• No results found

MRI-BASED NETWORK ANALYSIS ON BRAIN CONNECTIVITY AND ITS DISRUPTIONS IN SCHIZOPHRENIA

N/A
N/A
Protected

Academic year: 2021

Share "MRI-BASED NETWORK ANALYSIS ON BRAIN CONNECTIVITY AND ITS DISRUPTIONS IN SCHIZOPHRENIA"

Copied!
37
0
0

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

Hele tekst

(1)

AND ITS DISRUPTIONS IN SCHIZOPHRENIA

                June,  2012   Sofie  Valk,  0569100   Supervisor:       Boris  Bernhardt,  PhD  

Max  Planck  Institute  of  Human  Cognitive  and  Brain  Sciences   Co-­‐assessor:    

Mike  Cohen,  PhD,  University  of  Amsterdam   Word  count:  

15.266   No.  of  References:  

199  

Master  Brain  and  Cognitive  Sciences,  University  of  Amsterdam     Track:  Cognitive  Sciences  

University  of  Amsterdam  

(2)
(3)

 

1.0  INTRODUCTION  ...  4  

2.0  DEFINING  AND  CONSTRUCTING  BRAIN  NETWORKS  ...  5  

2.1  WHAT  ARE  NETWORKS?  ...  5  

2.2  ANIMAL  CONNECTIVITY  MAPPING.  ...  5  

2.3  HUMAN  IN  VIVO  CONNECTIVITY  MAPPING  THROUGH  MRI  ...  7  

Structural  networks  ...  7  

Functional  networks  ...  9  

3.0  NETWORK  ANALYSIS  ...  10  

3.1  RELEVANT  PARAMETERS  ...  11  

3.2  FINDINGS  IN  HEALTHY  SUBJECTS  ...  13  

Graph  theoretical  characteristics  of  structural  brain  networks  ...  13  

Graph  theoretical  analysis  of  functional  resting  state  MRI  data  ...  14  

Interpretation  of  findings  ...  15  

3.3  HOW  GRAPH  THEORY  SYNTHESIZES  DATA  ...  16  

4.0  NETWORK  DISRUPTIONS  IN  SCHIZOPHRENIA  ...  17  

4.1  HISTORY  OF  SCHIZOPHRENIA;  A  CLASSIC  EXAMPLE  OF  DYSCONNECTIVITY  DISORDER  ...  17  

4.2  NETWORK  STUDIES  ON  SCHIZOPHRENIA  PATIENTS  ...  19  

Low-­‐level  connectional  differences  ...  19  

Topological  connectional  differences  ...  22  

5.0  CHALLENGES  OF  CURRENT  NETWORK  APPROACHES  ...  24  

6.0  SUMMARY  AND  CONCLUSION  ...  26  

7.0  LITERATURE  ...  27  

   

(4)

1.0  Introduction    

One  of  the  reasons  why  modern  neuroscience  attracts  such  a  wide  interest  stems  from  the  inherent  and   challenging  complexity  of  the  human  brain.  Currently,  it  is  believed  that  when  the  structural  architecture   and  the  functional  dynamics  of  brain  networks  are  better  understood,  open  questions  related  to  both   normal   and   abnormal   mental   processes   can   be   answered.   With   the   use   of   graph   theory,   the   mathematical  branch  that  is  used  to  describe  the  topology  of  complex  networks,  different  organizational   characteristics   of   the   brain   network   can   be   formalized   and   investigated   to   probe   novel   hypotheses   about  human  brain  structure  and  function.    

Already   in   the   19th   century,   several   neurological   disorders   such   as   schizophrenia   and   aphasia   were   referred  to  as  disconnection  syndromes,  indicating  that  an  impoverished  connectivity  between  different   brain  regions  leads  to  dysfunction  (Dejerine,  1891;  Liepmann,  1977;  Wernicke,  1874).  However,  these   theories   were   largely   forgotten   during   the   first   half   of   the   20th   century.   In   the   1960s,   Norman   Geschwindt  reintroduced  disconnection  as  a  theme  in  neuroscience  (Geschwindt,  1965).    This  time,  his   reformulations  triggered  explorations  of  cortico-­‐cortical  connectivity  in  both  animals  and  later,  with  the   advent  of  in  vivo  imaging,  in  humans  (Catani  &  Ffytche,  2005)    

The  conceptual  framework  of  connectivity  and  its  disruptions  were  used  to  explain  psychiatric  diseases,   in  particular  schizophrenia.  The  idea  that  schizophrenia  is  linked  to  disintegration  of  the  psyche  is  as  old   as  its  name  (Bleuler,  1913).  On  the  other  hand,  approaches  attempting  to  localize  psychopathology  to   individual   cortical   areas   have   so   far   not   provided   a   sufficient   explanation   for   cardinal   symptoms   of   schizophrenia  such  as  hallucinations  and  delusions  (Tandon,  Nasrallah,  &  Keshavan,  2009).  Indeed,  it  has   been   suggested   that   these   symptoms   relate   to   the   abnormal   connections   between   and   within   brain   regions  (Bullmore  et  al.,  1998).  

This   review   addresses   the   concepts   and   techniques   that   are   currently   applied   to   unravel   general   principles   of   functional   and   structural   connectivity   in   the   human   brain,   and   their   disruptions   in   schizophrenia.   The   first   part   focuses   on   general   methodological   principles   of   network   analysis   in   neuroscience.  We  will  explain  how  a  brain  network  can  be  formalized  and  how  network  properties  are   quantified   across   several   levels   of   analysis.   Contrasting   network   properties   that   have   been   found   in   healthy   human   brains   to   those   in   patients   with   schizophrenia,   we   will   outline   clinical   applications   of   graph   theoretical   network   analysis   to   this   classical   dysconnectivity   disorder.   We   will   conclude   by   summarizing  current  and  future  challenges  of  network  analysis  in  neuroscience.    

(5)

2.0  Defining  and  constructing  brain  networks      

2.1  What  are  networks?  

The   brain   is   a   paramount   example   of   a   complex   network.   Complex   networks   are   networks   that   are   neither   completely   organized   nor   completely   unpredictable.   In   the   brain,   these   characteristics   of   connectivity   found   between   elements   exist   at   multiple   scales.   The   microscopic   scale   is   based   on   individual   neurons   and   synapses.   A   more   macroscopic   scale   focuses   on   anatomically   distinct   regions,   such   as   the   white   matter   tracts   and   interregional   projections,   and   in   between,   the   mesoscale,   which   consists  of  neuronal  populations  and  assemblies  (Sporns,  2011).  

While   animal   research   has   provided   a   wealth   of   information   about   the   wiring   of   individual   neurons,   recent  in  vivo  studies  have  rather  focused  on  brain  regions  and  large-­‐scale  network  (Sporns,  Tononi,  &   Kötter,   2005).   Alongside   steep   advancements   in   acquisition   and   processing   technology   of   cellular   neuroscience  and  human  neuroimaging,  there  was  a  rise  of  modern  network  theory  in  the  late  1990s.   The  conjoint  advancements  in  these  domains  stimulated  the  systematic  investigation  and  description  of   the   brain   as   a   network.     Accordingly,   brain   networks   are   formalized   as   a   complex   graph,   which   is   a   collection   of   nodes   (i.e.,   regions   or   neurons)   and   edges   (i.e.,   connections).     Such   graphs   are   formally   equivalent  to  connectivity  matrices,  in  which  the  network’s  nodes  are  represented  as  row  and  column   indices,  with  the  scalar  value  in  an  individual  cell  quantifying  the  edge  strength  between  a  given  pair  of   nodes.     The   connectivity   matrix   is   analyzed   using   mathematical   methods   derived   from   graph   theory   (Bassett  &  Bullmore,  2009;    Bullmore  &  Sporns,  2009;  Sporns,  Tononi,  &  Kötter,  2005).    

2.2  Animal  connectivity  mapping.  

A  landmark  study  in  both  applied  neuroscience  and  theoretical  network  science  was  the  work  of  Duncan   Watts   and   Steven   Strogatz   (1998).   They   classified   the   topology   of   different   complex   networks   on   the   basis  of  two  global  parameters;  average  clustering  coefficient  and  path  length.  In  a  graph,  the  clustering   coefficient  defines  the  strength  of  connections  between  nodes  in  a  typical  neighborhood  are  connected   among  each  other.    It  thus  represents  a  measure  of  local  network  efficiency  and  robustness  (Latora  &   Marchiori,  2001).  Path  length,  on  the  other  hand,  relates  to  the  global  efficiency  of  the  network,  and   quantifies   the   average   separation   between   any   two   nodes   (Bullmore   &   Sporns,   2012;   Guye,   Bettus,   Bartolomei,   &   Cozzone,   2010).   Given   these   two   parameters,   Watts   and   Strogatz   defined   random   networks  as  those  with  random  connections  between  nodes  and  a  low  average  path  length,  and  regular   networks   as   those   with   only   connections   to   nearest   neighbors.   These   networks   are   highly   clustered.  

(6)

They   also   categorized   a   third,   intermediary   class:   small-­‐world   networks.   Indeed,   these   networks   have   both   a   high   clustering   coefficient   and   a   low   average   path   length.   The   low   average   path   length   is   guaranteed  by  the  presence  of  short  cut  edges  that  connect  clusters  of  nodes  that  would  otherwise  be   much  farther  apart.  Both  characteristics  of  small-­‐world  networks  signify  a  high  local  and  global  efficiency   of  information  transfer.  These  short  cuts  contribute  to  a  higher  interconnectivity  of  different  strongly   inter-­‐connected  sub-­‐networks,  the  so-­‐called  small  worlds.  Small-­‐world  topology  is  generally  associated   with  global  and  local  parallel  information  processing,  sparse  connectivity  between  nodes,  and  low  wiring   costs  (Bassett  &  Bullmore,  2006).  To  illustrate  their  ideas  about  the  topology  of  small-­‐world  networks,   Watts  and  Strogatz  chose  to  analyze  the  neuronal  network  of  the  nematode  worm,  C.  Elegans.  At  the   time  of  their  study,  this  was  the  sole  example  of  a  completely  mapped  neuronal  network  in  an  organism.    

C.  Elegans  has  a  relatively  simple  network  topology.    It  consists  of  959  cells,  302  of  which  are  neurons  

that  are  connected  by  6393  synapses  and  890  electrical  junctions.  They  represented  neurons  as  nodes   and   neural   connections   as   edges.   For   their   analysis,   Watts   and   Strogatz   considered   a   sub-­‐graph   consisting  of  282  neurons.    They  observed  that  a  single  neuron  was,  on  average,  connected  to  14  other   neurons.   Importantly,   comparing   the   neural   network   of  C.   Elegans   to   an   equivalent   random   network,   they   observed   a   five-­‐fold   increase   in   clustering   coefficient   but   virtually   no   increase   in   overall   path   length.  Their  findings  thus  indicated,  for  the  first  time,  that  the  brain  network  of  an  organism  adheres  to   a  small-­‐world  topology.  

Anatomical   connection   matrices   of   larger   animals,   such   as   the   cat   and   the   macaque   monkey,   were   constructed  with  the  use  of   tract-­‐tracing  methods  (Hilgetag,  O’Neill,  &  Young,  2000;  Sporns  &  Kötter,   2004).  This  invasive  method  is  used  to  trace  axonal  projections  and  is  considered  the  gold  standard  for   defining  neuronal  network  anatomy.  However,  these  methods  generally  only  allow  the  tracing  of  fairly   large   cell   populations   and   single   axonal   pathways.   Connections   can   be   visualized   in   an   anterograde   fashion,   by   injecting   tracer   in   the   soma,   or   cell   body,   and   visualizing   their   transport   along   the   axonal   projections  to  the  synapses.  To  trace  projections  from  a  specific  regional  cell,  a  genetic  construct,  virus   or  protein  can  be  locally  injected,  after  which  it  is  allowed  to  be  transported  anterogradely  (Kuypers  &   Ugolini,  1990).  Other  tracers  consist  of  protein  products  that  can  be  taken  up  by  the  cell  and  transported   across  the  synapse  into  the  next  cell.  An  example  is  the  wheat-­‐germ  agglutinin  and  Phaseolus  vulgaris   leucoagglutinin   (Smith,   Hazrati,   &   Parent,   1990).   Conversely,   retrograde   tracing   maps   neural   connections   from   their   synapses   to   the   soma.   Retrograde   tracing   includes   the   use   of   viral   strains   as   markers  of  a  cell’s  connectivity  to  the  injection  site.  Another  technique  is  injecting  “beads”  into  the  brain   nuclei  of  anesthetized  animals.  After  a  few  days  the  animals  are  euthanized  and  the  cells  in  the  origin  of  

(7)

injection  are  visualized  through  a  fluorescence  microscope  (Luo  &  Aston-­‐Jones,  2009;  O’Donnell,  Lavin,   Enquist,  Grace,  &  Card,  1997).  Several  seminal  studies  have  mapped  the  connections  of  macaque  visual   cortex  (Felleman  &  Van  Essen,  1991)  and  the  feline  thalamo-­‐cortical  system  (Scannell,  Burns,  Hilgetag,   O’Neil,   &   Young,   1999).   In   macaque   monkeys,   tract-­‐tracing   studies   have   provided   a   basic   map   of   anatomical   links   between   major   cortical   areas   (Stephan,   2001)   and   provided   valuable   data   on   rostro-­‐ caudal   and   dorsal-­‐ventral   connectivity   gradients   between   major   prefrontal   and   parietal   cortices   (Petrides   &   Pandya,   2009).   Furthermore,   a   broad   range   of   characteristics   related   to   the   topology   of   complex   networks   have   been   identified   in   both   the   macaque   and   cat,   including   the   existence   of   clustering  (Hilgetag  et  al.,  2000),  segregation  (Young,  1992),  together  with  small-­‐world  organization  as   well  as  other  important  topological  attributes  (Sporns,  Tononi,  &  Edelman,  2000;  Sporns  &  Zwi,  2004;   Young,  1992;  Sporns  &  Kötter,  2004;  Sporns  et  al.,  2000;  Sporns  &  Kötter,  2004;  Kötter  &  Stephan,  2003;     Hilgetag,  O’Neill,  &  Young,  1996;  Hilgetag  et  al.,  2000)  

2.3  Human  in  vivo  connectivity  mapping  through  MRI  

Tracer-­‐based   connectivity   mapping   cannot   be   applied   to   the   living   human   brain   due   to   the   invasive   nature   of   this   method.   In   humans,   noninvasive,   in   vivo,   network   mapping   can   nevertheless   be   performed  in  both  structural  and  functional  domains.  Structural  networks  can  be  inferred  on  the  basis  of   diffusion-­‐weighted   MRI   tractography   or   through   analysis   of   inter-­‐regional   covariance   patterns   of   structural   measures   across   subjects.   Functional   networks   can   be   defined   on   the   basis   of   electrophysiological   and   metabolic   signal   correlations,   with   recent   work   being   mostly   based   on   time-­‐ series  correlations  of  task-­‐free,  resting-­‐state,  functional  MRI  signals.  In  the  following  paragraphs,  these   MRI-­‐based   methods   are   explained   in   more   detail,   as   well   as   compared   with   each   other,   in   order   to   illustrate  the  benefits  and  problems  associated  with  each  of  them.    

Structural  networks  

Diffusion-­‐weighted   MRI   relies   on   special   pulse   sequences,   which   elicit   signals   that   scale   with   the   orientation  and  magnitude  of  diffusion  processes  in  soft  tissue,  specifically  water  molecules  (Le  Bihan,   2003;  Basser,  Mattiello,  &  Le  Bihan,  1994;  Basser  &  Pierpaoli,  1996).  In  the  human  brain,  the  Brownian   motion  of  water  molecules  is  strongly  constrained  by  myelinated  fiber  tracts,  which  hinder  any  motion   across  them.  The  signals  recorded  therefore  indicate  two  main  features:  the  deviation  from  randomness   of   diffusion   within   each   measured   voxel,   which   is   expressed   by   the   fractional   anisotropy,   indicating   white   matter   integrity,   and   the   three-­‐dimensional   orientation   of   the   diffusion   tensor   of   the   corresponding   probability   functions.   High   anisotropy   can   also   be   observed   in   un-­‐myelinated   nerves,  

(8)

indicating   that   anisotropy   is   mainly   defined   by   the   axon,   and   myelination   is   not   required   to   detect   anisotropy   (Beaulieu   &   Allen,   1994).   By   analyzing   paths   of   diffusion   direction   throughout   adjacent   voxels,   one   can   estimate   the   course   of   different   fiber   tracts   by   means   of   tractography   algorithms   (Behrens  et  al.,  2003;  Mori,  Crain,  Chacko,  &  van  Zijl,  1999).  In  various  psychiatric  populations  such  as   autism,   (Catani   et   al.,   2008;   Cheng   et   al.,   2010;   Cheung   et   al.,   2009;   Thakkar   et   al.,   2008)   and   schizophrenia  (van  den  Heuvel  et  al.,  2010;  Fornito,  Zalesky,  Pantelis,  &  Bullmore,  2012;  Zalesky,  Fornito,   et   al.,   2010)   diffusion   MRI   tractography   has   been   assessed   to   localize   and   quantify   white   matter   disturbances.   Despite   generating   visually   impressive   results   and   closely   approximating   the   course   of   fiber  tracts  in  deep  white  matter  regions,  such  as  the  corpus  callosum,  diffusion  tractography  remains  an   indirect  measure  of  structural  connectivity.  Indeed,  the  reconstruction  of  the  fibers  relies  mostly  on  the   parameters   used   for   the   reconstruction   algorithm.   There   are   several   additional   limitations,   such   as   finding  the  exact  termination  of  connections,  detecting  collaterals,  tracking  the  very  dense  network  of   horizontal  intra-­‐cortical  connections,  discriminating  between  afferents  and  efferents  and  the  fact  that   the   uncertainty   of   the   location   of   a   tract   increases   when   the   tract   spans   a   larger   distance   (Jbabdi   &   Johansen-­‐Berg,   2011).   Furthermore,   seeding   from   grey   matter   regions   remains   challenging   for   conventional   diffusion-­‐based   tractography   given   the   high   uncertainty   in   fiber   directions   within   and   around  grey  matter  (Jbabdi  &  Johansen-­‐Berg,  2011).  

An  alternative  approach  to  diffusion  MRI  tractography  in  the  structural  domain  infers  networks  based  on   inter-­‐regional  covariance  patterns  in  morphological  measures.    Covariance  mapping  (Bullmore  &  Sporns,   2009;  Lerch  et  al.,  2006),  commonly  based  on  T1-­‐weighted  MRI,  takes  advantage  of  a  high  anatomical   resolution   that   is,   unlike   standard   echo-­‐planar   images   used   for   diffusion   imaging,   only   minimally   affected   by   imaging   artifacts   in   anteroventral   brain   regions.   In   addition,   correlations   of   structural   markers   such   as   cortical   thickness   directly   seed   from   grey   matter   regions,   complementing   diffusion   tractography.   Lastly,   cortical   measurements   have   undergone   surface-­‐based   registration   that   aligns   cortical  folding  patterns  of  each  individual  and  thus  improves  between-­‐subject  correspondence.  While   high   structural   correlations   between   two   regions   do   not   necessarily   signify   a   physical   link,   structural   coupling   has   been   suggested   to   reflect   common   genetic   and   developmental   influences,   a   shared   vulnerability  to  pathology,  and  the  presence  of  persistent  functional-­‐trophic  interactions  (Bernhardt  et   al.,  2008;  Bernhardt,  Chen,  He,  Evans,  &  Bernasconi,  2011;  Bullmore  &  Sporns,  2009;  Lerch  et  al.,  2006;   Raznahan   et   al.,   2010).   Indeed,   previous   work   has   demonstrated   a   high   correspondence   between   structural   covariance   networks   and   those   obtained   from   resting-­‐state   functional   connectivity   (Seeley,   Crawford,  Zhou,  Miller,  &  Greicius,  2009;  Segall,  Allen,  et  al.,  2012).    

(9)

Functional  networks  

Structural   connectivity   is   neither   a   sufficient   nor   a   complete   description   of   brain   networks   (Friston,   2011).   In   fact,   synaptic   connections   in   the   brain   are   in   a   state   of   constant   flux   showing   flexible   and   context-­‐sensitive   modulations,   together   with   time-­‐   and   activity-­‐dependent   effects   (Friston,   2011;   Saneyoshi,   Fortin,   &   Soderling,   2010).   Functional   connectivity   research   addresses   inter-­‐regional   signal   associations,  which  have  been  quantified  using  a  wide  range  of  measures,  including  simple  correlations   and   coherence,   partial   measures,   as   well   as   higher-­‐order   and   non-­‐linear   measures   of   association   (Friston,  2011).    These  analyses  are  performed  within  the  context  of  task  free,  resting-­‐state  functional   MRI  acquisitions,  experimental  tasks,  or  biophysical  simulations  (Smith,  2012).  The  analysis  of  resting-­‐ state  functional  MRI  generally  focuses  on  low-­‐frequency  (e.g.,  0.01  to  0.1  Hz),  spontaneous  oscillations   in   the   blood-­‐oxygen-­‐level-­‐dependent   (BOLD)   responses   (Biswal,   Zerrin   Yetkin,   Haughton,   &   Hydes,   1995).  These  patterns  of  BOLD  activity  have  been  suggested  to  reflect  intrinsic  functional  connectivity   (Greicius  et  al.,  2009;  Pawela  et  al.,  2008).    One  of  the  most  reproducible  networks  identified  in  resting   state   functional   connectivity  analysis   is   the   default   mode   network,   a   collection   of   regions   such   as  the   medial   prefrontal   cortex   and   posterior   midline   regions,   together   with   lateral   parietal   cortices,   which   show   extensive   deactivation   during   externally   focused   cognitive   tasks   (Greicius,   Krasnow,   Reiss,   &   Menon,  2003;  Raichle  et  al.,  2001).  Other  studies  have  identified  intrinsic  networks  encompassing  brains   involved  in  visual,  motor,  language,  and  auditory  processing  that  are  consistent  across  subjects  (Cordes   et  al.,  2000;  Damoiseaux  et  al.,  2006;  De  Luca,  Beckmann,  De  Stefano,  &  Smith,  2006;  Fox  et  al.,  2005;   Greicius  et  al.,  2003;  Hampson,  Olson,  Leung,  Skudlarski,  &  Gore,  2004;  Hampson,  Peterson,  Skudlarski,   Gatenby,  &  Gore,  2002;  Lowe,  Mock,  &  Sorenson,  1998).  In  a  seminal  study,  Smith  and  colleagues  (2009)   compared  spatial  patterns  of  task  activation  networks  based  on  meta-­‐analyses  of  nearly  30,000  subjects   to  the  major  networks  in  the  resting  brain  of  36  subjects.    Observing  close  correspondences  between   task-­‐free  and  task-­‐related  network  components,  Smith  and  colleagues  concluded  that  the  full  repertoire   of  functional  networks  used  by  the  brain  in  action  is  continuously  and  dynamically  active  even  when  at   rest.      

As   mentioned   earlier,   structural   and   functional   network   data   in   humans   is   commonly   inferred   from   indirect   methods.     Imaging-­‐based   findings   from   animals,   where   gold-­‐standard   tracing   techniques   are   available,  can  thus  be  used  to  cross-­‐validate  different  network  metrics  (Kötter,  2007).  In  the  macaque   monkey,  diffusion  MRI  tractography  has  been  shown  to  produce  networks  that  generally  overlap  with   traditional   anatomical   tract-­‐tracing   findings   (Schmahmann,   Pandya,   Wang,   Dai,   &   D’Arceuil,   2007).   Extending   these   results,   Hagmann   and   colleagues   found   significant   overlap   between   macaque  

(10)

connectivity  data  derived  from  diffusion  data  and  from  tract  tracing  in  the  same  monkeys  (Hagmann  et   al.,  2008).  However,  tract  tracing  studies  in  animals  and  imaging  studies  in  humans  are  made  difficult  by   the   uncertainty   of   cross   species   homologies   between   functionally   defined   brain   regions   and   brain   structure  in  general  (Bressler  &  Menon,  2010;  Orban,  Van  Essen,  &  VanDuffel,  2004).    

3.0  Network  analysis  

Despite  variations  in  defining  an  inter-­‐regional  link,  all  imaging  techniques  mentioned  above  allow  the   generation  of  connectivity  matrices.  These  matrices  are  equivalent  to  brain  graphs;  collections  of  nodes   interconnected  by  edges,  which  can  be  readily  analyzed  by  graph  theory,  the  mathematical  analysis  of   complex,  interconnected  networks.    

In   a   brain   graph,   nodes   represent   distinct,   homogeneous   elements   based   on   a   given   structural,   functional,   or   otherwise   objective   criterion.   A   variety   of   techniques   have   been   used   to   generate   anatomical   parcellations   of   the   brain,   which   can   then   be   used   to   define   nodes   structurally.   In   post  

mortem  data,  quantitative  cyto-­‐architectonic  features  (Schleicher,  Morosan,  Amunts,  &  Zilles,  2009),  as  

well   as   neurotransmitter   profiles   (Zilles   &   Amunts,   2009),   have   been   used   towards   this   end.   Using   human   imaging   analysis,   parcellations   have   been   proposed   that   are   based   on   macroscopic   landmarks   (Tzourio-­‐Mazoyer,  Landeau,  Papathanassiou,  Crivello,  &  Etard,  2002),  gyral  folding  patterns  (Desikan  et   al.,  2006),  and  more  recently,  myelin  density  profiles  (Glasser  &  van  Essen,  2011).  The  choice  of  nodal   parcellation   has   important   consequences   for   the   determination   of   network   connectivity,   and   diverse   tradeoffs  between  exactness  and  descriptive  power  have  to  be  made  (Zalesky,  Fornito,  et  al.,  2010;  Van   Dijk   et   al.,   2010).   In   addition   to   the   aforementioned   structural   criteria,   parcellations   have   also   been   performed   purely   in   the   functional   domain.     Examples   are   task   activation   (i.e.,   localizer-­‐based)   parcellations.   Specifically,   Eguiluz   and   colleagues   (2005)   used   a   finger-­‐tapping   task   to   reconstruct   correlation  matrices.  Moreover,  nodes  derived  from  consistently  activated  regions  in  meta-­‐analyses  can   be  used  to  define  nodes  (Fox  et  al.,  2005).  Decompositions  have  also  been  provided  using  data  driven   approaches,   such   as   independent   component   analysis   and   hierarchical   clustering   (Beckmann,   DeLuca,   Devlin,  &  Smith,  2005;  Damoiseaux  et  al.,  2006;  Salvador  et  al.,  2005).  

In   both   structural   and   functional   domains,   one   can   also   apply   random   parcellations   (Hagmann   et   al.   2007;  2008),  as  well  as  voxel-­‐wise  parcellations,  which  offer  a  tremendous  boost  in  resolution  but  also   significantly   increase   computation   time   (Friston,   2011;   Smith,   2012;   Lohmann   et   al.,   2010;   Tomasi   &  

(11)

Volkow,   2010).   Altogether,   the   choice   of   definition   of   a   node   has   a   strong   influence   on   the   network   properties  found,  and  wrongly  placed  nodes  can  influence  the  connections  found  between  nodes.    

Connections   based   on   structural   imaging   data   are   often   described   as   undirected,   that   is,   equivalently   strong   in   both   directions.   In   the   framework   of   structural   MRI   covariance   analysis,   connectivity   is   assumed  to  be  reflected  by  the  cross-­‐subject  correlation  between  regional  morphometric  parameters,   such  as  grey  matter  volume  or  cortical  thickness  (Bullmore  &  Sporns,  2009;  Lerch  et  al.,  2006;  Bernhardt   et  al.,  2011;  Seeley  et  al.,  2009).  Using  diffusion-­‐weighted  MRI,  connectivity  is  typically  inferred  from  a   tractographic  estimate  of  fiber  tracts  between  regional  pairs.  Connectivity  strength  may  be  quantified   using  the  proportion  of  seed  fibers  terminating  on  a  target  region,  some  index  of  fiber  integrity  averaged   over   the   reconstructed   tract,   or   a   composite   measurement   (Jbabdi   &   Johansen-­‐Berg,   2011;   Le   Bihan,   2003).  In  functional  MRI  studies,  edge  strength  is  commonly  inferred  from  the  correlation  of  the  activity   time  courses  (  Bassett  &  Bullmore,  2009).  The  more  similar  the  time  courses  are  between  any  given  pair   of   nodes,   the   likelier   it   is   that   there   is   a   functional   connection   between   those   nodes.   Functional   connectivity,   thus,   reflects   statistical   associations   of   regions,   and   does   not   signify   any   causal   link.   Effective  connectivity,  on  the  other  hand,  refers  explicitly  to  the  influence  that  one  neural  system  exerts   over  another,  either  at  the  synaptic  or  population  level  (Friston,  2011).  Methods  that  are  used  to  infer   effective   connectivity   in   functional   networks   include   structural   equation   modeling   (Bullmore   et   al.,   2003;  McIntosh  &  Gonzalez-­‐Lima,  1994),  dynamic  causal  modeling  (Friston,  2003)  or  Granger  causality   (Brovelli   et   al.,   2004).   However,   in   practice   most   studies   have   been   based   on   undirected   functional   connectivity   (Bullmore   &   Sporns,   2009;   Smith,   2012).   In   the   future,   it   is   aimed   to   unite   functional   connectivity  and  effective  connectivity  (Friston,  2011).  

3.1  Relevant  parameters    

Once  nodes  and  edges  of  a  brain  network  are  defined,  graph-­‐theoretical  parameters  can  be  calculated   that   characterize   various   topological   aspects   of   the   network.   Despite   methodological   challenges   put   forward  by  the  sheer  complexity  of  human  and  animal  brains,  these  parameters  allow  the  identification   of   general   patterns   that   cogently   summarize   organizational   network   principles.   Analyzing   those   characteristics   in   the   unified   framework   of   graph   theory   may   help   us   to   better   understand   the   relationship   between   different   structural   configurations,   functional   brain   dynamics,   and   behavior.   Ultimately,  this  approach  promises  to  shed  light  on  normal  variations,  developmental  alterations,  and   pathological  disruptions  in  brain  networks.  Network  properties  can  be  calculated  on  a  global  level  that   characterizes   the   organization   and   efficiency   of   the   whole   network   (Bullmore   &   Sporns,   2012;   Guye,  

(12)

Bettus,   Bartolomei,   &   Cozzone,   2010).   However,   a   network   can   also   be   described   on   a   local   and  

intermediate  level.    

On  a  local  level,  one  can  use  centrality-­‐based  metrics  that  quantify  the  integration  of  a  node  into  the   whole   network,   such   as   degree   centrality,   betweenness   centrality,   and   eigenvector   centrality.   Degree   centrality  is  the  total  number  of  the  connections  a  node  has.  The  betweenness  centrality  is  defined  by   the  factor  of  shortest  paths  from  one  point  to  another  passed  through  a  node,  and  thus  quantifies  the   location   of   a   node   on   efficient   pathways   of   information   transfer.   These   two   measures   give   complementary   notions   of   the   relevance   of   a   node   for   the   whole   network.     The   additional   metric   eigenvector   centrality   uses   a   recursive   formalization,   where   nodes   have   high   centrality   if   they   are   connected  to  nodes  that  are  central  themselves  (Lohmann  et  al.,  2010).  Although  local,  these  centrality   based  metrics  may  also  be  used  to  quantify  the  resilience  of  a  network  with  respect  to  targeted  network   attacks,   by   which   one   removes   central   nodes   from   the   network   and   assesses   the   impact   on   global   topological   parameters   (Achard   &   Bullmore,   2007;   Honey   &   Sporns,   2008;   Kaiser,   Martin,   Andra,   &   Young,  2007).    Moreover,  it  has  been  found  that  the  degree  distribution  of  the  brain  follows  a  power   law,  with  many  nodes  having  a  low  degree  centrality  and  a  few  nodes  having  a  high  degree  centrality   (Bullmore  &  Sporns,  2009).  Such  highly  central  nodes  are  called  hubs,  and  it  has  been  suggested  that   such  central  nodes  are  crucial  for  network  dynamics  (Freeman,  1977).    

An   intermediate   level   of   network   analysis   operates   on   the   scale   of   modules,   which   are   aggregates   of   nodes.  The  human  brain  small-­‐world  network  is  typically  characterized  by  a  high  number  of  clusters  of   local  interconnected  nodes,  called  modules  (Newman  &  Girvan,  2004),  and  a  relatively  low  number  of   extra-­‐modular   connections.   The   modularity   of   a   network   is   often   based   on   hierarchical   clustering   (Girvan   &   Newman,   2002).   A   modular   parcellation   allows   further   characterization   of   different   nodes.     For  example,  nodes  that  have  a  high  centrality  within  a  module  are  called  provincial  hubs;  conversely,   nodes   that   have   a   high   centrality   between   modules   are   called   connector   hubs.   Modular   organization   provides  the  brain  with  the  capacity  to  modify  and  adapt  one  module,  without  affecting  the  function  of   others.   Furthermore,   a   modular   structure   allows   functional   segregation   within   modules   as   well   as   functional  integration  between  modules  (Sporns,  2000).   The  close  association  of  areas  within  clusters   makes   efficient   recurrent   processing   possible   (Sporns,   2011).   Finally,   the   modular   structure   might   support  synchronous  processing  (Kaiser  &  Hilgetag,  2004)  and  efficient  information  exchange  (Latora  &   Marchiori,   2001).   Therefore,   modularity   can   be   seen   as   a   highly   suitable   characteristic   for   brain  

(13)

networks  that  must  adapt  very  rapidly  across  different  time  scales  (Bullmore  &  Sporns,  2012;  Guye  et   al.,  2010).

3.2  Findings  in  healthy  subjects    

The   parameters   described   above   have   been   used   to   characterize   structural   as   well   as   functional   networks  in  healthy  individuals.      

Graph  theoretical  characteristics  of  structural  brain  networks      

In   humans,   structural   connectivity   studies   of   diffusion   networks   have   revealed   highly   clustered   large-­‐ scale   cortical   networks.   These   networks   have   strong   connections   between   areas   that   are   spatially   proximal  and  functionally  related,  as  well  as  a  relatively  high  local  efficiency,  but  with  a  similar  global   efficiency  in  comparison  to  random  networks  (Gong  et  al.,  2009;  Iturria-­‐Medina  et  al.,  2008;  Hagmann  et   al.,  2008;    Zalesky,  Fornito,  et  al.,  2010).  Hagmann  and  colleagues  (2008)  identified  structural  modules   interconnected  by  highly  central  hub  regions,  as  well  as  a  structural  core  of  highly  interconnected  brain   regions   in   posterior   medial   frontal   cortex.   Regions   in   the   structural   core   share   high   degree   and   betweenness   centrality,   and   they   contain   connector   hubs   that   connect   to   all   other   main   structural   modules.   This   structural   core   also   contained   brain   regions   corresponding   to   the   posterior   part   of   the   default  mode  network.  These  findings  indicate  that  the  structural  core  may  be  an  important  basis  for   shaping   large-­‐scale   brain   dynamics   (Hagmann   et   al.,   2008).   Furthermore,   hubs   regions   such   as   the   precuneus,   posterior   cingulate   gyrus,   putamen,   insula,   superior   parietal,   and   superior   frontal   cortex   have   been   identified,   indicating   the   importance   of   these   regions   within   the   structural   network   of   the   brain  (Gong  et  al.,  2009;  Iturria-­‐Medina  and  colleagues.,  2008).  

Using   the   complementary   framework   of   correlation   analysis   of   cortical   thickness   measurements,   multiple  studies  (He,  Chen,  &  Evans,  2007;  He,  Chen,  &  Evans,  2008;  Yao  et  al.,  2010)  found  robust  small-­‐ world   properties   with   cohesive   neighborhoods   in   the   cortex   and   they   found   the   brain   network   had   truncated   power-­‐law   distributions.   Furthermore,   a   modularity   analysis   of   the   relationships   between   structural   cortical   networks   identified   modules   similar   to   known   functional   domains,   such   as   sensorimotor,  visual,  auditory/language,  strategic/executive,  and  mnemonic  processing  (Chen,  He,  Rosa-­‐ Neto,  Germann,  &  Evans,  2008).    

Age,  gender  and  intelligence  have  also  been  examined  from  the  perspective  of  topological  patterns  of   structural  brain  networks.  Using  diffusion  tractography,  it  was  found  that  age  was  negatively  correlated   with  overall  connectivity,  and  a  shift  of  regional  efficiency  from  parietal  and  occipital  regions  to  frontal  

(14)

and  temporal  regions  (Gong  et  al.,  2009).  Also,  a  structural  covariance  study  found  that  an  elder  group   (mean   age   =   66.6   years)   had   higher   local   clustering   but   lower   global   efficiency   in   comparison   to   a   younger   group   (mean   age   =   46.7)   (Zhu   et   al.,   2012).   Gender   differences   have   also   been   found.   In   a   diffusion   tensor   imaging   study,   it   was   found   that   females   had   higher   local   efficiency   than   males,   and   small   brains   had   greater   local   efficiencies   in   females   but   not   in   males   (Yan   et   al.,   2011).   Previously   another  diffusion  tensor  study  had  also  found  females  showed  greater  overall  cortical  connectivity  and  a   more   efficient   organization   both   locally   and   globally   (Gong   et   al.,   2009).   The   relation   between   intelligence   and   graph   structure   has   also   been   investigated.   In   a   diffusion   tensor   imaging   study,   a   significantly  higher  global  efficiency  and  a  shorter  characteristic  path  length  were  found  in  the  networks   of   the   high   intelligence   groups   (Li   et   al.,   2009).   The   findings   of   these   studies   illustrate   the   relation   between  the  structural  network  topology  in  the  brain  with  age,  gender  and  intelligence.    

Graph  theoretical  analysis  of  functional  resting  state  MRI  data  

Network-­‐level  findings  in  the  functional  domain  generally  correspond  to  those  in  the  structural  domain.     Salvador  and  colleagues  were  the  first  to  demonstrate  small-­‐world  properties  in  functional  MRI  resting   state  data  (Salvador  et  al.,  2005).  Their  study  calculated  an  undirected  graph  derived  by  thresholding  the   healthy  group  mean  based  on  partial  correlation  measurements  of  90  cortical  and  subcortical  regions.  At   around  the  same  time,  an  independent  study  reported  small  world  properties  in  a  set  of  activated  voxels   in  fMRI  data  as  well  (Eguiluz,  Chialvo,  Cecchi,  Baliki,  &  Apkarian,  2005).  Other  studies  have  explored  the   (modular)  community  structure  of  fMRI  networks  using  hierarchical  clustering  analysis  (Ferrarini  et  al.,   2009;   Meunier,   Achard,   Morcom,   &   Bullmore,   2008)   and   have   shown   that   functionally   related   brain   regions  are  more  densely  interconnected,  with  relatively  few  connections  between  functional  clusters.    

Moreover,   it   has   been   found   that   a   higher   global   efficiency   of   functional   networks   can   be   associated   with  a  higher  IQ  (Langer  et  al.,  2012;  Li  et  al.,  2009;  van  den  Heuvel,  Mandl,  Kahn,  &  Hulshoff  Pol,  2009),   suggesting   potential   benefits   of   shorter   path   lengths   in   the   cerebral   cortex   for   higher   cognitive   functioning.   Age-­‐related   efficiency   characteristics   of   functional   networks,   closely   related   to   clustering   and  path  length,  have  also  been  investigated  (Achard,  Salvador,  Whitcher,  Suckling,  &  Bullmore,  2006).   The  authors  used  resting-­‐state  fMRI  data  from  young  and  elderly  adults  and  found  that  brain  networks   in   the   younger   group   were   of   small-­‐world   and   efficient   layout,   indicated   by   high   local   and   global   efficiency   despite   relative   low   connection   cost.   In   the   elderly   group,   efficiency   was   reduced   disproportionately  to  the  wiring  cost  of  adding  a  connection  (Achard  et  al.,  2006),  albeit  still  showing  a   small-­‐world   layout.   This   study   illustrates   how   normal   processes   of   brain   maturation   are   represented  

(15)

through  quantifiable  changes  in  functional  network  topology.  Gender  related  differences  have  also  been   found  in  resting  state  networks,  with  women  showing  more  symmetric  functional  organization  than  men   (Liu,  Stufflebeam,  Sepulcre,  Hedden,  &  Buckner,  2009),  as  well  as  a  gender-­‐by-­‐hemisphere  interaction   (Tian,  Wang,  Yan,  &  He,  2011).  

Interpretation  of  findings  

There   is   now   strong   evidence   that   the   human   brain   generally   has   small   world   properties,   with   high   clustering  and  global  efficiency  (Achard  &  Bullmore,  2007),  a  modular  community  structure  (Chen  et  al.,   2008;  He  et  al.,  2007;  Meunier  et  al.,  2008;  Meunier,  Lambiotte,  &  Bullmore,  2010),  and  an  increased   proportion  of  central  hubs  relative  to  random  graphs  (Achard  et  al.,  2006;  Eguiluz,  Chialvo,  Cecchi,  Baliki,   &  Apkarian,  2005;  Sporns,  Honey,  &  Kötter,  2007).  Generally  speaking,  brain  networks  are  topologically   complex  and  efficiently  organized  given  likely  evolutionary  constraints  on  wiring  cost.  The  combination   of  high  clustering,  high  efficiency,  and  modular  small-­‐world  architecture  can  deliver  both  specialized  and   distributed  processes  (Sporns,  2011).  Specialized  and  local  processes  could  possibly  benefit  from  a  vast   amount  of  local  connections,  while  the  integration  of  segregated  processes  could  rather  benefit  from  a   high  global  information  transfer  efficiency  that  makes  rapid  communication  between  several  specialized   regions  possible  (Bullmore  &  Sporns,  2012;  Tononi,  Spons,  &  Edelman,  1994;  Tononi  &  Sporns,  2003).   For   example,   it   has   been   found   that   functionally   specialized   regions   show   high   clustering   with   areas   having   the   same   functional   specialization   in   the   same   anatomical   neighborhood   (Bullmore   &   Sporns,   2012).  Even  more  so,  the  brain  is  characterized  by  a  hierarchical  group  of  modules,  with  each  node  often   sharing  functional  specializations  with  other  nodes  in  the  same  module  (Chen  et  al,  2008;  Chen  et  al.,   2011;  He  et  al.,  2009;  Meunier  et  al.,  2009).  Moreover,  topological  measures  can  be  used  to  describe   neurodevelopmental  processes,  which  currently  have  already  been  assessed  in  recent  studies  on  aging   in  healthy  adults  (Achard,  Salvador,  Whitcher,  Suckling,  &  Bullmore,  2006;  Gong  et  al.,  2009;  Yan  et  al.,   2011).  For  example,  it  has  been  found  that  location  of  cortical  hubs  in  the  (±39  week  old)  infant  brain   stands  in  stark  contrast  to  the  situation  in  adults.  In  infants,  cortical  hubs  and  their  associated  cortical   networks  are  largely  limited  to  primary  sensory  and  motor  brain  regions,  supporting  perception-­‐action   tasks   (Fransson,   Aden,   Blennow,   &   Langercrantz,   2011).   In   contrast,   the   cortical   hubs   in   adults   are   located   in   areas   supporting   more   higher   order   cognitive   processing   (Fransson,   Aden,   Blennow,   &   Lagercrantz,  2011).  A  series  of  resting-­‐state  functional  MRI  studies  in  children  of  age  7  and  older  have   suggested  that  large-­‐scale  cortical  networks  develop  from  a  “local  and  segregated”  to  a  “distributed  and   integrated”  organization  (Fair  et  al.,  2009;  Fair  et  al.,  2008;  Fair  et  al.,  2007;  Supekar,  Musen,  &  Menon,   2009).   These   findings   suggest   that   brain   organization   may   have   evolved   following   an   adaptive   and  

(16)

dynamic  balance  of  network  cost  and  network  efficiency  (Bullmore  &  Sporns,  2012;  Latora  &  Marchiori,   2001).    

3.3  How  graph  theory  synthesizes  data    

Graph   theoretical   parameters   allow   a   seamless   integration   of   topological   information   from   both   structural   and   functional   data   across   different   modalities,   and   in   principle,   spatial   scales.   Moreover,   graph  theoretical  analysis  makes  it  possible  to  study  the  characteristics  of  the  brain  beyond  the  strength   of  low-­‐level  connections  and  disconnections  between  nodes  (Bullmore  &  Sporns,  2012).    

Structure-­‐function  relationships  in  the  brain  are  still  not  fully  understood.  However,  recent  studies  have   begun  to  use  network  analysis  to  fill  this  gap  (Hagmann  et  al.,  2008;  Honey  et  al.,  2009;  Skudlarski  et  al.,   2008;  Greicius,  Supekar,  Menon  and  Dougherty,  2009;  Seeley  et  al.,  2009;  van  den  Heuvel  et  al.,    2009).   In   their   seminal   study,   Honey   et   al.   compared   functional   networks   based   on   resting   state   signal   correlations   and   structural   networks   based   on   diffusion   tractography   in   the   same   individuals   and   observed   a   moderately   strong   correlation   (Honey   et   al.,   2009).   However,   they   also   found   functional   connectivity   between   regions   that   were   not   directly   linked.   Some   of   the   variance   in   functional   connectivity  could  be  explained  for  by  indirect  connections  and  interregional  distance,  as  well  as  high   variability  within  and  across  both  scanning  sessions  and  model  runs.  Another  study  compared  resting-­‐ state  functional  MRI  data  with  diffusion  tractography  specifically  in  the  default  mode  network  (Greicius   et  al.,  2009).  They  obtained  ROIs  of  the  default  mode  network  based  on  resting  state  data,  and  used   diffusion   tensor   fiber   tractography   to   estimate   the   connections   between   the   regions.   Although   they   found   high   consistency   between   resting-­‐state-­‐based   and   diffusion-­‐based   connectivity,   they   too   concluded  that  there  was  a  dichotomy  between  functional  and  structural  data,  illustrated  by  an  absent   structural   connection   between   the   medial   prefrontal   and   the   medial   temporal   ROI.   Extending   this   analysis  approach  to  more  networks,  van  den  Heuvel  and  colleagues  (2009)  found  that  eight  out  of  nine   functionally  linked  resting-­‐state  networks  were  interconnected  by  white  matter  tracts.    

Intrinsic  functional  connectivity  networks  have  also  been  found  to  be  related  to  structural  covariance   patterns.  Seeley  and  colleagues  (2009)  discovered  a  direct  link  between  intrinsic  connectivity  and  grey   matter   structure.   Across   healthy   individuals,   nodes   within   each   functional   network   showed   closely   correlated  grey  matter  volumes.  Furthermore,  Kelly  and  colleagues  (2012)  specifically  investigated  the   functional   architecture   of   the   insula.   They   parcellated   the   insula   on   the   basis   of   three   distinct   neuroimaging   modalities   (task-­‐evoked   co-­‐activation,   intrinsic   functional   connectivity,   and   gray   matter  

(17)

structural   covariance)   and   demonstrated   convergence   among   modalities   at   a   finer   resolution.   The   convergence   among   large-­‐scale   networks   defined   in   multiple   modalities   supports   the   hypothesis   of   a   fundamental  brain  architecture  governing  both  structure  and  function.    

4.0  Network  disruptions  in  schizophrenia  

Recent  years  have  seen  graph  theoretical  applications  in  clinical  settings  to  neurological  and  psychiatric   disorders.   Such   approaches   promise   to   provide   a   phenotypical   description   of   these   conditions   that   complement  accounts  that  rather  focus  on  the  localization  of  pathology.  In  the  next  sections,  we  will   outline  such  clinical  applications,  using  the  example  of  schizophrenia.    

4.1  History  of  schizophrenia;  a  classic  example  of  dysconnectivity  disorder

 

Schizophrenia  has  been  characterized  as  a  prototypical  disorder  of  brain  connectivity.  The  notion  that   schizophrenia   is   not   caused   by   focal   brain   abnormalities,   but   results   from   pathological   interaction   between   brain   regions,   is   an   old   and   influential   notion   in   schizophrenia   research   (Bleuler,   1913;   Wernicke,  1906).  Carl  Wernicke  hypothesized  that  the  origin  of  psychosis  was  sejunction,  the  anatomical   disruption   of   fiber   tracks   that   blocks   the   regular   associative   processes   and   shunts   them   into   an   abnormal   direction.   These   factors   of   sejunction   are   theoretically   localizable.   In   1908   Bleuler   came   up   with   the   term   schizophrenia,   which   is   a   conjunction   of   the   Greek   words   schizein   (σχίζειν)   and   phrēn,   phren-­‐  (φρήν,  φρεν-­‐),  which  translates  to  the  “splitting  of  the  mind”,  in  order  to  describe  the  separation   of  function  between  personality,  thinking,  memory,  and  perception.  

There   is   no   agreement   regarding   the   exact   nature   of   the   cognitive/neuropsychological   impairment   in   schizophrenia   (Rund,   1998;   Rund,   2009).   To   capture   the   heterogeneous   nature   of   the   disorder,   schizophrenia  has  been  classified  into  a  number  of  subtypes,  including  simple,  catatonic,  disorganized,   paranoid,   schizoaffective,   undifferentiated,   residual   and   latent   schizophrenia   (Tandon   et   al.,   2009;   American  Psychiatric  Association,  2000;  ICD-­‐10,  1993).  Furthermore,  symptoms  of  schizophrenia  can  be   grouped   as   positive,   negative,   and   cognitive   symptoms   (Broome   et   al.,   2005;   Kurtz,   2005).   Positive   symptoms   include   delusions,   disordered   thoughts   or   speech   and   hallucinations.   Negative   symptoms   encompass  deficits,  such  as  flat  affect  and  emotions,  lack  of  motivation  and  poverty  of  speech   (Sims,   2002;   van   Os   &   Kapur,   2009).   Moreover,   disease   severity   and   the   relative   proportions   of   different   symptomatologies  can  vary  across  patients  and  through  the  course  of  the  illness.    Schizophrenia  has  also   been  associated  with  cognitive  deficits  such  as  impaired  theory  of  mind  (Penn,  Sanna,  &  Roberts,  2008;  

(18)

Sprong,   Schothorst,   &   Vos,   2007;   Vauth,   Rusch,   Wirtz,   &   Corrigan,   2004)   and   anxiety   (Bleuler,   1913;   Kraepelin,   1919;   Rund,   2009).   Difficulties   in   working   and   long-­‐term   memory,   attention,   executive   functioning,  and  speed  of  processing  also  occur  (van  Os  &  Kapur,  2009).  Finally,  schizophrenics  are  likely   to  have  comorbid  conditions  such  as  major  depression  and  substance  abuse  (Buckley,  Miller,  Lehrer,  &   Castle,  2009).  

On   a   macroscopic   level,   imaging   studies   have   shown   that   schizophrenia   is   associated   with   ventricular   enlargement   (Daniel,   Goldberg,   Gibbons,   &   Weinberger,   1991;   Lawrie   &   Abukmeil,   1998;   van   Horn   &   McManus,  1992)  and  decreased  cortical  volume,  mainly  in  lateral  and  medial  temporal  regions  (Lawrie   &  Abukmeil,  1998).  Overall,  grey  matter  atrophy  appears  to  be  more  marked  than  white  matter  changes   (Lawrie  &  Abukmeil,  1998;  Zipursky,  Lambe,  Kapur,  &  Mikulis,  1998).  Furthermore,  Voets  and  colleagues   (2008)   identified   folding   abnormalities,   supporting   the   hypothesis   of   abnormal   cortical   development.   Similarly,   young   adults   at   high   risk   of   developing   schizophrenia   by   virtue   of   their   family   history,   also   show  enlarged  ventricles  (Cannon  et  al.,  1993)  and  smaller  medial  temporal  lobes  (Lawrie  et  al.,  1999).   These   findings   suggest   structural   abnormalities   prior   to   the   onset   of   disease   (Harrison,   1999)   and   support  a  neurodevelopmental  model  of  schizophrenia.  Post-­‐mortem  studies  on  the  cellular  level  have   demonstrated  abnormal  reduction  in  presynaptic  and  dendritic  parameters  (Harrison,  1999).  However,   there   are   many   contradictory   findings   and   difficulties   in   replicating   the   findings,   possibly   because   studied  patient  groups  often  suffer  comorbid  diseases  (Harrison,  1999),  which  makes  it  difficult  to  infer   definitive  conclusions.    

Functional  imaging  studies  have  studied  brain  activation  changes  related  to  the  positive,  negative,  and   cognitive  symptoms.  Comparing  hallucinations,  a  cardinal  positive  symptom,  to  non-­‐hallucinatory  states   revealed  largely  increases  in  brain  activity  in  right  temporal  regions,  such  as  the  right  superior  temporal   cortex   (Hoffman,   Anderson,   Varanko,   Gore,   &   Hampson,   2008;   Lennox,   Park,   Jones,   &   Morris,   1999;   Shergill,  Brammer,  Williams,  Murray,  &  McGuire,  2000)  and  frontal  regions,  including  inferior  frontal  gyri   (Shergill   et   al.,   2004,   2000;   Simons   et   al.,   2010).   Further   studies   suggest   that   disordered   memory   functioning,   mediated   by   mesial   temporal   regions,   may   contribute   to   hallucinations   (Seal,   Aleman,   &   McGuire,  2004).    

Negative   symptoms,   on   the   other   hand,   have   also   been   related   to   a   frontal   lobe   abnormalities   (Andreasen   et   al.,   1992;   Volkow   et   al.,   1987;   Wolkin   et   al.,   1992).   In   studies   where   patients   were   to   participate   in   cognitively   challenging   tasks,   especially   when   performing   working   memory   tasks,   abnormalities   in   frontal   activation   have   indeed   been   found   (Brown   et   al.,   2009;   Callicott   et   al.,   2000;  

Referenties

GERELATEERDE DOCUMENTEN

Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of

Dutch National Bicycle Safety Research Agenda: the way to safer cycling.. 25%, the recreational use of bicycles is low in the

Zij hebben zich tot doel gesteld 'het aanjagen van de modernisering en verduurzaming van het teeltareaal door het maken van een gereedschapskist met nieuwe instrumenten en

Magnesium wordt erg sterk beïnvloed door de Ca/Mg verhoudingen en door de EC in de eerste proef.. De K/Ca

fiscale eenheid in de vennootschapsbelasting.. 34 maatregelen die desgewenst strikter kunnen worden geïmplementeerd door de lidstaten. Op de EBITDA-hoofdregel zijn een aantal

Emotionele Empathie Taak als uit de Emotional Contagion Scale naar voren kwam dat mensen met sociale angst juist meer emotionele empathie lijken te hebben als het om negatieve

Although the logical order for a traditional Information Extraction (IE) system is to complete the extraction process before commencing the disambiguation, we start with an

De toekomst zal moeten uitwijzen of er toch juridische mogelijkheden gevonden worden om op basis van de onrechtmatige daad de gebruiker van een volledig zelfrijdende