cardiac  beta-­‐myosin  heavy   chain   and   in   the   MYBPC3   gene   encoding   cardiac   myosin-­‐binding   protein   C   are   the   most   common   ones   and   comprise   more   than   70%   of   the   known   HCM-­‐causing   genetic   defects   (Richard  et  al.,  2006,  Maron  et  al.,  2012).  The  phenotypic  expression  of  MYBPC3  mutations  is   largely   heterogeneous.   In   contrast   to   most   MYH7   mutations,   which   cause   early   onset   and   extensive   LVH   (Watkins   et   al.,   1992),   mutations   in   MYBPC3   have   been   first   shown   to   be   associated  with  delayed  onset,  incomplete  penetrance  and  mild  hypertrophy  (Charron  et  al.,   1998,  Niimura  et  al.,  1998).  However,  severe  cases  with  a  poor  outcome  and  high  sudden   cardiac  death  risk  profile  have  been  reported  as  well  (Erdmann  et  al.,  2001,  Oliva-­‐Sandoval   et   al.,   2010).   Especially   childhood   HCM   is   often   associated   with   an   extreme   LVH   and   the   presence   of   sinus   and   supraventricular   tachycardia   (Morita   et   al.,   2008,   El-­‐Saiedi   et   al.,   2013).   Besides   the   heterogeneity   of   the   causal   genes   and   mutations,   there   is   a   high   variability  in  the  phenotype  expression  of  HCM  (Marian,  2002).  Even  within  single  families,   affected   individuals   with   identical   causal   mutations   can   show   significant   variability   in   the   disease   penetrance,   age   of   onset   and   clinical   manifestation.   In   fact,   20-­‐30%   of   HCM   mutation-­‐positive  patients  do  not  reveal  any  cardiac  phenotype  (Richard  et  al.,  2003).  This   suggests  that  phenotypes  are  not  exclusively  gene  or  mutation  specific  and  distinct  modifiers   must   exist,   such   as   environmental   factors,   epigenetic   signaling,   microRNAs,   gene   polymorphisms   or   posttranslational   modifications,   which   modulate   disease-­‐causing   mechanisms  (Marian,  2002,  Richard  et  al.,  2006,  Schlossarek  et  al.,  2011).  For  example,  it  has   been   reported   that   a   polymorphism   in   the   promoter   of   the   CALM3   gene,   encoding   calmodulin  III,  has  modifying  impact  on  the  HCM  phenotype  by  affecting  the  expression  level   of  CALM3  and  consequently  the  calcium  handling  and  development  of  LVH  (Friedrich  et  al.,   2009).  Other  potential  modifier  genes  are  polymorphisms  in  genes  encoding  angiotensin  I-­‐ converting  enzyme  (Tesson  et  al.,  1997),  as  well  as  the  AT1  and  AT2  receptors  (Osterop  et   al.,  1998,  Deinum  et  al.,  2001)  

   

(13)

 

Table  1:  Summary  of  sarcomeric  gene  mutations  implicated  in  HCM  (Schlossarek  et  al.,  2011).  

                                               

1.2 The  sarcomere  –  fundamental  contractile  unit  

The  basic  functional  and  structural  unit  of  contractile  muscles  is  the  sarcomere  with  a  highly   ordered  assembly.  The  cardiac  sarcomere  is  formed  by  three  types  of  myofilaments,  thin  and   thick  filaments  and  titin  (Figure  3).  Their  organization  within  the  sarcomere  is  defined  by  two   neighboring   Z-­‐lines,   which   enclose   the   I-­‐Band   (thin   filaments)   and   the   A-­‐band   (thick   filaments)   with   the   M-­‐line   and   C-­‐zone.   The   thin   filaments   contain   α-­‐cardiac   actin,   α-­‐ tropomyosin   and   the   troponin   complex.   The   latter   is   composed   of   troponin   C   (binds   calcium),   troponin   I   (inhibits   contraction)   and   troponin   T   (binds   to   α-­‐tropomyosin).   Actin   filaments  interact  with  α-­‐actinin  in  the  Z-­‐line  and  extent  to  the  A-­‐band.  The  thick  filaments   consist  of  mainly  myosin  and  are  located  in  the  A-­‐band.  Myosin  consists  of  two  heavy  chains   (α/ß   myosin   heavy   chain,   MHC)   and   four   light   chains   (MLC;   two   essential   light   chains   and   two   regulatory   light   chains).   The   third   filament   titin   stabilizes   the   thick   filaments   and   connects  them  to  the  Z-­‐line.    

(14)

 

                                                                                                                               

Figure   3:   Scheme   of   the   cardiac   sarcomere   representing   the   localization   of   the   cMyBP-­‐C   protein.   The   thin  

filaments   are   composed   of   α-­‐cardiac   actin,   α-­‐tropomyosin   and   the   troponin   complex.   Thick   filaments   are   located   in   the   A-­‐band   and   consist   of   myosin   with   α/ß   myosin   heavy   chains,   essential   and   regulatory   myosin   light  chains.  cMyBP-­‐C  is  a  thick  filament-­‐associated  protein,  aside  titin,  which  itself  is  considered  as  the  elastic   component   of   the   sarcomere.   The   I-­‐band,   A-­‐band,   with   the   C-­‐zones   and   M-­‐line   are   confined   by   two   Z-­‐lines   (Schlossarek  et  al.,  2011).  

The  interaction  of  actin  and  myosin  is  responsible  for  the  muscle  contraction  and  is  triggered   by   an   increase   in   cytosolic   calcium   through   voltage-­‐dependent   L-­‐type   calcium   channels   during   action   potential.   This   inward   calcium   flux   activates   the   ryanodine   receptors   in   the   sarcoplasmic   reticulum   and   induces   calcium   release.   In   the   resting   cardiomyocyte,   α-­‐ tropomyosin   blocks   the   myosin-­‐binding   site   on   actin   and   is   anchored   by   troponin   T   and   I.   When  intracellular  calcium  binds  to  troponin  C,  it  initiates  a  conformational  change  in  the   troponin   complex   and   releases   α-­‐tropomyosin   from   the   myosin-­‐binding   site   on   actin,   allowing   the   interaction   of   myosin   and   actin.   Upon   adenosine   triphosphate   hydrolysis,   myosin  is  subjected  to  a  series  of  conformational  changes  resulting  in  the  motion  of  the  thick   along   the   thin   filaments.   Several   calcium   pumps,   such   as   sarcoplasmic   reticulum   calcium-­‐ ATPase   (SERCA)   and   sodium-­‐calcium   exchanger   contribute   to   the   removal   of   calcium   to   return   to   the   relaxed   state.   Finally,   α-­‐tropomyosin   is   again   locked   in   its   actin-­‐blocking   position  by  troponin  T  and  I.  

(15)

 

1.3 Cardiac  myosin-­‐binding  protein  C  

1.3.1 Structural  role  

The   cardiac   myosin-­‐binding   protein   C   (cMyBP-­‐C)   is   a   multidomain   protein   of   the   thick   filaments  and  is  located  in  doublets  in  the  C-­‐zone  of  the  A-­‐band  of  the  sarcomere  (Figure  3).   cMyBP-­‐C  protein  bundles  the  thick  filaments  transversally  over  nine  clear  stripes  in  each  half   A-­‐band  (Luther  et  al.,  2008).  The  MYBPC3  gene  encoding  the  cMyBP-­‐C  protein,  is  located  on   the  human  chromosome  11p11.2  (Gautel  et  al.,  1995).  Its  complete  structure  and  sequence   was  established  in  1997  (Carrier  et  al.,  1997).  The  gene  comprises  more  than  21,000  bp  and   contains   35   exons,   of   which   34   are   coding.   The   transcript   is   translated   into   a   150-­‐kDa   protein,   which   exhibits   fundamental   structural   and   regulatory   functions   (Winegrad,   1999,   Flashman   et   al.,   2004,   de   Tombe,   2006,   Granzier   and   Campbell,   2006,   Schlossarek   et   al.,   2011,  Sadayappan  and  Tombe,  2012).    

Beside  the  cardiac  isoform  of  MyBP-­‐C  there  are  two  others,  which  have  been  identified  in   adult   cross-­‐striated   muscle,   each   encoded   by   different   genes:   the   slow-­‐skeletal   (MYBPC1)   and   the   fast-­‐skeletal   (MYBPC2)   isoforms.   All   three   isoforms   belong   to   the   intracellular   immunoglobulin   superfamily   and   show   a   conserved   pattern   in   their   main   structure.   Generally,  MyBP-­‐C  protein  consists  of  seven  immunoglobulin  (Ig-­‐1-­‐like)  and  three  fibronectin   (FN-­‐3)   domains   (C1-­‐C10).   The   cardiac   isoform   is   exclusively   expressed   in   the   mammalian   heart  (Gautel  et  al.,  1995,  Fougerousse  et  al.,  1998)  and  differs  from  the  other  isoforms  in   distinct  characteristics.  It  contains  an  extra  Ig-­‐1-­‐like  domain  (C0)  located  at  the  N-­‐terminus,   four  phosphorylation  sites  located  in  the  MyBP-­‐C  motif,  which  is  a  conserved  105-­‐residues   linker  between  C1  and  C2,  a  28-­‐amino  acid  insertion  in  the  C5  domain  and  a  proline-­‐alanine-­‐ rich   extension   between   C0   and   C1   (Figure   4;   Gautel   et   al.,   1995,   Flashman   et   al.,   2004,   Oakley  et  al.,  2004).  

(16)

 

 

Figure  4:  Schematic  representation  of  MYBPC3  gene,  MYBPC3  mRNA  and  structure  of  cMyBP-­‐C  protein.  In  the  

upper  part  the  organization  of  the  MYBPC3  gene  5’  to  3’  is  shown  with  localization  of  exons  indicated  by  boxes.   In   the   middle   panel   the   mRNA   of   the   joined   exons   after   cis-­‐splicing   is   displayed.   cMyBP-­‐C   protein   domains   involved  in  sarcomeric  protein  interactions  are  indicated  by  arrows.  Abbreviations:  P,  phosphorylation  site;  S2,   myosin  subfragment  S2;  LMM,  light  meromyosin  (Schlossarek  et  al.,  2011).  

The   cardiac   isoform   of   MyBP-­‐C   protein   interacts   with   different   sarcomeric   proteins   via   specific   motifs   or   domains   (Figure   4).   The   MyBP-­‐C   motif   is   known   to   interact   with   the   subfragment  S2  of  myosin  (Gruen  and  Gautel,  1999),  whereas  the  C10  domain  binds  to  light   meromyosin   (Okagaki   et   al.,   1993)   and   the   C8-­‐C10   domains   to   titin   (Freiburg   and   Gautel,   1996).   Potential   actin-­‐binding   sites   were   described   at   the   proline-­‐alanine-­‐rich   extension   between   C0   and   C1   domains   (Squire   et   al.,   2003),   at   the   C0   domain   (Kulikovskaya   et   al.,   2003)  and  at  the  C1-­‐C2  domains  (Razumova  et  al.,  2006).  Interactions  of  cMyBP-­‐C  protein   with  myosin  and  titin  are  important  for  an  optimal  arrangement  of  the  sarcomere  and  the   complex  binding  enables  to  form  a  very  stable  structure  (Flashman  et  al.,  2004).  The  precise   cMyBP-­‐C   protein   incorporation   into   the   thick   filament   setup   still   remains   unclear,   but   the   ‘trimeric  collar’  model  is  preferred.  This  model  suggests  the  trimerization  of  three  cMyBP-­‐C   molecules   in   a   staggered   parallel   orientation   around   the   backbone   of   the   thick   filaments  

(17)

 

(Moolman-­‐Smook  et  al.,  2002).  This  ‘trimeric  collar’  formation  leads  to  interactions  between   C5  and  C8,  C6  and  C9,  C7  and  C10  of  two  cMyBP-­‐C  molecules  and  C10  of  the  third  one  binds   to   the   myosin   rod   (Figure   5).   The   distinct   C0-­‐C4   domains   interact   with   the   myosin   subfragment  S2  and  actin  thin  filaments.  

1.3.2 Functional  role  

cMyBP-­‐C  protein  regulates  the  cross-­‐bridge  cycling  via  titin,  myosin  and  actin  interactions,   the   myofilament   calcium   sensitivity   and   relaxation   of   the   sarcomere.   Its   regulatory   role   is   mediated   through   four   phosphorylation   sites   located   in   the   MyBP-­‐C   motif   (Barefield   and   Sadayappan,   2010).   Phosphorylation   occurs   in   response   to   beta-­‐adrenergic   stimulation   by   cAMP-­‐dependent   protein   kinase   (PKA)   (Gautel   et   al.,   1995),   by   calcium/calmodulin-­‐ dependent   kinase   II   (CaMKII)   in   a   calcium-­‐dependent   manner   (McClellan   et   al.,   2001),   by   protein   kinase   C   ε   (PKCε)   (Kooij   et   al.,   2010),   by   protein   kinase   D   (PKD)   (Bardswell   et   al.,   2010)  and  by  90-­‐kDa  ribosomal  S6  kinase  (RSK)  (Cuello  et  al.,  2011).    Upon  phosphorylation   of   cMyBP-­‐C   protein,   the   binding   to   actin   and   myosin   subfragment   S2   is   abolished,   which   increases   the   cross-­‐bridging   between   myosin   and   actin   and   therefore   the   force   of   contraction  (Figure  5)  (Flashman  et  al.,  2004).  It  is  therefore  essential  for  a  normal  cardiac   function  and  was  shown  to  be  cardioprotective  (Sadayappan  et  al.,  2005,  Sadayappan  et  al.,   2006).   It   has   been   reported   that   the   level   of   phosphorylated   cMyBP-­‐C   protein   is   low   in   human   and   experimental   models   of   heart   failure   (El-­‐Armouche   et   al.,   2007)   as   well   as   in   myocardial  tissue  of  HCM  patients  (van  Dijk  et  al.,  2009,  Marston  et  al.,  2012,  van  Dijk  et  al.,   2012).  Moreover,  cMyBP-­‐C  protein  is  crucial  to  allow  a  complete  diastolic  relaxation  of  the   sarcomere  at  low  intracellular  calcium  concentrations  by  inhibiting  the  interaction  of  actin   and   myosin   through   reversible   binding   to   the   myosin   subfragment   S2  (Kulikovskaya   et   al.,   2003,  Pohlmann  et  al.,  2007).  Residual  cross-­‐bridge  cycling  in  diastole,  incomplete  relaxation   and   increased   calcium   sensitivity   of   the   myofilaments   may   lead   to   diastolic   dysfunction,   hypercontractility   and   increased   energy   usage   (Crilley   et   al.,   2003,   Javadpour   et   al.,   2003,   Keller  et  al.,  2004,  Pohlmann  et  al.,  2007).  The  detailed  role  of  cMyBP-­‐C  protein  is  still  not   fully   understood,   but   alteration   in   protein   level   and   phosphorylation   may   lead   to   severe   cardiac  dysfunction  and  structural  abnormalities,  in  particular  to  cardiac  hypertrophy.  

(18)

 

 

Figure   5:   Sarcomeric   organization   of   cMyBP-­‐C   protein   in   the   dephosphorylated   and   phosphorylated   state.  

Three   cMyBP-­‐C   molecules   trimerize   around   the   thick   filament   backbone   of   light-­‐meromyosin   (Myosin   LMM)   and   titin.   The   N-­‐terminal   C1-­‐M-­‐C2   domains   in   the   dephosphorylated   state   are   tightly   attached   to   myosin-­‐S2   and   actin   via   reversible   linkage   to   prevent   the   cross-­‐bridge   formation.   Upon   phosphorylation,   the   motif   (M   domain)  releases  the  interaction  with  myosin-­‐S2  and  actin  and  results  in  an  attachment  of  the  myosin  heads   (myosin-­‐S1)  to  the  thin  filaments,  which  promotes  strong  actin-­‐myosin  interaction  (Schlossarek  et  al.,  2011).  

1.3.3 MYBPC3  mutations  and  pathophysiological  mechanisms  

More  than  460  different  mutations  associated  with  HCM  have  been  found  in  the  MYBPC3   gene  (source  HGMD;  http://www.hgmd.org/).  However,  the  expression  of  the  mutations  and   therefore  the  consequences  at  mRNA  and  protein  levels  are  not  known  for  most  of  them.   The   majority   (˃70%)   of   MYBPC3   mutations   are   frameshift   or   nonsense   and   predicted   to   cause   altered   splicing   (Richard   et   al.,   2006,   Carrier   et   al.,   2010,   Marian,   2010).   Nonsense   mutations  directly  introduce  a  premature  termination  codon  (PTC)  in  the  transcribed  mRNA,   whereas   frameshift   mutations   are   the   consequence   of   point   mutations,   insertions   or   deletions,   which   lead   to   a   PTC   downstream   of   the   mutation   in   the   transcript.   C-­‐terminal   truncated  cMyBP-­‐C  proteins,  likely  lacking  myosin-­‐  and/or  titin-­‐binding  sites  (Carrier  et  al.,   1997,   Richard   et   al.,   2006)   have   never   been   detected   in   myocardial   tissue   of   patients   carrying  MYBPC3  mutations  at  the  heterozygous  state  (Rottbauer  et  al.,  1997,  Moolman  et   al.,   2000,   Marston   et   al.,   2009,   van   Dijk   et   al.,   2009,   Marston   et   al.,   2012,   van   Dijk   et   al.,   2012).  The  full-­‐length  cMyBP-­‐C  protein  level  in  myectomy  samples  from  HCM  patients  with   frameshift  MYBPC3  mutations  is  20-­‐30%  lower  than  in  normal  hearts  (Marston  et  al.,  2009,   van   Dijk   et   al.,   2009,   Marston,   2011).   Apparently,   the   gene   product   from   the   remaining  

(19)

 

functional  wild-­‐type  allele  cannot  fully  compensate  for  the  defect  transcript  from  the  mutant   allele.  

Therefore,   the   molecular   mechanisms   of   MYBPC3-­‐associated   HCM   mutations   and   their   impact  at  mRNA  and  protein  levels  are  not  completely  conclusive.  The  reduced  amount  of   normal  cMyBP-­‐C  protein  is  one  argument  that  haploinsufficiency  is  likely  the  HCM  disease   mechanism.   The   expression   of   the   functional   wild-­‐type   allele   in   the   case   of   heterozygous   mutation   does   not   produce   a   sufficient   amount   of   protein   to   maintain   the   wild-­‐type   phenotype.  Haploinsufficiency  has  been  also  reported  in  several  animal  models  of  MYBPC3-­‐ associated   HCM.   The   Mybpc3-­‐tageted   KI   mouse   model   express   low   levels   of   cMyBP-­‐C   protein,  80%  in  the  heterozygous  state  and  only  10%  in  the  homozygous  state  (Vignier  et  al.,   2009).  The  Maine  Coon  cats,  carrying  a  natural  MYBPC3  missense  mutation  reveal  69%  lower   level  of  cMyBP-­‐C  protein  in  the  heterozygous  state  and  88%  in  the  homozygous  state  than  in   wild-­‐type  controls  (Meurs  et  al.,  2005).  Since  sarcomere  stoichiometry  is  tightly  regulated,   reduced  levels  of  cMyBP-­‐C  protein  could  imbalance  the  assembly  of  the  thick  filament  and   therefore   affect   sarcomeric   structure   and   function,   leading   to   contractile   deficits.   Furthermore,   the   cMyBP-­‐C   protein   deficiency   is   likely   associated   with   higher   myofilament   calcium  sensitivity  due  to  the  altered  protein  expression  and/or  its  phosphorylation,  which  

represents  a  consistent  abnormality  in  HCM  (Harris  et  al.,  2002,  Cazorla  et  al.,  2006,  van  Dijk   et   al.,   2009).   The   myofilament   activation   results   in   basal   cardiomyocyte   hypercontractility   and   excessive   energy   usage   (Watkins   et   al.,   2011).   The   resultant   energy   deficiency   and   altered  intracellular  calcium  handling  combined  with  activation  of  signaling  pathways  likely   contribute   to   the   anatomic   (hypertrophy,   myocardial   disarray   and   fibrosis)   and   functional   (diastolic  dysfunction)  characteristics  of  HCM  (Ashrafian  et  al.,  2011).  

Additionally,  the  presence  of  truncated  mutant  cMyBP-­‐C  proteins  could  abnormally  alter  the   sarcomere  organization  and  potentially  provoke  damage  in  cardiomyocytes  acting  as  poison   peptides.   However,   despite   detectable   amounts   of   mutant   mRNA   (25-­‐45%   of   wild-­‐type),   truncated  proteins  have  not  been  detected  in  myocardial  tissue  of  HCM  patients  (Moolman   et  al.,  2000,  Marston  et  al.,  2009,  van  Dijk  et  al.,  2009,  Marston  et  al.,  2012,  van  Dijk  et  al.,   2012).  This  suggests  that  their  expression  is  regulated  at  mRNA  and  protein  levels.  In  the  cell   the   major   quality   control   systems   to   regulate   expression   of   nonsense   and   frameshift  

(20)

 

(UPS)   and   autophagy-­‐lysosomal   pathway   (ALP)   (Sarikas   et   al.,   2005,   Carrier   et   al.,   2010,   Schlossarek  et  al.,  2011).  The  NMD  degrades  nonsense  transcripts  at  mRNA  level  and  UPS   and/or   the   ALP   rapidly   eliminate   misfolded   or   mutant   proteins   (Vignier   et   al.,   2009).   The   permanent  degradation  of  mutant  proteins  by  the  UPS  and/or  ALP  to  protect  the  cell  from   their   deleterious   effects   may   lead   to   an   impairment   of   the   proteolytic   control   systems.   In   heterozygous   Mybpc3-­‐targeted   KI   and   KO   mouse   models,   it   has   been   reported   that   adrenergic  stress  or  aging  resulted  in  saturation  of  the  systems  (Schlossarek  et  al.,  2012a,   Schlossarek  et  al.,  2012b).  The  potential  accumulation  of  poison  polypeptides  was  sufficient   to  alter  cell  homeostasis  and  trigger  the  disease  progression  (Sarikas  et  al.,  2005,  Bahrudin  et   al.,   2008).   Impairment   of   the   UPS   has   been   reported   in   human   HCM   patients   as   well   (Predmore  et  al.,  2010).    

1.4 Current  status  of  treatment  and  potential  novel  therapies  for  

HCM  

As   mentioned   before,   HCM   is   a   very   complex   disease   with   heterogeneous   genetic,   morphologic,  functional  and  clinical  manifestation.  Although  it  is  a  life-­‐threatening  disease,   no  curative  treatment  exists  up  to  date  reversing  the  cardiac  hypertrophy  and  dysfunction   and/or  preventing  sudden  cardiac  death  (Carrier  et  al.,  2010,  Schlossarek  et  al.,  2011,  Frey  et   al.,   2012,   Spoladore   et   al.,   2012).   The   standard   clinical   management   is   basically   empiric.   Current  drug-­‐based  strategies  are  partially  capable  to  relieve  the  HCM-­‐associated  symptoms   and   slow   down   the   disease   progression   but   none   of   them   induces   regression   of   cardiac   hypertrophy   or   fibrosis   or   targets   the   genetic   cause.   Therefore,   potential   innovative   therapies   against   fundamental   pathophysiological   mechanisms   in   patients   with   inherited   HCM   mutations   urgently   need   to   be   established   and   may   display   a   new   paradigm   for   personalized  medicine.  

(21)

 

1.4.1 Drug  and  surgical  treatments    

Current  drug-­‐based  interventions  of  HCM  mainly  focus  on  symptomatic  management  and  on   the   control   of   ventricular   outflow   obstruction   and   arrhythmias   in   order   to   improve   the   patient’s  quality  of  life.  Although  clinical  guidelines  for  diagnosis  and  treatment  of  HCM  exist   (Maron  et  al.,  2003,  Gersh  et  al.,  2011),  the  benefit  of  pharmacological  intervention  is  not   evidence-­‐based   (Spoladore   et   al.,   2012).   Until   2012,   only   less   than   50   pharmacological   studies  have  been  performed,  most  of  them  involving  small  and  non-­‐randomized  groups  of   patients   and   none   of   them   has   prospectively   addressed   the   long-­‐term   outcome   (Figure   6)   (Spoladore  et  al.,  2012).  Long-­‐standing  drugs,  such  as  beta-­‐adrenergic  inhibitors  and  L-­‐type   calcium  channel  blockers  have  been  applied  in  the  majority  of  studies.    

 

Figure  6:  Number  of  HCM  pharmacological  studies  (left)  and  number  of  patients  included  (right)  based  on  the  

application  of  the  indicated  drugs.  ACE-­‐I,  angiotensin  converting  enzyme  inhibitors;  ARBs,  angiotensin  receptor   blockers  (Spoladore  et  al.,  2012).  

The   beneficial   effect   of   beta-­‐blockers,   such   as   propranolol,   is   due   to   improvement   of   ventricular   relaxation,   increase   in   time   for   diastolic   filling   and   reduction   of   excitability,   especially  in  patients  with  exercise-­‐induced  symptoms,  left  ventricular  outflow  obstruction   and  chest  pain  (Marian,  2009).  The  negative  inotropic  response  of  beta-­‐blocker  additionally   decreases   the   myocardial   oxygen   demand   and   outflow   gradient   during   exercise.   L-­‐type   calcium  channel  blocker,  such  as  verapamil  and  diltiazem  are  employed  to  lower  the  heart   rate,   reduce   excitability   and   lengthen   the   diastolic   filling   period   in   non-­‐obstructive   HCM   (Spirito  et  al.,  1997,  Marian,  2009).  Besides  the  first  line  therapy  of  cardiac  arrhythmias  with   beta-­‐blocker,   amiodarone   is   commonly   used   for   the   treatment   of   atrial   and   ventricular   arrhythmias   in   HCM   patients   (McKenna   et   al.,   1985,   Cecchi   et   al.,   1998).   Disopyramide,   a  

(22)

 

class  I   antiarrhythmic   drug   with   a   negative   inotropic   effect,   has   been   successfully   used   in   combination   with   beta-­‐blocker   to   attenuate   symptoms   in   patients   with   left   ventricular   outflow   obstruction   (Sherrid   et   al.,   2005).   Very   limited   data   exist   for   drugs   targeting   the   altered   energy   homeostasis   (perhexeline)   (Abozguia   et   al.,   2010),   the   impaired   calcium   cycling   and   sensitivity   of   the   myofilaments   (blebbistatin)   (Baudenbacher   et   al.,   2008),   the   increased   fibrosis   and   left   ventricular   remodeling   (aldosterone   antagonists,   such   as   spironolactone  or  angiotensin  II  receptor  blockers,  such  as  losartan,  irbesartan)  (Lim  et  al.,   2001,  de  Resende  et  al.,  2006).  If  symptoms  persist  despite  drug-­‐based  treatment,  surgical   septal  myectomy  or  percutaneous  septal  ablation  with  ethanol  could  be  employed  in  order   to   mechanically   diminish   the   outflow   tract   obstruction,   to   attenuate   the   severity   of   symptoms  and  reduce  the  risk  of  sudden  cardiac  death  (Maron,  2002,  Elliott  and  McKenna,   2004,   Ball   et   al.,   2011).   The   insertion   of   an   implantable   cardioverter-­‐defibrillator   (ICD)   as   prophylactic  intervention  may  be  indicated  in  individuals  who  have  survived  cardiac  arrest  or   those  with  an  increased  susceptibility  for  atrial  and  ventricular  arrhythmias  (Ho,  2010).  ICD   has   been   reported   to   be   effective   and   life-­‐saving   in   relevant   patients   (Begley   et   al.,   2003,   Maron  et  al.,  2007).  Patients  with  end-­‐stage  HCM,  which  is  characterized  by  left  ventricular   remodeling   with   progressive   wall   thinning,   cavity   enlargement   and   systolic   dysfunction,   should   be   medicated   with   appropriate   drugs   for   heart   failure   including   diuretics   and   angiotensin   converting   enzyme   inhibitors   (ACE-­‐I)   (Spirito   et   al.,   1987,   Spirito   et   al.,   1997).   Patients   with   severe   heart   failure   ultimately   require   heart   transplantation   (Shirani   et   al.,   1993).  

     

(23)

 

1.4.2 Molecular-­‐based  treatments  

The  concept  of  genetic  medicine  in  order  to  repair  or  modify  inherited  disorders  was  initially   evoked  in  the  late  1960s  with  the  development  of  virus-­‐based  transformation  of  mammalian   cells  and  the  progress  in  recombinant  DNA  techniques  (Friedmann,  1992,  Sheridan,  2011).   Molecular-­‐based   therapy   involves   the   use   of   DNA   or   RNA   for   the   treatment,   curing   or   prevention  of  disorders.  It  generally  aims  at  the  correction  of  key  pathologies,  which  are  out   of   reach   for   conventional   drugs.   Depending   on   the   nature   of   the   disease,   gene-­‐based   approaches   can   be   applied   to   deliver   a   functional,   therapeutic   gene   to   substitute   the   defective  or  missing  endogenous  gene  analogue  (conventional  gene  therapy)  or  to  reduce   the   level   of   defective   transcripts   using   innovative   RNA-­‐based   approaches,   such   as   spliceosome-­‐mediated   RNA   trans-­‐splicing   (SMaRT),   exon   skipping   and   exon   inclusion.   The   idea   is   to   utilize   site-­‐directed   gene   or   RNA   editing   strategies   to   target   critical   molecular   changes   in   the   endogenous   gene   or   pre-­‐mRNA   to   anticipate   for   causal   HCM   therapy.   Mutations,  causing  aberrant  splicing  represent  approximately  about  one  third  of  all  disease-­‐ causing  mutations  (Lim  et  al.,  2011,  Sterne-­‐Weiler  et  al.,  2011).  The  splicing  mechanism,  as   an  early  step  in  gene  expression,  is  an  attractive  intervention  point  for  therapeutic  purposes,   which  does  not  alter  the  genome.  During  the  last  10  years  approaches  targeting  mutant  pre-­‐ mRNA   in   a   splice-­‐switching   manner   have   been   intensively   studied   in   the   field   of   neuromuscular   genetic   disorders   (Le   Roy   et   al.,   2009,   Havens   et   al.,   2013).   These   studies   used   sophisticated   molecular   tools,   including   pre-­‐trans-­‐splicing   molecules   and   modified   antisense  oligonucleotides.    

1.4.2.1 Spliceosome-­‐mediated  RNA  trans-­‐splicing    

Spliceosome-­‐mediated  RNA  trans-­‐splicing  (SMaRT)  is  a  promising  therapeutic  approach  for   genetic  disorders.  It  is  a  post-­‐transcriptional  process  occurring  during  mRNA  maturation,  and   is  defined  as  a  splicing  reaction  between  two  independently  transcribed  RNA  molecules,  a   target  endogenous  mutant  pre-­‐mRNA  and  a  therapeutic  pre-­‐trans-­‐splicing  molecule  (PTM)   (Wally  et  al.,  2012).  Trans-­‐splicing  is  a  rare  process,  but  naturally  occurring  mechanism  and   was   first   discovered   in   lower   eukaryotes,   such   as   trypanosomes   (Sutton   and   Boothroyd,  

(24)

 

important   process   to   achieve   functional   diversity   (Dorn   and   Krauss,   2003)   and   has   been   recently   reported   in   human   leukocytes   (Chiu   et   al.,   2008).   Trans-­‐splicing   is   very   attractive   since  it  occurs  by  using  the  endogenous  spliceosome  machinery  in  the  nucleus,  it  does  not   require  the  introduction  of  the  complete  gene  and  it  is  restricted  to  those  cells  expressing   the   target   pre-­‐mRNA.   Briefly,   eukaryotic   post-­‐transcriptional   processing   of   pre-­‐mRNA   includes  5’-­‐capping,  3’-­‐polyadenylation  and  cis-­‐splicing  reactions,  which  occur  in  the  nucleus.   It   has   been   intensively   studied   that   addition   of   7-­‐methylguanosine   to   the   5’-­‐end   and   3’-­‐ polyadenylation   of   the   transcript   are   essential   for   proper   splicing,   RNA   transport   and   protection   from   degradation.   Additionally,   these   processes   enhance   translation   of   mRNA   (Sachs  and  Wahle,  1993,  Wahle  and  Keller,  1996,  Colgan  and  Manley,  1997,  Cowling,  2010).   In   general,   cis-­‐splicing   is   a   multistep   process   to   remove   introns   from   the   pre-­‐mRNA   and   ligate  remaining  exons  to  form  a  single  continuous  mRNA  molecule.  The  cis-­‐splicing  reaction   is  catalyzed  by  a  protein  complex  called  the  spliceosome,  consisting  of  several  proteins  and   small   nuclear   RNA   molecules   that   recognize   splice   sites   within   the   pre-­‐mRNA   sequence.   These  universally  conserved  sequences  in  eukaryotic  pre-­‐mRNA  are  the  5’-­‐  or  donor-­‐splice   site   (GU)   and   3’-­‐   or   acceptor-­‐splice   site   (AG)   at   exon-­‐intron   boundaries,   the   conserved   branch  point  (A)  and  a  pyrimidine-­‐rich  region  (Py)n  just  upstream  of  the  3'-­‐splice  site  (Figure  

7;  Pagani  and  Baralle,  2004).  

                                                                                 

Figure   7:   Classical   cis-­‐splicing   reaction   and   essential   intronic   splicing   signals.   The   image   shows   conserved  

nucleotides  at  the  exon-­‐intron  boundaries:  5’-­‐splice  site  (GU),  branch  point  (A),  polypyrimidine  tract  (Py)n  and  

3’-­‐splice   site   (AG).   Splicing   involves   two-­‐step   transesterification   reactions.   In   the   first   step   the   2’-­‐hydroxyl-­‐ group  of  a  specific  branch  point  nucleotide  within  the  intron  performs  a  nucleophilic  attack  on  the  phosphate   (p)   of   the   5'-­‐splice   site   forming   the   lariat   structure.   The   second   transesterification   reaction   involves   the   3’-­‐ hydroxyl   group   of   the   released   5’-­‐exon   and   the   phosphate   (p)   at   the   3’-­‐splice   site,   which   releases   the   lariat   structure  and  ligates  the  two  exons  (Pagani  and  Baralle,  2004).  

(25)

 

SMaRT  is  an  emerging  technology  carried  out  by  the  endogenous  spliceosome  complex.  The   engineered  PTM  is  exogenously  delivered  to  the  nucleus  and  after  successful  transcription  it   hybridizes  with  the  target  pre-­‐mRNA  to  finally  generate  a  chimeric  mRNA  molecule  (Figure   8).   The   PTM   is   designed   to   recode   a   specific   part   of   the   mRNA   by   suppressing   cis-­‐splicing   while  enhancing  trans-­‐splicing  in  a  competitive  manner;  therefore,  as  the  level  of  repaired   mRNA  increases  the  level  of  native  mRNA  should  decrease.  This  observation  suggests  SMaRT   as   a   general   approach   for   correction   of   a   targeted   pre-­‐mRNA   transcript   (Puttaraju   et   al.,   1999).  

                                                               

Figure  8:  Schematic  illustration  of  SMaRT  strategy.  Pre-­‐trans-­‐splicing  molecule  (PTM)  is  dispensed  to  the  cell  

and  transcribed  in  the  nucleus.  The  transcript  targets  the  pre-­‐mRNA  of  the  gene  of  interest  (GOI)  and  produces   repaired  mRNA  and  protein.  This  process  is  in  competition  with  cis-­‐splicing,  which  gives  rise  to  the  endogenous   mRNA  and  proteins  (Mearini  et  al.,  2013).  

PTMs   consist   of   three   main   domains:   i)   a   coding   domain   containing   the   wild-­‐type   exonic   information;   ii)   a   set   of   splice   signals   containing   5´-­‐   and/or   3´-­‐splice   sites,   including   the   branch   point   sequence;   iii)   a   binding   domain   complementary   to   intronic   sequences   of   the   target  pre-­‐mRNA  by  base  pairing.  The  latter  is  essential  to  put  the  splicing  sites  of  both  pre-­‐ mRNA   and   PTM   close   to   each   other   in   order   to   induce   trans-­‐splicing.   The   length   of   the   binding  domain  defines  the  specificity  of  the  PTM  and  has  a  profound  impact  on  the  trans-­‐ splicing  activity  (Mansfield  et  al.,  2004).  Usually  the  size  of  the  binding  domain  comprises  70-­‐ 150  nucleotides  (Puttaraju  et  al.,  1999),  although  no  formula  for  its  length  has  been  defined  

Referenzen

ÄHNLICHE DOKUMENTE

Fig. 4 Total and neutralization antibody responses in the immunized mice. Both vaccine constructs elicited significant IgG1 anti- body responses in immunized BALB/c mice. Although

In the case described above, the surgical approach allowed precise and complete resection of the hypertrophic septum and was able to abolish (and not only to reduce) the LVOT

The RV free wall peak systolic and diastolic velocities, strain rate values and strain at basal and mid segment are significantly reduced in patients with HCM in comparison

To investigate whether the higher UPS activity in KO is associated with increased ATP usage and therefore increased AMP-kinase activity, the concentration of phosphorylated

Here we describe the generation of two induced pluripotent stem cell (iPSC) clones from a HCM patient, heterozygous for the p.Arg723Gly (c.2169C > G) mutation in the MYH7

In WT cardiomyocytes, stimulation with 30 nM ISO normally led to a 2-fold increase of the contraction amplitude at 1-Hz pacing frequency (fig. An additional increase

After four weeks of allergen cessation eosinophilic inflammation, goblet cell hyperplasia and collagen deposition were resolved, full resolution of lymphocyte

It is commonly caused by mutations in the MYBPC3 gene encoding cardiac myosin-binding protein C (cMyBP-C). About 61% of MYBPC3 mutations are frameshift or