• No results found

Neurofibromatosis Type 1: From Cognitive Profile to Future Therapeutic Targets

N/A
N/A
Protected

Academic year: 2021

Share "Neurofibromatosis Type 1: From Cognitive Profile to Future Therapeutic Targets"

Copied!
31
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Neurofibromatosis Type 1: From Cognitive

Profile to Future Therapeutic Targets

Paul L.C. Feyen - 6323391

Supervisor: Dr. Omrani, Azar

Co-Assesor: Dr. Werkman, Taco

Friday October 19

th

Literature Thesis - MsC in Brain and Cognitive

Sciences, Neuroscience Track, University of

(2)

Content Overview

 

  Introduction  

  NF1  Cognitive  Phenotype  

  The  NF1  genotype  and  how  this  relates  to  the  phenotype  

    Neurofibromin  is  expressed  as  several  variants  encoded  by  the  NF1    

    gene  

 

    The  relationship  between  mutations  of  NF1  and  NF1  pathogenesis  

 

  Molecular  and  Cellular  View  on  NF1-­related  Cognitive  Problems  

 

    Upstream  modulators  of  neurofibromin  

 

    Downstream  signaling  pathways  of  neurofibromin  

 

 

Animal  Models  of  NF1:  Insights  on  Etiology

   

    Behavioral  tasks  assigned  to  specific  brain  areas  highlight  the  parallels    

    of  NF1  mice  and  human  phenotype  

 

    Increased  inhibition  and  altered  dopamine  system  underlie  NF1      

    cognitive  phenotype    

 

Treatment of NF1 in Human Patients

Neurodevelopmental  disorders  can  be  treated,  even  in  adulthood

    Phase  1  clinical  trials  highlight  the  complexity  of  NF1  pathogenesis  

 

    Modulation  of  ion  channel  function  as  potential  therapies  for  NF1  

 

    Benefits  of  a  behavioral  component  in  the  treatment  of  NF1  

 

(3)

Introduction    

With   an   incidence   of   1   in   3000   individuals,   Neurofibromatosis   type   1   (NF1)   is   among   the   most   common   autosomal   dominant   diseases   (Huson   et   al,   1989).   NF1,   also   known   as   von   Recklinghause   disease,   is   characterized   by   the   presence   of   neurofibromas,  optic  pathway  glioma,  Lisch  nodules,  axillary  freckling  and  café-­‐au-­‐ lait  macules  (NIH,  1988).    NF1  symptomology  develops  with  age,  but  a  diagnosis  can   be  made  before  the  age  of  six  (Huson  et  al  1989).  In  both  child  and  adult  patients  of   NF1,   cognitive   impairments   are   frequently   observed   (North   et   al,   1997;   Ozonoff,   1999;   Hyman   et   al,   2005).   NF1   patients   have   a   low-­‐average   IQ   (Kayl   and   Moore,   2000),  and  display  problems  in  specific  cognitive  domains  like  memory,  visuospatial   skills,  language,  executive  functioning,  and  attention  (Hyman  et  al  2005).  In  children   with  NF1  these  deficits  are  the  most  common  complication  affecting  quality  of  life   (Hyman   et   al.   2005).   Overall,   the   cognitive   problems   and   associated   learning   disabilities   pose   one   of   the   most   significant   sources   of   lifetime   morbidity   for   patients  (North  et  al.,  1997).    Given  the  impact  which  cognitive  deficits  have  on  the   lives  of  NF1  patients,  this  literature  study  aims  to  bring  together  clinical  and  animal   model   data   to   develop   a   clear   view   of   the   molecular   underpinnings   of   cognitive   deficits  in  NF1,  in  the  aim  of  defining  reliable  molecular  targets  for  future  research   and  potentially  for  pharmaceutical  therapy.    

NF1  Cognitive  Phenotype  

The   cognitive   phenotype   of   NF1   is   quite   diverse,   and   is   characterized   by   domain   specific  cognitive  impairment  rather  than  a  global  mental  deficit.    This  is  reflected  in   the  slightly  downward  shift  of  IQ  in  the  NF1  population  which  lies  within  the  normal   distribution,   as   well   as   in   the   fact   that   several   other   measures   sensitive   to   global   cognitive  impairments  do  not  distinguish  at  all  between  NF1  patients  and  unaffected   siblings  (Kayl  and  Moore  2000).  This  shift  in  IQ  does  however  result  in  a  two-­‐fold   increase  for  risk  of  mental  retardation  (North  et  al,  1997).      On  the  other  hand,  more   than   80%   of   individuals   affected   by   NF1   show   impairments   in   specific   cognitive   domains  as  assessed  by  neuropsychological  evaluations  (Hyman  et  al,  2005,  Krab  et   al.   2008).     There   is   also   a   high   correlation   between   NF1   and   attention   deficit   disorder   (Hofman   et   al.1994,   Hyman   et   al.   2005,   Mautner   et   al.   2002).     Learning   disabilities   are   diagnosed   in   up   to   65%   of   individuals   affected   by   NF1   (Rosser   &   Packer  2003),  and  individuals  with  NF1  are  fourfold  more  likely  to  require  special   education  (Krab  et  al.  2008a).    

 

The   fourth   edition   of   the   Diagnostic   and   Statistical   Manual   of   Mental   Disorders   (DMS-­‐IV)   define   a   specific   learning   difficulty   as   an   academic   achievement   that   is   more  than  twice  the  standard  deviation  on  an  individual’s  IQ,  which  can  occur  in  one   of   5   major   categories   including   reading   disorders,   mathematical   disorders,   disorders   of   written   expression,   nonverbal   learning   disorders,   and   other   learning   disorders   which   lead   to   academic   underachievement,   However,   rather   than   fitting   one  such  description,  different  types  of  learning  disabilities  have  been  observed  in  

(4)

individuals  with  NF1  (Levine  et  al  2006).  Early  reports  suggested  that  the  cognitive   profile   of   NF1   resembled   that   of   children   with   non-­‐verbal   learning   disorders   (Ozonoff   1999).   Not   only   do   NF1   patients   show   consistently   poor   performance   in   tests  of  visuospatial  functioning  and  spatial  learning  (North,  2000;  Kayl  and  Moore,   2000),   these   individuals   also   present   other   components   of   non-­‐verbal   learning   disorders  like  increased  impulsiveness,  poor  organizational  skills  and  a  poor  social   cue  perception  (North  et  al,  1997;  North,  2000).  Yet  NF1  cannot  be  strictly  classified   into  one  of  the  categories  defined  in  DMS-­‐IV.  Literacy  based  learning  impairments   are   also   prevalent,   as   are   mathematical   ones   in   NF1   (Descheermaeker   et   al   2005,   Hyman   et   al,   2006).     Interestingly,   after   controlling   for   IQ,   children   with   NF1   no   longer  show  a  higher  incidence  of  math  based  learning  problems  (Kayl  and  Moore,   2000).   On   the   other   hand,   deficits   in   reading   and   writing   have   repeatedly   been   found   to   be   more   severe   in   NF1   patients   than   predicted   from   IQ   scores,   and   is   associated   with   specific   deficits   in   expressive   and   receptive   language,   vocabulary,   visual  naming,  and  phonologic  awareness  (Hyman  et  al  2005).    This  occurrence  of   multiple  learning  disorders  in  NF1  is  indicative  of  a  more  fundamental  impairment   to  learning  (Shilyansky  2010b).  Now  the  aim  has  shifted  to  understanding  the  parse   behavioral  symptoms  as  more  specific  and  stable  phenotypes,  or  'endophenotypes'   with  a  clear  anatomical,  molecular,  and  genetic  connection.  In  this  light,  the  hurdles   to  social  and  academic  performance  can  be  seen  to  stem  from  learning  disabilities   and   cognitive   impairment   in   five   specific   areas,   namely,   reading/vocabulary,   attention,  executive  function  and  planning,  visuospatial  function,  and  motor  deficits   (Hyman   et   al,   2005).     These   in   turn   are   likely   caused   by   one   or   more   general   mechanism/s  affecting  multiple  brain  systems  (Shilyansky  2010b).  

 

For   one,   the   poor   performance   of   NF1   patients   on   visuospatial,   vocabulary,   and   reading  comprehension  suggest  a  dysfunction  of  hippocampus-­‐dependent  memory   (Shilyansky   2010b).   Damage   to   hippocampus   results   in   specific   and   profound   cognitive   defects.   Such   patients   often   have   difficulty   learning   and   retrieving   new   vocabulary  (Verfaelli  et  al,  2000),  and  they  are  often  reported  to  have  severe  deficits   in   spatial   memory   and   spatial   navigation   tasks   (Burgess   et   al,   2002;   Squire   et   al   2004).   Interestingly,   these   symptoms   are   also   commonly   presented   in   individuals   with   NF1.   For   example,   patients   display   poor   performance   in   tests   of   visuospatial   functioning   and   spatial   learning   (Kelly,   2004).     As   such,   abnormal   hippocampal   function  in  NF1  patients  could  contribute  to  their  deficits  in  learning  and  memory.   Though   the   parallels   are   striking   and   suggest   role   for   this   brain   area   in   NF1,   dysfunction  in  this  region  alone  cannot  explain  all  the  impaired  cognitive  functions   presented  in  NF1.    

 

Attention  deficit  disorders  are  another  cognitive  hallmark  of  NF1  patients,  and  are   diagnosed  as  a  co-­‐morbid  condition  in  30%  to  50%  of  patients  (Hofman  et  al,  1994;   Mautner   et   al,   2002;   Hyman   et   al,   2005;   Pavol   et   al,   2006).   The   impact   of   this   condition  is  noticeable  not  only  in  NF1  children’s  academic  performance  but  also  in   their  social  development  (Barton  and  North,  2004).  About  40%  of  NF1  children  seen   as  academic  underachievers  show  no  deficit  in  evaluations  of  learning  disabilities,  

(5)

and   attention   deficit   disorders   may   well   contribute   to   their   poor   academic   performance  (Kayl  and  Moore,  2000).  NF1  patients  also  face  social  problems  and  are   reported   to   appear   socially   awkward   and   withdrawn.   These   problems   with   interpersonal   skills   are   thought   to   stem   from   a   decreased   attention   to   social   cues   (Kayl   and   Moore   2000).     Attention   deficit   disorders   have   been   associated   with   fronto-­‐striatal   and   fronto-­‐cortical   abnormalities,   two   brain   systems   that   are   also   associated   with   normal   attention   processing   (Fox   et   al   2006;   Cubillo   et   al   2012).     Thus  abnormalities  in  these  same  brains  areas  may  contribute  to  the  difficulty  with   planning  and  organization  seen  in  NF1  patients.      

 

Attention,  planning,  and  organization  are  components  of  executive  functioning,  and   the  functioning  of  this  multi-­‐component  cognitive  faculty  is  tightly  reliant  on  frontal   cortex  (Stuss  and  Levine  2002).    Of  note,  patients  suffering  lesions  to  this  area  share   symptoms   with   NF1   patients   including   limitations   in   planning,   organization,   and   control  of  attention  (Powel  and  Voeller  2004).  The  severity  of  attention  deficits  is   however   independent   of   the   degree   of   organizational   and   planning   deficits,   suggesting  that  NF1  affects  each  component  of  executive  function  independently  of   the  other  (Hyman  et  al  2005).  Importantly  though,  these  components  of  executive   function   all   rely   on   frontal   brain   networks,   and   dysfunction   in   this   brain   area   has   even  been  observed  by  functional  imaging  of  NF1  patients.  For  instance,  one  study   showed  hypoactivation  of  prefrontal  cortex  during  a  rhyming  task  (Billingsley  et  al.,   2003).  Here  participants  are  visually  presented  two  nonsense  words  and  asked  to   assess   whether   the   set   rhymes   or   not.   Interestingly,   hypoactivation   of   the   corticostriatal  pathway  has  also  been  observed  during  a  delayed  response  task.  In   this  visuospatial  working  memory  task  subjects  must  fixate  on  the  central  point  of  a   screen   and   remember   the   location   of   stimuli   that   are   presented   peripherally.   NF1patients   show   an   impaired   working   memory   performance   on   this   task   that   is   associated  with  hypoactivation  of  corticostriatal  networks  (Shilyanski  et  al  2010a).      

Impairments   in   visuospatial   cognition   occur   in   over   50   percent   of   NF1   patients   (Hyman  et  al  2005;  Zöller  et  al  1997).    In  fact  these  deficits  are  so  common  that  NF1   children   can   be   distinguished   from   the   normal   population   by   evaluating   their   performance   on   multiple   visuospatial   tasks   (Schrimsher   et   al   2003).     The   NF1   population  performance  is  most  frequently  low  in  the  Judgment  of  Line  Orientation   task   (Hyman   et   al   2005),   in   which   patients   must   discriminate   visual   stimuli   presented  at  varying  degrees  of  rotation.  The  task  is  thought  to  rely  on  an  extensive   brain  system  comprising  the  prefrontal,  parietal  and  visual  cortices  (Ng  et  al  2000).   More   recently   spatial   working   memory   and   visuospatial   processing   was   also   associated  to  activity  in  distinct  cerebellar  regions  (Lee  et  al  2005).    The  degree  of   contribution  that  these  different  areas  make  to  the  visuospatial  cognition  phenotype   seen   in   NF1   may   well   vary   from   area   to   area,   and   importantly   from   patient   to   patient.    Though  similar  brain  areas  appear  to  be  involved  in  different  components   of   the   NF1   cognitive   phenotype,   the   occurrence   of   visuospatial   impairments,   just   like   problems   with   planning   and   organization,   is   independent   of   the   presence   of   attention  disorders  (Schrimsher  et  al  2003).    

(6)

 

Motor   deficits   are   another   component   of   the   NF1   cognitive   phenotype,   and   the   impairments   suggest   a   role   for   the   cerebellum   in   the   disorder.   Children   with   NF1   are  frequently  described  as  clumsy  (Cnossen  et  al  1998,  Dilts  et  al  1996)  and  tend  to   reach  developmental  milestones  at  a  late  time  (Chapman  et  al,  1996).  The  specific   deficits   that   have   been   observed   include   many   fine   and   gross   motor   functioning   impairment  such  as  problems  with  balance  and  gait,  and  with  tasks  like  writing  that   require   a   visuomotor   integration   and   fine   motor   skills   (Hyman   et   al,   2003).   Additionally,   NF1   children   were   found   to   be   impaired   in   several   motor   learning   paradigms   that   are   sensitive   to   cerebellar   dysfunction,   though   they   also   showed   normal   performance   in   other   cerebellum-­‐dependent   tasks   (Krab   et   al   2011).   Further   suspicion   of   the   cerebellum   comes   from   the   strikingly   high   level   of   expression   of   neurofibromin,   the   protein   encoded   by   the   NF1   gene,   in   cerebellar   purkinje  cells  (Gutmann  et  al,  1995).  

 

Participation   from   an   early   age   in   remedial   teaching   have   been   associated   with   enhanced   long   term   outcomes   on   both   cognitive   and   quality   of   life   measures   in   individuals   with   Nf1-­‐associated   learning   disorders   (Hyman   et   al,   2006).   Unfortunately,   remedial   teaching   is   not   always   available,   pressing   the   need   for   alternative   treatment   options.   For   one,   attention   impairments   can   be   treated   pharmaceuticaly.   NF1   patients   presenting   with   co-­‐morbid   Attention-­‐Deficit-­‐ Hyperactivity-­‐Disorder   (ADHD)   are   known   to   benefit   from   methylphenidate   treatment  (Mautner  et  al,  2002).  The  development  of  pharmaceutical  treatments  for   the  other  NF1  associated  cognitive  symptoms  may  greatly  benefit  the  quality  of  life   of  individuals  for  whom  remedial  teaching  is  not  available  or  effective.    

 

Summarizing,   the   range   of   symptoms   presented   by   NF1   patients   lends   to   a   presentation   of   a   highly   diverse   and   variable   disease   phenotype.       In   the   central   nervous   system,   the   effects   of   NF1   on   cognitive   functioning   can   be   categorized   as   those   affecting   reading/vocabulary,   attention,   executive   function   and   planning,   visuospatial  function,  and  motor  function  (Hyman  et  al,  2005).    In  turn,  dysfunction   in   brain   areas   underlying   these   cognitive   processes   may   well   contribute   to   the   impairments  observed  in  NF1  patients.    The  overall  diversity  of  the  phenotypes  and   its   many   clinical   presentations   can   for   a   large   part   be   understood   by   gaining   an   understanding  on  the  genetics  of  the  pathology.            

     

The  NF1  genotype-­phenotype  relationship    

Underlying  neurofibromatosis  are  mutations  to  the  Nf1  gene  (Wallace  et  al  1990).   The  resulting  hereditary  syndrome  afflicts  a  variety  of  tissues  and  causes  the  myriad   of   clinical   features   described   above.   Though   the   disease   is   fully   penetrant,   individuals   show   large   differences   in   the   presentation   of   their   phenotype,   even   within  families  (Huson  1994).  One  major  factor  contributes  critically  to  the  diverse   clinical   manifestation   of   NF1   and   that   is   genetic   background.   Evidence   has   long   demonstrated  a  role  for  trait-­‐specific  modifier  genes  in  modulating  the  phenotypic  

(7)

expression  of  individuals  with  NF1  (Easton  et  al  1993).  NF1  gene  expression  itself  is   also  a  highly  regulated  and  dynamic  process  involving  several  post-­‐transcriptional   modifications   such   as   the   alternative   splicing   mediated   by   the   gene’s   various   splicing   sites   (Cappione   et   al,   1997).   The   result   is   that   neurofibromin's   final   biochemical  and  functional  properties  are  regulated  at  multiple  levels,  both  during   transcription   and   translation   and   by   its   modifier-­‐gene   dictated   environment   following  expression.  Modulation  of  NF1  also  occurs  at  different  levels,  though  the   impact  of  such  regulation  on  NF1  pathogenesis  is  less  clear.  For  instance,  differences   in   allelic   expression   levels,   suggest   that   epigenetic   mechanisms   modulate   neurofibromin  protein  levels  and  thereby  physiological  functionality  (Stephens  et  al   1992).     More   recently,   the   regulation   of   NF1   expression   by   microRNA   has   been   described   in   cultures   of   primary   hippompampal   neurons   (Paschou   and   Doxakis   2012).    

 

Neurofibromin  is  expressed  as  several  variants  encoded  by  the  NF1  gene  

 

The  neurofibromin  protein  is  encoded  by  the  NF1  gene  located  at  17q11.2,  a  gene   whose   mass   accounts   for   more   than   300kb   of   human   chromosome   17   (Li   et   al,   1995).     The   length   of   the   gene   is   composed   of   60   exons   that   are   expressed   as   alternatively   spliced   transcripts.   Of   the   8   known   variants,   several   are   preserved   across   species,   and   6   are   expressed   in   humans.   The   isoforms   are   distinct   in   their   temporal  and  tissue  expression  pattern,  and  further  differ  in  functional  properties   (Skuse  and  Cappione  1997).  

 

The  naming  of  NF1  variants  is  not  standardized  and  reflects  the  order  of  discovery   and  species  in  which  it  was  discovered.  More  recently,  the  variants  have  also  been   referred   to   on   the   basis   of   exon   inclusion/exclusion.   The   naturally   occurring   alternative   splicing   of   the   NF1   pre-­‐mRNA   results   in   differential   inclusion   of   three   alternative  exons  (23a,  and  48a,  9a/9br)  (Danglot  et  al,  1995;  Andersen  et  al,  1993;   Gutman   et   al,   1993).   Type   1   and   Type   2   NF1   exclude   and   include   exon   23a,   respectively  (Andersen,  1993).  The  sequence  of  NF1  Type  1  shows  a  high  degree  of   homology  to  the  mammalian  GTPase  activating  proteins  and  is  referred  to  as  NF1's   guanosine   triphosphatase-­‐activating   protein   (GAP)   domain   (Xu   et   al   1990a).   This   domain   modulates   the   inhibitory   action   of   neurofibromin   on   the   intracellular   signalling   molecule,   RAS   (Xu   et   al   1990a),   and   has   been   implicated   in   NF1   pathogenesis  in  humans  (Klose  et  al  1998).  The  inclusion  of  exon  23a  in  NF1  Type  2   adds   63   bases   within   the   GAP   domain,   and   the   ensuing   conformational   change   weakens  the  ability  of  NF1  to  regulate  RAS  (Uchida  et  al  1992,  Andersen  et  al,  1993).     Both  Type  1  and  Type  2  isoforms  are  expressed  ubiquitously,  though  expression  is   enriched   in   neurons,   Schwann   cells,   oligodendrocytes,   astrocytes,   leukocytes   and   adrenal  medulla  (DeClue  1991,  Daston  1992b,  Gutman  1999).  

 

NF1  Type  3  and  Type  4  are  defined  by  their  inclusion  of  exon  48a  and  exon  23a.  The   Type  3  isoform,  which  includes  exon48a,  is  largely  expressed  in  cardiac  and  skeletal   muscle  (Gutmann  et  al,  1995).  Type  4,  which  includes  exon  23a  in  addition  to  exon  

(8)

48a,   presents   a   similar   expression   profile.   Exon   48a   encodes   18   amino   acids   inserted  in  carboxyl  terminus  of  neurofibromin,  and  therefore  does  no  effect  on  the   GAP  domain  (Skuse  and  Cappione  1997),  but  a  role  has  been  suggested  for  this  exon   in  the  development  and  differentiation  of  heart  and  skeletal  muscles  (Gutmann  et  al   1993,  Gutmann  et  al,  1995,  Skuse  and  Cappione,  1997).    

 

Lastly,  the  inclusion  of  alternative  exon  9a/9br  in  NF1  results  in  the  addition  of  10   amino  acids  to  the  amino  terminal  region  of  the  transcript  (Danglot  et  al,  1995).  The   presence   of   exon   9a/9br   does   not   affect   the   GAP   function   of   neurofibromin.   Its   expression   seems   to   be   restricted   to   the   central   nervous   system,   though   it   shows   low   expression   in   NF1-­‐associated   brain   tumors   (Gutman   et   al,   1999   &   Geist   &   Gutman,   1996).   Interestingly,   the   presence   of   exon   9a/9br   on   a   protein   level   is   limited  to  neurons  of  the  forebrain  (Gutmann  et  al,  1999),  suggesting  a  role  for  the   isoform  in  regulating  neuronal  function.      

 

The  Relationship  between  mutations  of  NF1  and  NF1  pathogenesis  

 

Convention   has   it   that   as   a   single   gene   disorder,   NF1   has   enormous   potential   for   elucidating   gene–brain–behavior   connections.   This   is   based   on   the   fact   that   such   relationships   are   many   times   more   difficult   and   problematic   to   establish   in   multigenic  disorders  that  are  associated  with  mechanisms  and  phenotypes  that  are   far  more  diverse  and  complex  (Acosta  et  al,  2004).  What  has  become  clear  though  is   that  'single-­‐gene'  disorders  can  be  quite  complex  and  as  explained  in  the  following  2   subsections,   that   the   NF1   disorder   is   not   as   'monogenic'   as   once   thought.   What   is   clear   is   that   NF1   deficits   ultimately   manifest   in   a   wide   variety   of   clinical   features   with  large  differences  in  individual  phenotypic  severity.    

 

Early   studies   showed   that   neurofibromin   expression   is   essential   in   mammals.   Genetically   engineered   mice   modified   to   have   a   constitutive   deletion   of   the   NF1   gene   face   embryonic   lethality   between   embryonic   days   12.5   and   13.5   due   to   cardiovascular   defects   (Lakkis   et   al   1999).     Importantly,   NF1   and   its   downstream   targets   are   highly   conserved   among   different   species,   including   mice   and   humans   (Bernards  et  al  1993,  Hajra  et  al  1994).  Accordingly,  homozygous  mutations  are  also   lethal   in   humans,   and   individuals   with   the   NF1   disorder   carry   heterozygous   mutations    (Friedman,  1999).    

 

Because  of  its  large  size,  the  NF1  gene  is  particularly  susceptible  to  mutations.  The   NF1  gene  has  one  of  the  highest  mutation  rates  described  for  any  human  gene,  and   as   a   consequence   some   30   to   50%   of   NF1   cases   represent   de   novo   mutations   (Riccardi  et  al  1992,  Huson  et  al  1994).  NF1  mutations  affect  one  allele  at  the  DNA,   mRNA,   or   protein   level   and   this   can   result   in   various   effects   on   the   functions   of   neurofibromin.   70%   of   Nf1   Patients   have   mutations   that   lead   to   truncated,   non-­‐ functional   version   of   the   neurofibromin   protein   (Shen   et   al   1996,   Thomson   et   al   2002).   A   recent   comprehensive   mutation   analysis   showed   that   27%   of   NF1   mutations  affect  premRNA  splicing  (Messiaen  2008).  In  contrast  to  the  abundance  of  

(9)

data  available  on  mutations  and  their  implications  on  a  molecular  level,  it  remains   difficult   to   identify   clear   phenotype-­‐genotype   relationship   due   to   the   broad   mutational  spectrum  of  NF1  and  the  heterogeneity  of  symptom  expression  (Szudek   et   al   2002).   Despite   of   these   difficulties,   two   strong   relationships   have   appeared,   though  these  are  derived  from  NF1  microdeletions  cases  that  are  associated  with  a   stronger   NF1   phenotype.   (Descheemaeker   et   al   2004;   Upadhyaya   et   al   2007).   The   first   is   a   3bp   deletion   outside   of   the   GAP   that   leads   to   notably   mild   clinical   phenotype   and   may   also   lead   to   a   lower   frequency   of   learning   disabilities   (Upadhyaya  et  al  2007).  The  second  involves  a  microdeletion  region  of  17q11  that   includes  RNF135  gene,  a  candidate  gene  for  the  overgrowth,  facial  dimorphism,  and   possibly  the  more  severe  learning  disabilities  of  NF1  microdeletic  patients  (Douglas   et   al   2007).   Clinically   presented   mutations   have   also   associated   GAP   dysfunction   with  NF1  pathogenesis.  In  one  specifically  interesting  case,  a  missense  mutation  of   arginine1276   into   proline   was   found   in   a   female   member   of   a   family   with   multi-­‐ symptomatic  NF1  phenotypes  that  includes  malignant  schwannomas.  The  mutation   that  was  discovered  neither  impairs  secondary  or  tertiary  structure  of  the  protein,   nor  does  it  largely  affect  expression  levels  or  RAS  binding  of  neurofibromin.  Instead   the  mutation  of  the  arginine  finger  of  the  NF1  GAP  abolishes  the  enzymatic  activity   of  GAP  critical  for  NF1  cellular  and  molecular  function,  implicating  GAP  activity  as  a   critical  element  of  NF1  pathogenesis  (Klose  et  al  1998).  Despite  the  fact  that  NF-­‐1  is   a  single-­‐gene  disease,  it  manifests  with  a  variable  expressivity  of  symptoms  within   families,  even  if  individuals  are  affected  by  the  same  NF1  mutation  (Sabbagh  et  al   2009).    

 

As  alluded  to  above,  a  significant  factor  contributing  to  the  variable  expressivity  of   cognitive   and   somatic   symptoms   in   the   NF1   populations   is   the   contribution   of   modifying  genes.  Indeed  a  majority  of  the  variability  of  NF1  symptom  expression  in   mouse  models  and  NF1  patients  can  be  accounted  for  by  inherited  modifiers  (Easton   et   al   1993;   Sabbagh   et   al   2009).   Modifying   genes   are   genes   whose   aberrant   functioning   as   proteins   are   of   no   pathological   consequence   in   a   wild-­‐type   background,   but   exacerbate,   or   increase   the   severity   of   disease   symptoms   in   the   presence  of  mutant  NF1.    One  study  examined  750  NF1  patients  from  275  families  to   analyze  phenotype  correlations  that  included  neurofibromas,  cafe  au  lait  spots,  and   learning  disabilities  (Sabbagh  et  al  2009).  Of  the  12  traits  that  were  investigated,  11   showed   a   strong   genetic   component   with   no   apparent   influence   of   the   NF1   mutations   that   were   identified   in   the   families.   In   other   words,   the   study   strongly   supports   a   model   whereby   the   inheritance   of   modifier   genes   modulates   NF1   symptom  expressivity  more  than  mutations  in  the  NF1  gene  itself.  These  modifiers   can  be  thought  to  interact  directly  with  the  gene  or  to  modulate  expression  levels  or   tissue  expression  pattern.  Alternatively  the  modifier  genes  could  also  be  responsible   for  regulation  of  NF1  mediated  signaling  at  the  molecular  or  cellular  level.    Studies   have  found  evidence  for  both  these  and  several  other  mechanisms  being  utilized  by   putative  NF1  modifier  genes.  

(10)

Molecular   and   Cellular   View   on   NF1-­Related   Cognitive  

Problems  

 

As  expected  from  the  wide  range  of  symptoms  and  the  variability  in  the  severity  of   expression   in   the   NF1   population,   neurofibromin   is   a   highly   dynamic   regulator   of   multiple   molecular   pathways   across   various   types   of   tissues   and   species.     The   neurofibromin  protein  contains  a  notorious  RASGAP-­‐  related  domain,  or  GAP,  which   spans   some   350   amino   acids   of   this   rather   large   2818   amino   acid   long   protein   encoded  by  the  NF1  gene  (Xu  et  al  1990a).  The  consequent  GAP  mediated  regulation   of  RAS  activity  is  believed  to  account  for  the  tumor  suppressor  activity  of  Nf1  (Xu  et   al  1990a,  Xu  et  al  1990b,  Jacks  et  al  1994,  Brannan  et  al  1994).    Though  the  GAP  is   the   most   well   researched   and   characterized   Nf1   domain,   several   more   functional   domains  have  been  identified  in  the  NF1  gene  and  protein.  Importantly  among  them   is   an   adenylyl   cyclase–activating   domain   (Tong   et   al.,   2002).   Additional   domains   have   been   assigned   roles   in   various   cellular   processes   that   include   cell   adhesion,   melanosome   transport,   promotion   of   dendritic   spine   formation,   and   of   neurite   outgrowth   (Hsue   et   al,   2012).   One   NF1   domain,   the   CSRD   or   cysteine-­‐and   serine-­‐ rich  domain,  is  thought  to  be  critical  for  the  association  of  neurofibromin  with  actin   (Gregory  1993).  Lastly,  a  domain  encoded  by  exons  28  to  33  is  thought  to  mediate   the  formation  of  a  complex  between  neurofibromin  and  caveolin-­‐1  (Boyanapalli  et   al   2006).   Caveolin   proteins   are   found   uniquely   in   morphologically   distinct   invaginations  at  the  plasma  membrane,  where  they  function  as  scaffolding  protein   that   bind   several   receptors,   signaling   molecules,   and   adaptor   proteins   (Rajendran   and  Simons,  2005;  Williams  and  Lisanti  2004).  The  relationships  between  most  Nf1   domains   and   NF1   associated   cognitive   impairments   remain   to   be   elucidated.   However,   certain   domains   have   been   well   characterized.   Knowledge   on   the   molecular   pathways   in   which   they   allow   NF1   to   participate   continues   to   advance   our  understanding  of  NF1  pathogenesis,  and  paint  a  picture  of  Nf1  as  an  important   contributor  to  the  regulation  of  neuronal  function.    

 

Upstream  modulators  of  neurofibromin  

 

Though  it  is  not  yet  fully  clear  which  receptors  are  upstream  of  Neurofibromin,  the   current   line   of   evidence   has   demonstrated   a   strong   role   for   growth   factors   as   modulators  of  Nf1  activity.  For  instance  protein  kinase  C  (PKC)  has  been  observed  to   quickly   inhibit   Nf1   in   response   to   growth   factors   whose   pathways   utilize   the   tyrosine   kinase   receptors   (Bernards   &   Settleman,   2005).   Neurofibromin   loss   of   function  can  lead  to  a  de-­‐coupling  of  RAS  signalling  from  extracellular  triggers.  This   was  demonstrated  by  the  finding  that  sensory  neurons  with  a  heterozygous  loss  of   function  of  Nf1  no  longer  required  the  extracellular  growth  factor  BDNF  to  survive   and   mature   (Vogel   et   al   1995).     Likely   underlying   this   cellular   effect   is   the   up-­‐ regulation  of  the  RAS  pathway,  which  is  disinhibited  following  loss  of  NF1  function   (Vogel  et  al  1995).    

(11)

Some   insight   on   upstream   modulators   of   NF1   has   come   from   the   study   of   closely   related  pathologies.  For  instance  Legius  Syndrome,  a  neurodevelopmental  disorder   characterized  by  a  cognitive  and  somatic  phenotype  similar  to  that  of  NF1,  results   from  mutations  in  the  SPRED1  gene  (Stowe  et  al  2012).  Like  NF1,  Legius  Syndrome   is   associated   with   deregulated   RAS   signaling.   Interestingly,   neurofibromin   is   a   Spred1-­‐interacting  protein  that  is  necessary  for  Spred1's  inhibitory  function  on  the   RAS  pathway  (Stowe  et  al  2012).  This  effect  is  mediated  by  the  binding  of  Spred1  to   NF1,  which  induces  the  localization  of  NF1  to  the  plasma  membrane,  where  NF1  can   act  to  down-­‐regulate  RAS  signaling  (Stowe  et  al  2012).    

 

Downstream  signaling  pathways  of  neurofibromin    

 

The  NF1  gene  encoding  neurofibromin  is  expressed  in  a  wide  array  of  cell  types  in   the  body,  though  protein  levels  are  most  abundant  in  central  nervous  system  cells   including   neurons   (Daston   et   al   1992a;   Daston   et   al   1992b).   This   tissue-­‐specific   enrichment  of  NF1  is  indicative  of  a  role  for  the  protein  in  central  nervous  system   functions.  Indeed  three  down-­‐stream  targets  of  Nf1  signalling,  RAS,  cAMP,  and  VCP,   position   neurofibromin   as   a   modulator   of   processes   involved   in   the   regulation   of   neuronal   morphogenesis   and   synaptic   plasticity.   These   functions   of   NF1   help   explain  the  role  that  its  mutant  variants  may  have  in  the  cognitive  impairments  and   learning  disabilities  experienced  by  NF1  patients.    

 

The   RAS   proto-­‐oncogene   is   a   small   GTP-­‐binding   signalling   protein   that   cycles   between   the   inactive   GDP-­‐bound   state   and   the   active   GTP-­‐bound   state.   The   GAP   domain  of  NF1  increases  the  endogenous  GTP  hydrolyzing  activity  of  RAS,  thereby   promoting   the   GDP-­‐bound   state   of   RAS.   In   this   way,   neurofibromin   acts   as   a   negative   regulator   of   RAS   signaling   (Weiss   et   al,   1999).   This   effect   of   NF1   on   RAS   inhibits   critical   intracellular   signalling   cascades   like   the   Extracellular   signal   Regulated  Kinase  (RAS/ERK)  pathway  (Habib  et  al  2008),  and  the  Phosphoinositide   3-­‐kinase   (RAS-­‐PI3K)   /   mammalian   target   of   rapamycin   (MTOR)   pathway   (Tohda,   2006).   RAS-­‐ERK   signalling   modulates   synaptic   plasticity   by   regulating   both   presynaptic  and  postsynaptic  mechanisms.  The  presynaptic  RAS-­‐ERK  pathway  has   been  ascribed  a  role  in  the  control  of  neurotransmitter  release,  which  is  realized  by   ERK  phosphorylation  of  synapsin-­‐1  (Tyler  et  al  2002,  Kushner  et  al  2005).    On  the   post-­‐synaptic   side   the   RAS-­‐ERK   pathway   functions   as   an   important   signal   integrator,  being  activated  by  calcium  influx,  BDNF  binding  to  TRKB,  and  by  indirect   cAMP-­‐PKA   dependent   pathway.   The   post-­‐synaptic   RAS-­‐ERK   signalling   has   been   reported  to  be  involved  in  AMPA  receptor  dynamics  (Zhu  et  al,  2002)  and  regulation   of   protein   transcription   and   translation   events   including   CREB,   or   cAMP   response   element-­‐binding   (Thomas,   2004),   both   of   which   are   processes   implicated   in   long   term   potentiation   (LTP).   LTP   serves   as   an   in   vitro   measure   of   synaptic   plasticity,   which  in  turn  is  the  cellular  mechanism  believed  to  underlie  learning  and  memory   (Kandel,   2001).   Similarly,   MTOR   pathway   has   been   implicated   in   the   protein   synthesis-­‐dependent  phase  of  LTP  (Tang  et  al  2002).    

(12)

 In  addition  to  down-­‐regulating  the  RAS  pathway,  NF1  has  also  been  implicated  in   up-­‐regulating   levels   of   cyclic   AMP   (cAMP).   This   regulation   is   well   reflected   by   the   reduced   cAMP   levels   in   Nf1-­‐deficient   central   nervous   system   neurons   and   Nf1-­‐ deficient  astrocytes  (Tong  et  al,  2002;  Dasgupta  et  al,  2003;  Warrington  et  al  2007).   NF1  mediation  of  the  cAMP/PKA/Rho  pathway  is  crucial  for  development  of  normal   neuronal   morphology   (Brown   et   al   2012),   and   heterozygous   loss   of   function   mutation  in  Nf1  lead  to  cell  autonomous  reductions  in  neurite  lengths,  growth  cone   areas,   and   cell   survival   (Brown   et   al   2010a).   Though   the   mechanism   is   not   yet   defined,   neurofibromin   functions   in   the   cAMP   pathway   at   the   level   of   adenylyl   cyclase  (AC)  activation  (Dasgupta  et  al,  2003).  Two  mechanisms  utilizing  receptor   tyrosine   kinase   pathway   and   the   heterotrimeric   G-­‐protein   pathways   respectively   are  likely  to  contribute  to  NF1  mediated  AC  signalling.  The  former  requires  the  GAP   activity   of   NF1,   and   the   latter   is   a   Gs   alpha   subunit   (Gαs)-­‐dependent   process   and  

requires  the  C-­‐terminal  region  of  neurofibromin  (Hannan  et  al  2006).  In  addition  to   controlling  neuronal  cell  function  by  regulating  morphology,  cAMP  can  also  exert  a   more  direct  effect  on  the  electrical  properties  of  neurons  by  modulating  the  gating   kinetics  and  surface  expression  of  voltage-­‐gated  ion  channels  (Biel  et  al,  2009).      

Lastly,   an   important   NF1   interacting   partner   has   been   discovered   by   the   study   of   inclusion  body  myopathy  with  Paget's  disease  of  bone  and  frontotemporal  dementia   (IBMPFD).   Like   Nf1,   IBMPFD   is   autosomal   dominant   and   though   fully   penetrant,   patient   populations   show   high   degree   of   phenotype   heterogeneity.The   disorder   is   caused   by   mutations   in   the   valosin-­‐containing-­‐protein   (VCP),   a   direct   protein   interaction   partner   of   neurofibromin.       Both   NF1   and   VCP/p97   are   critical   for   dendritic  spine  formation,  and  disruption  of  the  interaction  between  neurofibromin   and  VCP  impairs  dendritic  spinogenesis  (Wang  et  al  2011).    

 

All  of  the  examples  given  above  highlight  the  broad  spectrum  of  NF1  signalling  in   the   regulation   of   neuronal   function.   In   this   way,   neurofibromin   is   likely   to   affect   brain   function   by   modulating   control   of   dendritic   spinogenesis,   neuron   morphogenesis,   synatpic   plasticity,   and   possibly   by   modulating   ion   channel   function.    

 

Animal  Models  of  NF1:  Insights  on  Etiology  

   

Animal  models  have  given  invaluable  insight  on  the  genetic,  molecular,  and  cellular   mechanisms  of  NF1  pathogenesis.  The  power  of  these  models  comes  from  cleverly   combining   genomic   and   pharmacological   tools   with   appropriate   behavioral   and   physiological   assessments.     NF1   mouse   models   and   patient   populations   display   striking   genenetic   and   phenotypic   similarities.   The   NF1   gene   sequence,   transcriptional   regulation,   and   downstream   targets   are   conserved   across   species   (Bernards   et   al,   1993,   Hajra   et   al,   1994),   and   heterozygous   mutations   of   NF1   function   impair   similar   cognitive   functions   in   mice   (Li   et   al,   2005)   and   humans   (Hyman   et   al   2005).   Yet   the   results   from   such   studies   should   always   be   taken   in   context.  The  evolutionary  divide  between  mice  and  humans,  as  well  as  contributions  

(13)

of  modifier-­‐genes  and  NF1  mutation  type,  must  be  considered  when  assessing  the   translational  validity  of  the  findings.    

 

Behavioral  Tasks  Assigned  to  Specific  Brain  Areas  Highlight  the  Parallels  

of  NF1  mice  and  human  phenotype  

 

The   human   NF1   cognitive   profiles   suggest   dysfunction   in   hippocampus,   parietal/prefrontal   cortex,   and   cerebellum.   In   parallel,   animal   models   have   evaluated  cognitive  functions  using  behavioral  tasks  thought  to  rely  on  these  areas.     The  most  commonly  employed  mouse  model  is  a  heterozygous  null  mutant  mouse,  

Nf1+   /   -­,   developed   by   deletion   of   exon   30   (Jacks   et   al   1994).   The   resulting  

neurofibromin  proteins  are  unstable  and  are  rapidly  targeted  for  proteolysis  (Jacks   et  al  1994),  which  result  in  non-­‐functional  truncated  versions  of  the  protein  (Shen  et   al  1996).    Thereby  this  mouse  model  mimics  70%  of  heterozygous  mutations  seen  in   NF1   patient   population.   The   Nf1+   /   -­‐   mouse   displays   deficits   in   hippocampal-­‐

dependent   spatial   learning   and   memory   (Silva   et   al   1997).   The   hidden   platform   version   of   the   Morris   Water   Maze   Task   employed   in   these   studies   evaluates   a   mouse's  ability  to  learn  and  remember  the  location  of  a  submerged  platform.  A  pilot   study  on  children  with  NF1  revealed  a  similar  pattern  of  spatial  learning  deficits  in  a   virtual  Morris  maze  (Ullrich  et  al,  2010).  Although  the  performance  of  Nf1+  /  -­‐  mice  is  

significantly   lower   than   wild-­‐type   (WT)   mice   on   the   early   trials,   this   deficit   is   overcome  by  additional  training  (Costa  et  al.  2001,  2002;  Cui  et  al.  2008;  Silva  et  al.   1997).  Similarly,  the  beneficial  effects  of  extra  training  have  also  been  described  in   NF1-­‐associated   learning   disabilities,   in   the   form   of   remedial   teaching   (Fuchs   and   Vaughn,   2012).   The   Nf1+   /   -­‐   mouse   model   further   parallels   human   NF1   in   its  

sensitivity   to   modifier   genes,   so   that   the   severity   of   symptom   expression   was   modulated   by   the   genetic   background   of   the   mice   (Costa   2002).   This   was   most   clearly   shown   by   crossing   mice   with   a   heterozygous   knock   out   of   glutamate   receptor   (NMDAR+/-­‐)   with   Nf1+   /   -­‐   mouse.   Whereas   NMDAR+/-­‐   showed   no  

performance  deficit,  poor  task  performance  was  exacerbated  in  mice  carrying  both   mutations   (Silva   et   al   1997).   Another   task   employed   to   investigate   hippocampal   dependent  spatial  learning  is  contextual  fear  conditioning  (Cui  et  al,  2008;  Brown  et   al   2010b).     The   task   measures   a   form   of   spatial   learning   in   which   an   aversive   stimulus  (an  electrical  shock)  is  associated  with  a  particular  neutral  context  (spatial   location),   resulting   in   the   expression   of   fear   responses   to   the   originally   neutral   context.  Performance  on  this  task  was  found  to  be  impaired  in  multiple  NF1  mouse   models  (Cui  et  al,  2008;  Brown  et  al  2010b).    

 

Deficits  in  the  attention  domain  have  also  been  described  in  Nf1+   /   -­‐  mice  (Li  et  al,  

2005;   Brown   et   al   2010b).   These   are   evaluated   using   a   lateralized   reaction   time   task,  which  requires  sustained  attention  on  two  spatially  separated  location  where   visual  stimuli  may  be  presented  (Jentsch,  2003),  and  an  analogue  version  of  the  task   has   been   used   in   humans   to   reveal   attention   deficits   in   NF1   patients   (Robbins   2002).  

(14)

Given  the  vast  evolutionary  differences  between  humans  and  drosophila,  there  are   limitations  on  the  extent  that  human  cognition  can  be  mimicked  in  flies.  Nf1  animal   models  employing  Drosophila  therefore  investigate  the  effects  of  NF1  on  cognition   by  utilizing  olfactory  associative  learning  paradigms  (Guo  et  al,  2000).  In  this  task   flies  are  trained  by  exposure  to  electroshock  paired  with  one  odor  and  subsequent   exposure  to  a  second  odor  without  electroshock.  Learning  and  memory  is  evaluated   by  allowing  flies  to  choose  between  the  two  odors  used  during  training.    

 

Finally,  Nf1+  /  -­  mice  have  been  used  to  investigate  the  motor  deficits  that  are  present  

in  the  NF1  patient  populations  (van  der  Vaart  2011).  Neurofibromin  is  enriched  in   cerebellar  purkinje  cells  (Gutman  et  al,  1995)  and  the  cerebellum  has  a  modulatory   role   in   motor   function   (Mier,   2002).     As   such,   this   mouse   study   on   motor   deficits   utilized  the  Erasmus  Ladder  task  known  to  depend  on  the  cerebellum.  In  this  task   mice  must  move  along  a  horizontal  ladder,  and  learn  to  associate  motor  movements   with   the   presentation   of   a   tone   which   comes   just   prior   to   raising   or   lowering   a   ladder   rung   in   the   path   of   the   moving   animal.   Though   Nf1   model   mice   show   non-­‐ specific   motor   learning   deficits   in   a   rotarod   task,   in   which   mice   are   repeadetly   tested  for  endurance  in  running  on  a  rotating  cylinder  (van  der  Vaart  2011,  Costa  et   al   2001),   Nf1+   /   -­  showed   no   deficits   on   the   Erasmus   ladder   (van   der   Vaart   2011).  

These   results,   together   with   the   finding   that   NF1   children   do   not   display   deficits   across  all  cerebellum-­‐associated  tasks  argue  against  a  causal  role  of  the  cerebellum   in   NF1-­‐associated   motor   learning   problems   (Krab   et   al   2011).   For   now,   the   role   which   neurofibromin   plays   in   the   cerebellum   plays   in   NF1   phenotype   expression   remains   unresolved.     However,   a   recent   review   looking   over   fMRI   studies   of   the   cerebellum  reports  on  a  host  of  non-­‐motor  functions  that  are  mediated  by  this  brain   area   including   attention,   executive   control,   working   memory,   and   learning   (Strick   2009).   Future   studies   could   asses   the   effects   of   cerebellar   NF1   on   these   cognitive   functions  through  a  cre-­‐recombinase  mediated  restricted  expression  of  mutant  NF1   in  purkinje  cells.    

 

Increased   Inhibition,   Decreased   cAMP,   and   Altered   Dopamine   System  

Underlie  NF1  Cognitive  Phenotype      

 

Several   tools   have   been   applied   to   direct   the   cellular   and   molecular   mechanism   underlying   cognitive   deficits   in   Nf1   mice   and   drosophila.   These   studies   employed   four   approaches;   (1)   specific   deletions   of   neurofibromin   exon,   (2)   restrictive   expression   using   cre   recombinase   genetic   tools,   and   (3)   pharmacological   intervention.  

 

In   order   to   investigate   the   role   of   NF1   in   molecular   and   cellular   mechanism   underlying  the  cognitive  deficits,  a  combination  of  all  the  tools  has  been  used,  and   though   a   central   molecular   mechanism   has   emerged,   the   entirety   of   the   NF1   pathogenesis   mechanism   remains   elusive.     Nf1   heterozygous-­‐null   mice   show   impaired   hippocampal   LTP   (Costa   et   al   2002,   Li   et   al   2005).   Interestingly,   heterozygous   Nf1   deletions   driven   by   either   synapsin   I-­‐Cre   or   Dlx5/6-­‐Cre   which  

(15)

yield  expression  in  GABAergic  inhibitory  neurons  result  in  spatial  learning  deficits,   whereas  heterozygous  deletions  of  Nf1  in  excitatory  neurons  do  not  alter  learning   (Li   et   al,   2008).   Consequently,   LTP   deficit   was   found   to   result   from   abberant   NF1   function   in   GABAergic   inhibitory   neurons   (Cui   et   al   2008,   Shyliansky   et   al   2010).   Both   LTP   and   the   associated   deficit   in   Morris   Water   Maze   task   can   be   rescued   by   picrotoxin,   a   GABA   A   receptor   antagonist   (Costa   et   al   2002,   Li   et   al   2005),   which   normalized  the  heightened  inhibitory  transmission  observed  in  Nf1  mice  (Costa  et  al   2002).   Importantly,   mice   lacking   the   alternatively   spliced   NF1   exon   23a,   which   inhibits   the   GTPase-­‐activating   protein   (GAP)   domain   of   Nf1,   displayed   impaired   visuospatial   learning   and   thereby   implicate   enhanced   RAS   signalling   in   Nf1  +/-­  

phenotype.   Furthermore,   the   deficits   of   Nf1+/-­  mice   can   be   rescued   by   genetically  

reducing   the   level   of   H-­‐RAS   or   K-­‐RAS   (Costa   et   al   2002).   The   link   to   inhibitory   neurons   lies   in   neurofibromin's   regulation   of   ERK,   a   downstream   target   of   RAS   (Shilyanski   et   al   2010a),   and   the   finding   that   ERK   dependent   phosphorylation   of   synapsin  1  is  critical  for  GABA  release  (Cui  et  al,  2008).  In  parallel  to  these  findings   in  mice,  fMRI-­‐imaging  studies  in  humans  found  that  the  degree  of  hypoactivation  of   dorsal  lateral  prefrontal  circuits  is  correlated  with  working  memory  performance  in   humans   (Shilyansky   et   al   2010a),   reflecting   the   increased   inhibition   observed   in   multiple   mouse   models.   The   activity   or   RAS   is   dependent   on   its   insertion   in   the   membrane,   and   this   anchoring   requires   post-­‐translational   isoprenylation   (Konstantinopoulos   et   al,   2007).   The   rate-­‐limiting   enzyme   in   the   synthesis   of   isoprenoids  is  3-­‐hydroxy-­‐3-­‐methylglutaryl  coenzyme  A  (MHG-­‐CoA).  The  enzymatic   activity  of  MHG-­‐CoA  can  be  inhibited  by  Statins,  such  as  lovastatin,  a  commonly  used   drug  for  the  treatment  of  hypercholesterolemia.  Interestingly,  lovastatin  can  reverse   the  LTP  deficits  of  Nf1+/-­‐  mice  (Li  et  al  2005).  Furthermore  lovastatin  reverses  the   performance  deficits  of  these  mice  on  a  visuospatial  memory  task  as  well  as  some   components  of  their  attention  deficits  (Li  et  al  2005).  Two  clinical  studies  of  statin   therapy  for  NF1  cognitive  phenotype  are  currently  underway.    

 

Further   interesting   clues   on   NF1   cognitive   phenotype   have   come   from   an   NF1   mouse  model  which  carries  both  a  germline  Nf1  heterozygous  null  mutation  and  a   conditional  homozygous  knockout  restricted  to  glial  fibrillary  acidic  protein  (GFAP)   positive  cells,  or  glia  (Brown  et  al  2010b).  Unlike  heterozygous  null  mutants,  these   mice  develop  optic  glioma  (Bajenaru  et  al  2003,  Zhu  et  al  2005);  a  tumor  present  in   15%  of  NF1  afflicted  children  (Listernick  et  al  1997).  These  Nf1OPG  mice  presented   the   same   visuospatial   cognitive   profile   as   Nf1+/-­   mice   and   show   impairment   on  

Morris  Water  Maze  and  Contextual  Fear  Condition  Task.  Additionaly,  the  mice  were   also   found   to   have   deficits   in   selective   and   non-­‐selective   attention   (Brown   et   al   2010b).  The  attention  deficits  of  the  mice  were  reversed  by  methylphenidate  (MPH)   treatment  (Brown  et  al  2010b).    In  a  follow  up  study  it  was  shown  that  Nf1OPG  mice   have   reduced   dopamine   levels   in   the   striatum,   and   that   these   levels   can   be   normalized   by   MPH   (Brown   et   al   2011b).   Attention   deficit   hyperactivity   disorder   (ADHD)   is   a   frequent   co-­‐morbid   condition   in   NF1   (Hofman   et   al   1994),   and   a   one   year  follow  up  study  demonstrated  the  benefits  of  MPH  treatment  in  this  population   in  both  learning  ability  and  social  behavior  (Mautner  et  al,  2002).  

Referenties

GERELATEERDE DOCUMENTEN

Minimizing cognitive dysfunction and improving cognitive functioning in brain tumour patients may be achieved both by preserving cognitive functioning during antitumor

Also, findings from the individual fixed effects model could mean that both acceptance of homosexuality and level of education are partially confounded by levels of cognitive

The possible argumentative means available to protesters in the activity type of documents of demands can be to appeal to contradictions between actions of the target and

In order to show these apparent inconsistencies, I review studies involving functional magnetic imaging within four cognitive domains, well known to be affected by early life

maltreatment (Belsky, 1984; Beckerman et al., 2017; Coleman & Karraker, 1998; Ellis & Milner, 1981; Jones & Prinz, 2005; Sawrikar & Dadds, 2018); I hypothesized

environmental impacts of coal mining, particularly in regard to local water resources see chapter 4... The necessary provision of water from the Tana River for the MCMP will

H2: MS-patiënten die een zelfbevestigingstaak maken voordat zij informatie lezen over cognitieve klachten als gevolg van het hebben van MS, ervaren minder cognitieve

Wil men een dergelijke onderverdeling toch gebruiken, dan kan men terecht bij Koehne en Ingram (zie hiervoor de literatuurlijst). Vrucht met een groef, steen afgeplat,