Mini Disseration Kabir Nigam

37
1 A Survey of the Neuronal Mechanisms Underlying CNS Injury and the Barriers Preventing Successful Recovery By Kabir Nigam

Transcript of Mini Disseration Kabir Nigam

Page 1: Mini Disseration Kabir Nigam

 1  

 

 

 

A  Survey  of  the  Neuronal  Mechanisms  

Underlying  CNS  Injury  and  the  Barriers  

Preventing  Successful  Recovery  

 

 

By  Kabir  Nigam  

 

 

 

 

 

 

 

 

Page 2: Mini Disseration Kabir Nigam

 2  

TABLE  OF  CONTENTS  

 

 

Abstract…….……………………………………..…………………………..…………………………..…..pg.  3  

 

Introduction……………………………………..…………………………..…………………………..…..pg.  4  

 

Astrocytes……………………………………..…………………………..…………………………..…...…pg.  9  

 

Developmental  Molecules…………………………………………………………………….………..pg.  13  

 

Oligodendrocytes……………………………………..…………………………..……………………….pg.  15  

 

Regeneration……………………………………..…………………………..…………………………..…pg.  19  

 

Conclusion……………………………………..…………………………..…………………………..…….pg.  24  

 

References……………………………………..…………………………..……………………….…..……pg.  27  

 

 

 

 

 

Page 3: Mini Disseration Kabir Nigam

 3  

ABSTRACT  

 

Injuries  to  the  central  nervous  system  can  be  caused  by  either  physical  impact  or  by  

restriction  of  blood  flow  to  its  constituents.  These  injuries  often  have  very  serious  

consequences,  in  which  full  rehabilitation  is  rare.  The  main  reason  for  this  is  that  

upon  injury,  the  affected  area  turns  into  an  environment  where  axonal  regeneration  

is  not  possible.  This  includes  the  formation  of  a  dense  glial  scar,  the  secretion  of  

repulsive  growth  factors,  and  upregulation  of  glycoproteins  that  induce  actin  

depolymerization.  Strategies  to  overcome  such  inhibition  include  studying  the  

intercellular  mechanisms  by  which  these  molecules  act,  and  targeting  either  the  

molecule  itself  or  downstream  secondary  molecules  that  these  inhibitory  

components  affect,  with  the  greater  goal  of  blocking  the  transduction  of  their  signals  

and  creating  an  environment  that  is  permissible  to  neuronal  growth.    

 

 

 

 

 

 

 

 

 

 

Page 4: Mini Disseration Kabir Nigam

 4  

INTRODUCTION  

 

The  central  nervous  system  is  perhaps  the  most  complex  system  known  to  the  

human  race,  with  over  100  billion  nerve  cells  making  hundreds  of  trillions  of  

synaptic  connections  (Koch  &  Laurent,  1999).  It  governs  the  vital  functions  that  

sustain  life,  and  is  thus  a  priority  in  terms  of  biomedical  study.  Given  its  complexity,  

the  CNS  is  still  relatively  a  black  box,  a  massive  uncharted  territory  of  current  

exploration.  Damage  to  the  cells  that  encompass  the  CNS  often  has  devastating  

results  that  severely  impair  the  affected  individual,  if  not  resulting  in  death.  Thus,  it  

is  important  to  study  how  the  CNS  responds  to  such  insult,  as  through  

understanding  the  natural  mechanisms  that  mediate  recovery,  we  can  use  medical  

technology  and  research  to  facilitate  this  process.  

 

There  are  three  main  types  of  CNS  injury:  traumatic  brain  injury,  spinal  cord  injury  

and  ischemia.  The  first  is  defined  as  sudden  physical  damage  to  the  brain,  either  

through  impact  to  the  intact  skull  or  physical  penetration  of  actual  brain  tissue  

(Finnie  &  Blumbergs,  2002).  Spinal  cord  injury  is  physical  damage  to  the  spine,  

again  through  impact  or  penetration  that  damages  the  axons  that  encompass  the  

spinal  cord,  which  carry  information  from  the  brain  to  the  body  and  vice  versa  

(McDonald  &  Sadowsky,  2002).  Ischemia  involves  the  restriction  of  blood  to  

neurons,  often  as  a  result  of  a  stroke,  restricting  them  of  oxygen  and  glucose  that  is  

necessary  for  proper  cellular  function  (Aarts  &  Tymianski,  2005).  

 

Page 5: Mini Disseration Kabir Nigam

 5  

“Approximately  200,000  people  die  each  year  in  the  United  States  from  brain  

injuries,  with  an  additional  500,000  hospitalized  for  treatment.  About  10%  of  

surviving  individuals  have  continuing  disabilities  that  may  impair  their  

ability  to  live  independently…  In  the  US,  approximately  8,000  new  cases  of  

spinal  cord  injury  occur  each  year,  and  an  estimated  450,000  people  in  the  

country  live  with  the  condition.”  (Physicians’  Desk  Reference)  

 

Neuroinflammation  is  a  relatively  recent  term  within  neuroscience  that  defines  the  

CNS’s  response  process  to  injury.  It  has  only  been  within  the  past  decade  that  

scientists  discovered  that  the  brain  exhibits  immune  activity  unique  to  that  of  the  

rest  of  the  body.  This  response  process  is  highly  complex,  and  is  not  fully  

understood,  however  the  study  of  neuroinflammation  has  become  a  very  “hot”  topic  

in  modern  neuroscience,  with  many  labs  dedicating  their  research  to  studying  this  

process.    

 

Page 6: Mini Disseration Kabir Nigam

 6  

 

Figure  adapted  from  Nature  Reviews  Neuroscience  (Popovich  &  Longbrake,  2008)  

 

 

Page 7: Mini Disseration Kabir Nigam

 7  

There  are  a  multitude  of  different  aspects  of  the  neuroinflammatory  process  that  

have  been  identified,  and  research  shows  these  aspects,  though  acutely  beneficial  in  

the  healing  process,  can  also  be  harmful  (Nakajima  &  Kohsaka,  2001).  Such  aspects  

are  largely  mediated  by  the  activation  of  astrocytes  and  microglia,  two  types  of  glial  

cells  that  play  important  roles  in  maintaining  proper  neuronal  functionality.  

Astrocytes  help  provide  nutrients  to  nervous  tissue  and  maintain  extracellular  ion  

balances,  while  microglia  are  the  macrophages  of  the  CNS,  constantly  clearing  

pathogens  and  extracellular  debris  through  phagocytosis  (Nakajima  &  Kohsaka,  

2001;  Kimelberg  &  Nedergaard,  2010).  But  upon  CNS  insult,  these  cells  quickly  

respond  by  producing  inflammatory  mediators.  Microglia  are  activated  by  CNS  

damage  or  infection,  and  are  recruited  rapidly  to  the  area  of  injury.  The  presence  of  

damaged  cells  causes  microglia  to  produce  cytokines  and  chemokines  that  can  

either  damage  or  protect  neighboring  cells.  Cytokines  can  cross  the  blood–brain  

barrier  to  recruit  mediators  like  leukocytes  originating  from  the  periphery  (Lucas  et  

al.,  2009).  Astrocytes  are  most  known  for  their  ability  to  form  scars  that  isolate  

damaged  tissue  from  healthy  tissue,  minimizing  the  spread  of  infection  and  cellular  

damage  while  allowing  for  the  damaged  area  to  heal  through  the  reorganization  of  

blood  vessels  that  provide  factors  necessary  for  proper  healing  (Stichel  &  Müller,  

1998).    

 

However  despite  promoting  the  healing  of  damaged  tissue,  neuroinflammation,  

otherwise  known  as  gliosis,  has  effects  that  inhibit  axonal  regeneration,  severing  

proper  communication  between  the  affected  region  and  the  rest  of  the  brain,  thus  

Page 8: Mini Disseration Kabir Nigam

 8  

resulting  in  a  loss  of  function.  The  purpose  of  this  paper  will  be  to  investigate  what  

factors  inhibit  axonal  regeneration  and  to  discuss  therapeutic  techniques  that  show  

promise  with  regards  to  overcoming  regenerative  failure.    

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Page 9: Mini Disseration Kabir Nigam

 9  

ASTROCYTES  

 

Upon  CNS  insult,  astrocytes  quickly  react  by  forming  what  is  known  as  a  glial  scar.  

This  process  involves  the  proliferation  as  well  as  morphological  and  functional  

changes  within  astrocytes.  Such  cells  are  known  as  reactive  astrocytes  (Fawcett  &  

Asher,  1999;  Sofroniew,  2005).  Proliferation  has  been  marked  by  a  strong  increase  

in  the  synthesis  of  Glial  Fibrillary  Acidic  Protein  and  an  increase  in  extracellular  

matrix  molecules  (Wilhelmsson  et  al.,  2006).  The  proliferation  of  astrocytes  serves  

to  form  a  dense  web  of  interconnected  tissue  that  surrounds  and  fills  the  vacant  

spaces  caused  by  degenerating  and  dead  neuronal  tissue  (Stichel  &  Müller,  1998).  

This  layer  of  cells  functions  to  protect  surrounding  healthy  tissue  from  being  

damaged  by  potential  microbial  agents,  while  maintaining  a  homeostatic  

environment  and  protecting  the  damaged  are  from  other  proinflammatory  

molecules,  growth  factors,  and  free  radicals  (Rolls  et  al.,  2009).  Additionally,  

astrocytes  mediate  the  revascularization  of  the  damaged  tissue  that  promotes  repair  

by  providing  the  affected  area  with  nutritional  and  metabolic  support.  However,  

reactive  astrocytes  also  produce  molecules  that  chemically  inhibit  neuronal  growth,  

preventing  full  recovery  of  the  damaged  area  in  the  long-­‐term  (Huang  et  al.,  2014).  

 

The  most  obvious  explanation  for  this  inhibitory  mechanism  is  that  the  density  of  

the  glial  scar  is  so  strong  that  it  provides  a  mechanical  barrier  that  prevents  

anything  from  getting  into  the  area  it  surrounds  (Windle  &  Chambers,  1950).  

However  what  seems  to  play  a  bigger  role  are  the  inhibitory  molecules  that  are  

Page 10: Mini Disseration Kabir Nigam

 10  

upregulated  upon  reactive  astrogliosis,  as  it  has  been  shown  that  axonal  

regenerative  failure  will  still  occur  in  vitro  when  the  glial  scar  is  removed  from  the  

neuroinflammatory  environment  (Rudge  and  Silver,  1990).  Such  molecules  include  

Semaphorin  3  (Pasterkamp  et  al.,  2001),  ephrin-­‐B2  (Bundesen  et  al.,  2003)  and  

chondroitin  sulfate  proteoglycans  (Jones  et  al.,  2003).  Evidence  that  supports  the  

environment  as  a  factor  in  axonal  regenerative  failure  includes  micro-­‐

transplantation  of  adult  dorsal  root  ganglion  cells  into  either  intact  host  glial  

environment  or  a  damaged  one.  Though  “these  sensory  axons  regenerate  rapidly  

over  long  distances  within  adult  white  matter  tracts  of  the  brain  and  spinal  cord,  as  

the  growing  adult  neurons  reach  an  area  of  CNS  damage,  with  associated  

inflammatory  infiltrates  and  inhibitory  molecules,  the  axons  convert  into  a  

dystrophic  state  and  are  unable  to  continue”    (Fitch  &  Silver,  2008).  

 

Chondroitin  sulfate  proteoglycans  are  strongly  upregulated  following  CNS  injury.  

There  are  various  types  of  CPSG’s  that  are  differentially  expressed  that  have  been  

shown  to  inhibit  neuronal  growth  (Snow  et  al.,  1990;  Jones  et  al.,  2003).  Degradation  

of  CPSG’s  following  injury  is  shown  to  partially  enhance  axonal  growth  in  animals  

(Lee  et  al.,  2010).  Cleavage  of  the  glycosaminoglycan  side  chains  by  Chondroitinase  

ABC  also  makes  the  environment  substantially  more  permissive  to  neuronal  

outgrowth,  suggesting  that  the  inhibitory  effect  on  axonal  regeneration  is  primarily  

dependent  on  the  “sulfation  pattern  of  GAG  chains,  since  preventing  GAG  sulfation  

eliminates  much  of  the  inhibitory  activity  on  axon  growth  in  vitro”  (Sharma  et  al.,  

2012).    

Page 11: Mini Disseration Kabir Nigam

 11  

 

 

Figure  adapted  from  The  University  of  Leipzig  (Schnabelrauch  et  al.)  

 

The  mechanisms  through  which  CPSGs  mediate  their  inhibitory  effects  on  axonal  

regeneration  are  complex  and  not  fully  understood.  The  first  and  most  obvious  

reason  is  that  the  diverse  network  of  GAG  side  chains  produces  a  dense  barrier  

prohibiting  growth-­‐promoting  molecules  from  interacting  with  the  cell  surface  

(Silver  &  Miller,  2004).  Recent  studies  have  shed  light  on  the  intercellular  

mechanisms  through  which  CPSGs  may  act.  The  Epidermal  Growth  Factor  Receptor  

is  one  such  target,  and  addition  of  CPSGs  to  cerebral  granular  neurons  was  shown  to  

elicit  EGFR  phosphorylation.  Furthermore,  EGFR  inhibitors  were  found  to  neutralize  

the  growth  inhibitory  effects  of  CPSGs  (Koprivica  et  al.,  2005).  Another  molecule  

identified  that  binds  to  CPSGs  was  the  transmembrane  protein  tyrosine  

phosphatase,  PTPsigma,  a  receptor  for  heparin  sulfate  proteoglycans  that  serves  to  

guide  axons  during  development  (Aricescu  et  al.,  2002;  Johnson  &  Van  Vactor,  

2003).  Because  of  the  structural  homology  shared  between  HPSGs  and  CPSGs  

CPSG  

CPSG  

GAG  

Page 12: Mini Disseration Kabir Nigam

 12  

(Kjellén  &  Lindahl,  1991),  scientists  tested  whether  CPSGs  bind  to  PTPsigma  

receptors,  and  found  that  binding  occurs  with  high  affinity.    Dorsal  root  ganglion  

cells  from  PTPsigma  knockout  mice  were  then  probed  with  CPSGs,  with  results  

showing  that  the  knockout  PTPsigma  DRG  cells  exhibited  substantially  less  

inhibition  of  growth  than  the  wild-­‐type  DRG  cells  (Shen  et  al.,  2009).  Last,  it  has  

been  shown  that  when  CPSGs  inhibit  axon  growth,  there  is  activation  of  the  

Rho/ROCK  pathway,  a  pathway  that  will  be  described  later  when  discussing  myelin  

inhibitors.  Use  of  a  ROCK  inhibitor  was  able  to  sufficiently  suppress  the  inhibition  

induced  by  CPSGs  in  outgrowth  assays  (Monnier  et  al.,  2003).    

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Page 13: Mini Disseration Kabir Nigam

 13  

DEVELOPMENTAL  MOLECULES    

 

 

 

 

 

 

 

 

Figure  adapted  from  Riken  Brain  Science  Institute  (Kamiguchi)    

 

Studying  molecules  that  regulate  the  development  of  the  nervous  system  has  been  

useful  in  identifying  other  factors  that  play  a  role  in  inhibiting  neuronal  outgrowth  

after  CNS  injury.  One  such  molecule  is  ephrin-­‐B3,  which  is  shown  to  be  upregulated  

after  spinal  cord  injury  (Miranda  et  al.,  1999).  During  development,  expression  of  

this  molecule  prevents  crossing  of  neurons  in  the  coritospinal  tract  across  the  

midline,  as  these  neurons  possess  the  receptor  for  this  protein,  EphA4  receptor  

tyrosine  kinase,  and  knocking  out  of  this  receptor  results  in  undesired  crossing  

(Yokoyama  et  al.,  2001).  It  has  been  noted  that  ephrin-­‐B3  expression  is  limited  to  

myelinating  oligodendrocytes,  and  that  “postnatal  EphA4-­‐positive  cortical  neurons  

retain  their  sensitivity  to  ephrin-­‐B3  in  myelin  and  that  this  ligand  accounts  for  a  

fraction  of  the  inhibitory  activity  in  CNS  myelin”  (Benson  et  al.,  2005).  Similarly,  

ephrin-­‐B2  seems  to  be  expressed  by  astrocytes,  and  upregulation  of  its  expression  is  

Semaphorins,    ephrins  

Page 14: Mini Disseration Kabir Nigam

 14  

observed  to  occur  in  later  time  points  following  spinal  cord  injury.  Though  this  is  

thought  to  account  for  the  segregation  of  EphB2  receptor  -­‐bearing  fibroblasts  from  

ephrin-­‐B2  expressing  astrocytes  in  the  glial  scar,  it  is  possible  that  ephrin-­‐B2  

expressing  astrocytes  may  repulse  other  EphB2  receptor-­‐containing  axons  from  the  

injured  area  (Bundesen  et  al.,  2003).    

 

Another  factor  in  this  group  of  developmental  molecules  is  semaphorins,  a  family  of  

secreted  and  membrane-­‐bound  proteins  that  serve  as  repulsive  guidance  cues  

during  development  (Bagnard  et  al.,  1998).  In  the  injured  CNS,  semaphorins  are  

expressed  by  both  oligodendrocytes  and  by  meningeal  fibroblasts  that  occupy  a  

portion  of  the  glial  scar  (De  Winter  et  al.,  2002;  Moreau-­‐Fauvarque  et  al.,  2003).  This  

is  supported  by  the  fact  that  that  injured  CNS  neurons  express  semaphorin  receptor  

components,  potentially  making  them  sensitive  to  the  repulsive  effects  of  

semaphorins  (Pasterkamp  &  Verhaagen,  2001).  Sema3s  are  the  family  of  

semaphorins  found  in  the  glial  scar,  and  it  has  been  shown  that  “sprouting  sensory  

axons  are  responsive  to  Sema3-­‐mediated  axon  repulsion  in  vitro  and  in  vivo”  and  

that  “conditioning  peripheral  nerve  injuries  that  allow  dorsal  root  ganglia  axons  to  

regenerate  centrally  do  not  promote  regenerative  axon  growth  through  Sema3-­‐

expressing  scar  tissue”  (De  Winter  et  al.,  2002;  Pasterkamp  &  Verhaagen,  2006).  

Sema4D  and  Sema5A  are  two  other  semaphorins  that  are  expressed  in  

oligodendrocites  (Goldberg  et  al.,  2004;  Yamaguchi  et  al.,  2012).  Both  have  also  been  

shown  to  inhibit  axonal  growth,  with  Sema4D  being  strongly  upregulated  following  

CNS  injury  (Moreau-­‐Fauvarque  et  al.,  2003).  

Page 15: Mini Disseration Kabir Nigam

 15  

OLIGODENDROCITES  

 

Oligodendrocites  are  a  type  of  neuroglia  that  provide  support  and  insulation  to  

neurons  in  the  CNS.  These  cells  join  together  to  form  what  is  known  as  the  myelin  

sheath,  which  functions  to  enhance  conduction  of  the  action  potential  along  the  axon  

by  increasing  the  resistance  and  lowering  the  capacitance  of  the  axonal  membrane  

(Edgar  &  Sibille,  2012).    It  has  become  known  that  myelin  also  has  the  ability  to  

block  neuronal  regeneration  (Berry,  1982).  This  knowledge  was  furthered  through  

experiments  by  Martin  Schwab  using  a  monoclonal  antibody  IN-­‐1,  which  when  

“raised  against  a  fraction  of  myelin  that  did  not  support  neurite  extension,  allowed  

axons  to  grow  on  myelin  both  in  culture  and  in  vivo”  (Caroni  &  Schwab,  1988;  Filbin,  

2003).  

 

It  was  later  discovered  that  the  molecule  which  was  inhibited  by  this  antibody  was  a  

protein  called  Nogo-­‐A  which  associates  with  the  endoplasmic  reticulum  and  on  the  

oligodendrocyte  cell  surface  (Chen  et  al.,  2000;  Huber  et  al.,  2002;  Wang  et  al.,  

2002).  Nogo-­‐A  has  two  inhibitory  domains  where  one  is  extracellular  and  the  other  

is  in  a  cytoplasmic  compartment.  These  domains  are  specifically  a  66  amino  acid  

extracellular  domain  and  the  acidic  amino  terminus  of  Nogo-­‐A  (GrandPré  et  al.,  

2000).  Both  domains  have  been  shown  to  induce  growth  cone  collapse  regardless  of  

their  location.  Upon  CNS  injury,  it  is  possible  that  both  the  Nogo-­‐A  located  in  the  

endoplasmic  reticulum  as  well  as  the  Nogo-­‐A  located  in  the  cytoplasmic  

Page 16: Mini Disseration Kabir Nigam

 16  

compartment  would  become  exposed  to  an  extracellular  environment  that  upon  

interacting  with  axons  attempting  to  regenerate,  would  inhibit  them  (Filbin,  2003).    

 

Two  additional  glycoproteins  have  been  identified  that  inhibit  axonal  regeneration:  

Myelin-­‐associated  glycoprotein  (MAG)  and  oligodendrocyte  myelin  glycoprotein  

(Omgp).  MAG  is  a  member  of  the  immunoglobulin  super  family,  and  appears  to  

exhibit  bifunctionality,  promoting  growth  during  early  stages  of  development  while  

inhibiting  growth  in  older  stages  (Lai  et  al.,  1987;  McKerracher  et  al.,  1994;  Turnley  

&  Bartlett,  1998).  MAG  is  located  on  the  periaxonal  membrane,  allowing  for  easy  

interaction  with  an  axonal  receptor  (Trapp,  1988;  Filbin,  2003).  Omgp  has  also  been  

shown  to  induce  growth  cone  collapse  and  inhibit  neurite  outgrowth  (Wang  et  al.,  

2002).  In  the  spinal  cord,  Omgp  is  expressed  at  the  Nodes  of  Ranvier,  functioning  to  

maintain  the  structure  of  these  areas  by  inhibiting  axon  growth  (Nie  et  al.,  2006).  

This  glycoprotein  is  linked  to  the  cell  membrane  in  myelinating  oligodendrocytes,  

and  is  localized  to  the  glial–axonal  interface  of  myelinated  axons  (Kottis  et  al.,  2002).    

 

Interestingly,  although  bearing  no  structural  or  sequence  homology,  Nogo,  Mag  and  

Omgp  all  act  on  the  same  receptor,  the  glycosylphosphatidylinositol-­‐linked  Nogo  

receptor,  which  is  expressed  on  the  surface  of  various  neurons  (Fournier  et  al.,  

2001;  Domeniconi  et  al.,  2002;  Wang  et  al.,  2002).  Because  this  receptor  has  no  

transmembrane  or  cytoplasmic  domains,  it  requires  co-­‐receptors  to  transduce  the  

inhibitory  signal  (Filbin,  2003).  The  first  co-­‐receptor  identified  was  the  p75  

neurotrophin  receptor  (Yamashita  et  al.,  2002).  Similar  to  the  mechanism  

Page 17: Mini Disseration Kabir Nigam

 17  

underlying  CPSG  inhibition,  “binding  of  myelin  inhibitors  to  the  NgR1-­‐p75NTR  

receptor  complex  activates  protein  kinase  C  (PKC)  and  causes  the  small  GTPase  Rho  

to  assume  its  active,  GTP-­‐bound  state.  Rho-­‐mediated  activation  of  downstream  

effectors  such  as  Rho-­‐associated  kinase  (ROCK)  then  induces  actin  polymerization,  

which  leads  to  growth  cone  collapse  and  inhibition  of  neurite  outgrowth”  (Hannila  &  

Filbin,  2008).  In  order  for  p75  to  activate  Rho,  it  must  be  cleaved,  as  it  is  the  

molecule’s  intracellular  domain  that  is  functional  in  activating  Rho  (Domeniconi  et  

al.,  2005).    

 

Although  inactivation  of  p75  is  shown  to  promote  neuronal  outgrowth,  co-­‐

expression  of  p75  with  NgR1  in  non-­‐neuronal  cells  led  for  an  inability  for  Omgp  to  

activate  Rho,  suggesting  the  presence  of  another  protein  that  mediates  functionality  

in  this  complex  (Yamashita  et  al.,  2002;  Mi  et  al.,  2004).  Surveying  CNS  proteins  for  

their  ability  to  interact  with  NgR1  led  to  the  discovery  of  Lingo-­‐1.  The  existence  of  a  

ternary  complex  was  supported  by  binding  assays  and  experiments  that  showed  

occurrence  of  Rho  activity  only  when  all  three  proteins  were  present,  as  opposed  to  

the  lack  of  Rho  activity  in  binary  complexes  expressing  only  two  of  the  three  

proteins.  When  analyzing  the  structural  morphology  of  Lingo-­‐1,  the  presence  of  an  

EGFR-­‐like  tyrosine  phosphorylation  site  was  discovered.  Deletion  of  this  region  

resulted  in  a  diminished  inhibitory  effect  on  neuron  outgrowth  (Mi  et  al.,  2004).    

 

However,  p75  is  not  expressed  in  all  regions  of  the  brain,  suggesting  the  presence  of  

yet  another  protein  that  is  associated  with  NgR1.  This  led  to  the  discovery  of  Troy,  a  

Page 18: Mini Disseration Kabir Nigam

 18  

tumor  necrosis  factor  receptor  that  is  expressed  on  adult  neurons.  Similar  to  the  

experiments  with  Lingo-­‐1,  binding  assays  and  assessment  of  RhoA  activity  

supported  the  presence  of  another  ternary  complex  consisting  of  NgR1,  Lingo-­‐1  and  

Troy,  with  Troy  functioning  as  a  substitute  for  p75.  Troy-­‐knockout  mice  exhibit  

substantially  larger  neuronal  outgrowths  as  compared  to  wild-­‐type  mice,  further  

supporting  its  role  in  mediating  the  functional  activity  of  NgR1.    These  results  

indicate  Troy  as  a  functional  homolog  to  p75  in  areas  of  the  brain  where  p75  is  not  

expressed  (Park  et  al.,  2005;  Mandemakers  &  Barres,  2005).    

 

   

Figure  adapted  from  Experimental  Neurology  (Akbik  et  al.,  2012)  

 

 

Page 19: Mini Disseration Kabir Nigam

 19  

REGENERATION  

The  goals  of  attempts  to  promote  axonal  regeneration  all  involve  trying  to  make  the  

damaged  area  permissive  to  growth.  One  way  to  do  this  is  to  transplant  cells  from  

regions  known  to  promote  axonal  regeneration  to  the  areas  of  CNS  injury.  Data  

supports  the  regeneration  of  axons  in  the  olfactory  bulb.  “Normal  and  sectioned  

olfactory  axons  spontaneously  grow  within  the  adult  olfactory  bulb,  establishing  

synaptic  contacts  with  their  targets.  The  major  difference  between  a  regenerating  

system  such  as  the  olfactory  bulb  and  the  rest  of  the  CNS  is  the  presence  of  

ensheathing  glia  (EG)  in  the  former”  (Ramón-­‐Cueto  et  al.,  1998).    To  see  if  EGs  have  

the  ability  to  regenerate  axons  in  the  CNS,  pure  EGs  were  transplanted  to  the  dorsal  

root  entry  zone  of  a  transected  region  of  one  dorsal  root  at  the  cord  entry  point.  

“Three  weeks  after  transplantation,  numerous  regenerating  dorsal  root  axons  were  

observed  re-­‐entering  the  spinal  cord…  Neither  ensheathing  cells  nor  regenerating  

axons  invaded  those  laminae  they  did  not  innervate  under  normal  circumstances”  

(Ramón-­‐Cueto  &  Nieto-­‐Sampedro,  1994).  Further  experimentation  showed  

successful  recovery  of  forepaw  function  in  mice  after  the  region  of  the  spinal  cord  

controlling  this  area  was  transected  and  EGs  were  transplanted  to  the  area,  allowing  

for  a  permissive  growth  environment  that  restored  functional  connectivity  (Li  et  al.,  

1997).  

 

An  enzyme  discussed  earlier  called  Chondroitinase  ABC  (ChABC)  has  been  shown  to  

promote  functional  recovery  of  locomotor  behavior  in  mouse  models  of  dorsal  

Page 20: Mini Disseration Kabir Nigam

 20  

column  spinal  cord  injury.  When  administered,  ChABC  functions  to  delete  the  

glycosaminoglycan  side  chains  on  the  chondroitin  sulfate  proteoglycan,  which  is  

thought  to  render  the  protein  ineffective  with  regards  to  transducing  the  inhibitory  

signal.  In  a  rat  model  of  spinal  cord  injury,  electrophysiological  and  behavioral  

testing  showed  functional  recovery  of  synaptic  connections  through  the  area  of  

injury.  Furthermore,  upregulation  of  growth-­‐associated  protein  43  was  observed,  a  

protein  seen  in  PNS  injury  environments  thought  to  mediate  the  regenerative  state  

(Bradbury  et  al.,  2002).    

 

Another  approach  to  studying  ways  to  promote  axonal  growth  and  block  inhibition  

involves  studying  the  intracellular  mechanisms  by  which  myelin  associated  

inhibitors  and  CPSGs  act.  Based  on  the  bifunctionality  of  MAG  depending  on  the  

stages  of  development,  both  its  growth-­‐promoting  and  growth-­‐inhibiting  actions  

were  studied.  Comparing  the  structure  of  MAG  during  these  two  stages  reveled  no  

changes,  suggesting  the  role  of  intracellular  mechanisms  mediating  the  response  to  

MAG.  It  was  found  that  during  early  stages  when  MAG  is  growth-­‐promoting,  levels  of  

cyclic  AMP  are  high,  while  during  later  stages  when  MAG  is  inhibitory,  levels  of  

cyclic  AMP  were  low  (Cai  et  al.,  2001).    This  led  to  the  idea  of  using  pharmacological  

agents  to  stimulate  cAMP  activity  in  cells  in  order  to  promote  axonal  growth.  

Through  administering  a  phosphodiesterase  inhibitor  that  prevents  the  breakdown  

of  cAMP,  scientists  were  able  to  overcome  MAG-­‐associated  inhibition  in  a  rat  model  

of  spinal  cord  injury  (Nikulina  et  al.,  2004).  Elevated  levels  of  cAMP  activate  the  

Page 21: Mini Disseration Kabir Nigam

 21  

transcription  factor  cAMP  response  element  binding  protein  (CREB)  through  

phosphorylation  of  CREB  by  cAMP-­‐activated  protein  kinase  A.  This  leads  to  the  

transcription  of  genes  that  are  responsible  for  promoting  axonal  regeneration  (Gao  

et  al.,  2004;  Hannila  &  Filbin,  2008).  

 

In  an  attempt  to  identify  other  cellular  mechanisms  by  which  CNS  myelin  inhibition  

works,  one  study  screened  for  compounds  with  the  ability  to  prevent  axonal  growth  

inhibition  and  assessed  each’s  effect  on  neuronal  outgrowth  on  an  immobilized  

myelin  substrate.  “Several  EGFR  kinase  inhibitors  showed  a  remarkable  ability  to  

counter  the  effects  of  myelin  inhibition,  suggesting  the  involvement  of  EGFR  kinase  

activity  in  the  inhibitory  effects  of  myelin  inhibitors”  (Koprivica  et  al.,  2005).  

Further  experiments  showed  that  both  Nogo-­‐A  and  Omgp  induce  rapid  EGFR  

phosphorylation,  however  indirectly  through  events  that  occur  downstream  from  

the  NgR1  receptor,  as  coimmunoprecipitation  experiments  failed  to  detect  the  

presence  of  EGFR  with  the  NgR1  receptor.  Although  how  exactly  myelin  inhibitors  

induce  EGFR  phosphorylation  is  unclear,  treatment  of  CNS  lesion  sites  with  EGFR  

inhibitors  like  Erlotinib,  which  is  approved  for  the  treatment  of  cancer,  was  able  to  

block  the  neurite  outgrowth  inhibition  caused  by  myelin-­‐associated  inhibitors  

(Koprivica  et  al.,  2005).    

 

Based  on  the  Rho/ROCK  pathway  by  which  both  myelin  associated  inhibitors  and  

CPSGs  act,  one  study  looked  at  the  effect  of  using  both  Rho  and  ROCK  inhibitors  to  

promote  axonal  regeneration  on  inhibitory  substrates.  Using  C3,  a  Rho  inhibitor,  and  

Page 22: Mini Disseration Kabir Nigam

 22  

Y27632,  a  ROCK  inhibitor,  scientists  showed  that  both  had  an  effect  of  substantially  

increasing  the  degree  of  axonal  growth  when  plated  on  substrates  containing  only  

myelin  associated  inhibitors,  only  CPSGs,  and  both  myelin  associated  inhibitors  and  

CPSGs.  Inhibition  of  Rho  showed  significantly  better  effects  on  promoting  

regeneration  than  did  inhibition  of  ROCK,  suggesting  Rho  may  affect  another  

downstream  secondary  molecule  that  is  separate  from  ROCK.  The  authors  took  this  

a  step  forward  and  applied  the  Rho  inhibitor  to  mouse  spinal  cord  injury  models  to  

test  if  inhibition  of  Rho  allows  for  functional  recovery  of  motor  function.  Use  of  C3  

allowed  for  long-­‐distance  axonal  generation  extending  past  the  lesion  into  distal  

white  matter  and  showed  a  remarkable  recovery  of  motor  functions,  allowing  mice  

to  walk  with  weight  support  as  opposed  to  control  mice  who  moved  by  pulling  

themselves  forward  with  their  forelimbs  (Dergham  et  al.,  2002)  .        

 

Looking  further  into  the  intercellular  signalling  pathways  initiated  by  myelin  

associated  inhibitors,  one  group  found  a  molecule  that  when  bound,  promotes  down  

regulation  of  a  key  protein  in  the  receptor  complex  that  mediates  myelin  associated  

inhibition.  This  molecule  is  retinoic  acid,  and  binds  to  a  receptor  known  as  the  

retinoic  acid  receptor  beta.  When  cerebellar  granule  neurons  were  cultured  with  

both  myelin  substrates  and  retinoic  acid,  neurite  outgrowth  was  promoted.  With  

knowledge  of  the  myelin  associated  inhibition  signalling  pathway,  the  group  

assessed  the  activity  of  Rho  to  see  if  retinoic  acid  was  affecting  this  pathway,  and  

found  that  Rho  activity  was  decreased  when  retinoic  acid  was  bound  to  its  receptor.  

Further  exploration  of  the  molecular  mechanisms  mediating  this  decrease  in  activity  

Page 23: Mini Disseration Kabir Nigam

 23  

found  that  retinoic  acid  binding  resulted  in  the  down  regulation  of  the  Lingo-­‐1  

protein,  a  protein  part  of  the  NgR  complex  that  is  necessary  for  myelin  associated  

inhibitors  to  transmit  their  inhibitory  signal.  This  down  regulation  is  mediated  by  

RA-­‐bound  RAR-­‐β  occupying  “a  specific  RA  response  element  (RARE)  on  the  Lingo-­‐1  

promoter,  transcriptionally  repressing  Lingo-­‐1  myelin-­‐dependent  gene  activation”  

(Puttagunta  et  al.,  2011).    

 

This  same  group  took  an  epigenetic  approach  to  study  the  changes  in  gene  

expression  associated  with  recovery  from  PNS  injury  with  hopes  of  being  able  to  

induce  the  same  changes  after  CNS  injury.  They  found  that  “p300/CBP-­‐associated  

factor  (PCAF)-­‐dependent  acetylation  of  histone  3  lysine  9  (H3K9ac),  paralleled  by  a  

reduction  in  methylation  of  H3K9  (H3K9me2),  occurred  at  the  promoters  of  select  

genes  only  after  PNS  axonal  injury”  (Puttagunta  et  al.,  2014).    This  histone  

modification  was  promoted  by  extracellular  signal-­‐regulated  kinase  acting  as  a  

retrograde  signal  to  induce  activation  of  PCAF  that  then  causes  increased  expression  

of  regeneration  associated  genes  like  H3K9ac,  Galanin  and  BDNF.  The  group  then  

tested  whether  PCAF  overexpression  in  a  spinal  cord  injury  model  would  stimulate  

axonal  regeneration,  finding  a  significant  increase  in  the  number  of  regenerating  

fibres  across  the  lesion  (Puttagunta  et  al.,  2014).    This  study  represents  a  technique  

unique  to  the  other  regeneration  strategies  presented  as  opposed  to  using  inhibitors  

or  transplantation,  this  strategy  employs  overexpression,  potentially  the  least  

harmful  method  with  regards  to  downstream  negative  effects.    

 

Page 24: Mini Disseration Kabir Nigam

 24  

CONCLUSION  

 

CNS  injuries  are  a  very  serious  matter,  as  they  often  lead  to  irreversible  motor  

defects  due  to  the  inability  of  neurons  in  the  CNS  to  regenerate  through  the  injury.  

However,  current  neuroscience  is  targeting  this  area  with  hopes  of  both  shedding  

light  on  the  mechanisms  that  make  CNS  injuries  irreversible  and  how  to  manipulate  

these  mechanisms  as  to  restore  motor  functionality  by  promoting  CNS  axonal  

regeneration.  A  number  of  the  mechanisms  underlying  the  inhibition  of  axonal  

regeneration  come  from  the  neuroinflammatory  response  to  CNS  injury.  Reactive  

astrogliosis  is  the  process  by  which  astrocytes  respond  to  CNS  injury  by  

proliferating  and  upregulating  different  molecules  that  function  to  block  off  the  

injured  site  from  being  exposed  to  healthy  tissue.  Besides  the  formation  of  the  glial  

scar  that  acts  as  a  physical  barrier  to  axonal  regeneration,  there  is  a  large  increase  in  

the  number  of  CPSGs  that  are  found  in  the  extracellular  matrix.  CPSGs  exhibit  their  

inhibitory  effects  through  diverse  and  complex  signaling  pathways  that  are  not  yet  

fully  understood.  They  also  function  to  sequester  growth-­‐promoting  molecules,  

blocking  them  from  accessing  the  damaged  tissue  to  promote  regeneration.  Other  

molecules  that  play  important  roles  during  development  are  shown  to  be  

upregulated  following  CNS  injury  and  are  likely  to  play  a  role  in  inhibition.  Such  

molecules  include  ephrins,  semaphorins,  slit  proteins,  and  tenascin.    

 

The  other  major  component  of  axonal  inhibition  after  CNS  inhibition  are  the  

inhibitory  glycoproteins  associated  with  myelin.  These  are  proteins  that  are  always  

Page 25: Mini Disseration Kabir Nigam

 25  

present  on  myelin,  and  though  they  exhibit  no  change  in  expression  following  CNS  

injury,  they  still  play  a  major  role  in  inhibiting  axonal  regeneration.  These  proteins  

often  have  a  crucial  role  developmentally,  promoting  and  directing  growth  during  

development  while  inhibiting  growth  later  on  in  life.  These  proteins  are  myelin-­‐

associated  glycoprotein  and  oligodendrocyte  myelin  glycoprotein.  Both  of  these  

molecules  exhibit  their  inhibitory  effects  through  acting  on  the  same  receptor,  the  

Nogo  receptor.  Thus,  they  all  induce  a  similar  cascade  of  intercellular  events.  This  

pathway  includes  activation  of  protein  kinase  C  that  causes  RhoA  to  activate  Rho-­‐

associated  kinase  (ROCK),  which  then  induces  actin  polymerization,  leading  to  

growth  cone  collapse  and  inhibition  of  neurite  outgrowth.    

 

Strategies  to  overcome  the  inhibition  of  axonal  growth  following  CNS  injury  have  

come  from  understanding  the  mechanisms  by  which  the  molecules  that  are  

inhibiting  axonal  regeneration  are  acting.  Strategies  include  inhibition  of  the  

molecules  directly  involved  or  more  effectively,  of  the  downstream  signaling  

cascade  that  they  induce,  transplantation  of  growth-­‐promoting  regions  to  create  a  

growth-­‐permissive  environment,  as  well  as  upregulation  or  activation  of  different  

molecules  and  genes  that  are  shown  to  induce  growth  either  during  development  or  

in  the  PNS.  Though  many  of  these  strategies  have  shown  promise  in  restoring  

functionality  to  CNS  injury  models  in  animals,  clinical  effectiveness  of  any  of  the  

mentioned  methods  has  yet  to  be  established.  One  might  conclude  that  due  to  the  

multifaceted  nature  of  CNS  inhibition,  a  combination  of  many  strategies  would  be  

the  most  effective  in  promoting  CNS  recovery  after  injury.  Because  many  of  the  

Page 26: Mini Disseration Kabir Nigam

 26  

molecules  involved  in  axonal  inhibition  generally  serve  other  functional  roles  in  the  

CNS,  it  is  important  to  establish  a  methodology  that  takes  into  account  many  factors  

such  as  the  time  course  of  administration  of  such  inhibitors  or  activators,  as  failure  

to  do  so  may  result  in  downstream  negative  effects  due  to  ineffective  signaling.  

Another  issue  that  would  be  important  to  account  for  is  how  to  direct  axonal  

regeneration  such  that  original  connectivity  is  maintained,  and  so  that  new  

connections  that  might  functionally  hinder  the  affected  individual  are  not  made.    

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Page 27: Mini Disseration Kabir Nigam

 27  

REFERENCES  

 

1. "Brain  and  Spinal  Cord  Injury  Causes,  Symptoms,  Diagnosis,  Treatment."  PDR  

Health.  Physcians  Desk  Reference,  n.d.  Web.  04  Nov.  2015.  

2. Kamiguchi,  Hiroyuki.  "Neuronal  Growth  Mechanisms."  Faculty.  Riken  Brain  

Science  Institute,  n.d.  Web.  4  Nov.  2014.  

3. Schnabelrauch,  Matthias.  "Synthesis  and  Characterization  of  ECM-­‐specific  

Polysaccharides."  Funktionelle  Biomaterialien  Zur  Steuerung  Von  

Heilungsprozessen  in  Knochen-­‐  Und  Hautgewebe  -­‐  Vom  Material  Zur  Klinik.  

University  of  Leipzig,  n.d.  Web.  04  Nov.  2015.  

4. Aarts,  M.  M.  &  Tymianski,  M.  TRPM7  and  ischemic  CNS  injury.  Neuroscientist  

11,  116–23  (2005).  

5. Aricescu,  A.,  McKinnell,  I.,  Halfter,  W.  &  Stoker,  A.  Heparan  sulfate  

proteoglycans  are  ligands  for  receptor  protein  tyrosine  phosphatase  sigma.  

Mol.  Cell.  Biol.  22,  1881-­‐92  (2002).  

6. Akbik,  F.,  Cafferty,  W.  B.  J.  &  Strittmatter,  S.  M.  Myelin  associated  inhibitors:  a  

link  between  injury-­‐induced  and  experience-­‐dependent  plasticity.  Exp.  

Neurol.  235,  43–52  (2012).  

7. Bagnard,  D.,  Lohrum,  M.,  Uziel,  D.,  Puschel,  A.  &  Bolz,  J.  Semaphorins  act  as  

attractive  and  repulsive  guidance  signals  during  the  development  of  cortical  

projections.  Development  125,  5043–5053  (1998).  

Page 28: Mini Disseration Kabir Nigam

 28  

8. Benson,  M.,  Romero,  M.,  Lush,  M.,  Lu,  Q.,  Henkemeyer,  M.  &  Parada,  L.  Ephrin-­‐

B3  is  a  myelin-­‐based  inhibitor  of  neurite  outgrowth.  Proc.  Natl.  Acad.  Sci.  U.  S.  

A.  102,  10694–9  (2005).  

9. Berry,  M.  Post-­‐injury  myelin-­‐breakdown  products  inhibit  axonal  growth:  an  

hypothesis  to  explain  the  failure  of  axonal  regeneration  in  the  mammalian  

central  nervous  system.  Bibl.  Anat.  23,  1–11  (1982).  

10. Bradbury,  E.,  Moon,  L.,  Popat,  R.,  King,  V.,  Bennett,  G.,  Patel,  P.,  Fawcett,  J.  &  

McMahon,  S.  Chondroitinase  ABC  promotes  functional  recovery  after  spinal  

cord  injury.  Nature  416,  636–40  (2002).  

11. Bundesen,  L.,  Scheel,  T.,  Bregman,  B.  &  Kromer,  L.  Ephrin-­‐B2  and  EphB2  

Regulation  of  Astrocyte-­‐Meningeal  Fibroblast  Interactions  in  Response  to  

Spinal  Cord  Lesions  in  Adult  Rats.  J.  Neurosci.  21,  7789-­‐7800  (2003).  

12. Cai,  D.,  Qiu,  J.,  Cao,  Z.,  McAtee,  M.,  Bregman,  B.  &  Filbin,  M.  Neuronal  Cyclic  

AMP  Controls  the  Developmental  Loss  in  Ability  of  Axons  to  Regenerate.  J.  

Neurosci.  21,  4731–4739  (2001).  

13. Caroni,  P.  &  Schwab,  M.  E.  Antibody  against  myelin-­‐associated  inhibitor  of  

neurite  growth  neutralizes  nonpermissive  substrate  properties  of  CNS  white  

matter.  Neuron  1,  85–96  (1988).  

14. Chen,  M.,  Huber,  A.,  van  der  Haar,  M.,  Frank,  M.,  Schnell,  L.,  Spillmann,  A.,  

Christ,  F.  &  Schwab,  M.  Nogo-­‐A  is  a  myelin-­‐associated  neurite  outgrowth  

inhibitor  and  an  antigen  for  monoclonal  antibody  IN-­‐1.  Nature.  403,  434-­‐9  

(2000).  

Page 29: Mini Disseration Kabir Nigam

 29  

15. Dergham,  P.,  Ellezam,  B.,  Essagian,  C.,  Avedissian,  H.,  Lubell,  W.  &  

McKerracher,  L.  Rho  Signaling  Pathway  Targeted  to  Promote  Spinal  Cord  

Repair.  J.  Neurosci.  22,  6570–6577  (2002).  

16. De  Winter,  F.,  Oudega,  M.,  Lankhorst,  A.,  Hamers,  F.,  Blits,  B.,  Ruitenberg,  M.,  

Pasterkamp,  R.,  Gispen,  W.  &  Verhaagen,  J.  Injury-­‐induced  class  3  semaphorin  

expression  in  the  rat  spinal  cord.  Exp.  Neurol.  175,  61–75  (2002).  

17. Domeniconi,  M.,  Cao,  Z.,  Spencer,  T.,  Sivasankaran,  R.,  Wang,  K.,  Nikulina,  E.,  

Kimura,  N.,  Cai,  H.,  Deng,  K.,  Gao,  Y.,  He,  Z.  &  Filbin,  M.  Myelin-­‐associated  

glycoprotein  interacts  with  the  Nogo66  receptor  to  inhibit  neurite  

outgrowth.  Neuron  35,  283–290  (2002).  

18. Domeniconi,  M.,  Zampieri,  N.,  Spencer,  T.,  Hilaire,  M.,  Mellado,  W.,  Chao,  M.  &  

Filbin,  M.  MAG  induces  regulated  intramembrane  proteolysis  of  the  p75  

neurotrophin  receptor  to  inhibit  neurite  outgrowth.  Neuron  46,  849-­‐855  

(2005).  

19. Edgar,  N.  &  Sibille,  E.  A  putative  functional  role  for  oligodendrocytes  in  mood  

regulation.  Transl.  Psychiatry  2,  e109  (2012).  

20. Fawcett,  J.  &  Asher,  R.  The  glial  scar  and  central  nervous  system  repair.  Brain  

Res  Bull.  49,  377-­‐91  (1999).  

21. Filbin,  M.  T.  Myelin-­‐associated  inhibitors  of  axonal  regeneration  in  the  adult  

mammalian  CNS.  Nat.  Rev.  Neurosci.  4,  703–13  (2003).  

22. Finnie,  J.  W.  &  Blumbergs,  P.  C.  Traumatic  Brain  Injury.  Vet.  Pathol.  39,  679–

689  (2002).  

Page 30: Mini Disseration Kabir Nigam

 30  

23. Fitch,  M.  T.  &  Silver,  J.  CNS  injury,  glial  scars,  and  inflammation:  Inhibitory  

extracellular  matrices  and  regeneration  failure.  Exp.  Neurol.  209,  294–301  

(2008).  

24. Fournier,  A.,  GrandPre,  T.  &  Strittmatter,  S.  Identification  of  a  receptor  

mediating  Nogo-­‐66  inhibition  of  axonal  regeneration.  Nature  409,  341–346  

(2001).  

25. Gao,  Y.,  Deng,  K.,  Hou,  J.,  Bryson,  J.,  Barco,  A.,  Nikulina,  E.,  Spencer,  T.,  Mellado,  

W.,  Kandel,  E.  &  Filbin,  M.  Activated  CREB  is  sufficient  to  overcome  inhibitors  

in  myelin  and  promote  spinal  axon  regeneration  in  vivo.  Neuron  44,  609–21  

(2004).  

26. Goldberg,  J.,  Vargas,  M.,  Wang,  J.,  Mandemakers,  W.,  Oster,  S.,  Sretavan,  D.  &  

Barres,  B.  An  oligodendrocyte  lineage-­‐specific  semaphorin,  Sema5A,  inhibits  

axon  growth  by  retinal  ganglion  cells.  J.  Neurosci.  24,  4989–99  (2004).  

27. GrandPré,  T.,  Nakamura,  F.,  Vartanian,  T.  &  Strittmatter,  S.  Identification  of  

the  Nogo  inhibitor  of  axon  regeneration  as  a  Reticulon  protein.  Nature.  403,  

439-­‐44  (2000).  

28. Hannila,  S.  S.  &  Filbin,  M.  T.  The  role  of  cyclic  AMP  signaling  in  promoting  

axonal  regeneration  after  spinal  cord  injury.  Exp.  Neurol.  209,  321–32  (2008).  

29. Huang,  L.,  Wu,  Z.,  Zhuge,  Q.,  Zheng,  W.,  Shao,  B.,  Wang,  B.,  Sun,  F.  &  Jin,  K.  Glial  

scar  formation  occurs  in  the  human  brain  after  ischemic  stroke.  Int.  J.  Med.  

Sci.  11,  344–8  (2014).  

Page 31: Mini Disseration Kabir Nigam

 31  

30. Huber,  A.,  Weinmann,  O.,  Brösamle,  C.,  Oertle,  T.  &  Schwab,  M.  Patterns  of  

Nogo  mRNA  and  protein  expression  in  the  developing  and  adult  rat  and  after  

CNS  lesions.  J.  Neurosci.  22,  3553-­‐67  (2002).  

31. Johnson,  K.  &  Van  Vactor,  D.  Receptor  protein  tyrosine  phosphatases  in  

nervous  system  development.  Physiol.  Rev.  83,  1-­‐24  (2003).  

32. Jones,  L.  L.,  Margolis,  R.  U.  &  Tuszynski,  M.  H.  The  chondroitin  sulfate  

proteoglycans  neurocan,  brevican,  phosphacan,  and  versican  are  

differentially  regulated  following  spinal  cord  injury.  Exp.  Neurol.  182,  399–

411  (2003).  

33. Kimelberg,  H.  K.  &  Nedergaard,  M.  Functions  of  astrocytes  and  their  potential  

as  therapeutic  targets.  Neurotherapeutics  7,  338–53  (2010).  

34. Kjellén,  L.  &  Lindahl,  U.  Proteoglycans:  structures  and  interactions.  Annu.  Rev.  

Biochem.  60,  443-­‐75  (1991).  

35. Koch,  C.  &  Laurent,  G.  Complexity  and  the  nervous  system.  Science  284,  96–8  

(1999).  

36. Koprivica,  V.,  Cho,  K.,  Park,  J.,  Yiu,  G.,  Atwal,  J.,  Gore,  B.,  Kim,  J.,  Lin,  E.,  Tessier-­‐

Lavigne,  M.,  Chen,  DF.  &  He,  Z.  EGFR  activation  mediates  inhibition  of  axon  

regeneration  by  myelin  and  chondroitin  sulfate  proteoglycans.  Science  310,  

106–10  (2005).  

37. Kottis,  V.,  Thibault,  P.,  Mikol,  D.,  Xiao,  Z.,  Zhang,  R.,  Dergham,  P.  &  Braun,  P.  

Oligodendrocyte-­‐myelin  glycoprotein  (OMgp)  is  an  inhibitor  of  neurite  

outgrowth.  J.  Neurochem.  82,  1566–1569  (2002).  

Page 32: Mini Disseration Kabir Nigam

 32  

38. Lai,  C.,  Watson,  J.,  Bloom,  F.,  Sutcliffe,  J.  &  Milner,  R.  Neural  protein  

1B236/myelin-­‐associated  glycoprotein  (MAG)  defines  a  subgroup  of  the  

immunoglobulin  superfamily.  Immunol.  Rev.  100,  129-­‐51  (1987)  

39. Lee,  H.,  McKeon,  R.  J.  &  Bellamkonda,  R.  V.  Sustained  delivery  of  

thermostabilized  chABC  enhances  axonal  sprouting  and  functional  recovery  

after  spinal  cord  injury.  Proc.  Natl.  Acad.  Sci.  U.  S.  A.  107,  3340–5  (2010).  

40. Li,  Y.,  Field,  P.  &  Raisman,  G.  Repair  of  Adult  Rat  Corticospinal  Tract  by  

Transplants  of  Olfactory  Ensheathing  Cells.  Science  (80-­‐.  ).  277,  2000–2002  

(1997).  

41. Lucas,  S.-­‐M.,  Rothwell,  N.  J.  &  Gibson,  R.  M.  The  role  of  inflammation  in  CNS  

injury  and  disease.  Br.  J.  Pharmacol.  147  Suppl  ,  S232–40  (2006).  

42. Mandemakers,  W.  J.  &  Barres,  B.  A.  Axon  regeneration:  it’s  getting  crowded  at  

the  gates  of  TROY.  Curr.  Biol.  15,  R302–5  (2005).  

43. McKerracher,  L.,  David,  S.,  Jackson,  D.,  Kottis,  V.,  Dunn,  R.  &  Braun,  

P.  Identification  of  myelin-­‐associated  glycoprotein  as  a  major  myelin-­‐derived  

inhibitor  of  neurite  growth.  Neuron  13,  805–811  (1994).  

44. McDonald,  J.  W.  &  Sadowsky,  C.  Spinal-­‐cord  injury.  Lancet  359,  417–25  

(2002).  

45. Mi,  S.,  Lee,  X.,  Shao,  Z.,  Thill,  G.,  Ji,  B.,  Relton,  J.,  Levesque,  M.,  Allaire,  N.,  Perrin,  

S.,  Sands,  B.,  Crowell,  T.,  Cate,  R.,  McCoy,  J.  &  Pepinsky,  R.  LINGO-­‐1  is  a  

component  of  the  Nogo-­‐66  receptor/p75  signaling  complex.  Nat.  Neurosci.  7,  

221–8  (2004).  

Page 33: Mini Disseration Kabir Nigam

 33  

46. Miranda,  J.,  White,  L.,  Marcillo,  A.,  Wilson,  C.,  Jagid,  J.  &  Whittemore,  S.  

Induction  of  Eph  B3  after  Spinal  Cord  Injury.  Exp.  Neurology.  156,  218-­‐22  

(1999).  

47. Monnier,  P.  P.,  Sierra,  A.,  Schwab,  J.  M.,  Henke-­‐Fahle,  S.  &  Mueller,  B.  K.  The  

Rho/ROCK  pathway  mediates  neurite  growth-­‐inhibitory  activity  associated  

with  the  chondroitin  sulfate  proteoglycans  of  the  CNS  glial  scar.  Mol.  Cell.  

Neurosci.  22,  319–330  (2003).  

48. Moreau-­‐Fauvarque,  C.,  Kumanogoh,  A.,  Camand,  E.,  Jaillard,  C.,  Barbin,  G.,  

Boquet,  I.,  Love,  C.,  Jones,  E.,  Kikutani,  H.,  Lubetzki,  C.,  Dusart,  I.  &  Chédotal,  A.  

The  Transmembrane  Semaphorin  Sema4D/CD100,  an  Inhibitor  of  Axonal  

Growth,  Is  Expressed  on  Oligodendrocytes  and  Upregulated  after  CNS  Lesion.  

J.  Neurosci.  23,  9229–9239  (2003).  

49. Nakajima,  K.  &  Kohsaka,  S.  Microglia:  activation  and  their  significance  in  the  

central  nervous  system.  J.  Biochem.  130,  169–75  (2001).  

50. Nie,  D.,  Ma,  Q.,  Law,  J.,  Chia,  C.,  Dhingra,  N.,  Shimoda,  Y.,  Yang,  W.,  Gong,  N.,  

Chen,  Q.,  Xu,  G.,  Hu,  Q.,  Chow,  P.,  Ng,  Y.,  Ling,  E.,  Watanabe,  K.,  Xu,  T.,  Habib,  A.,  

Schachner,  M.  &  Xiao,  Z.  Oligodendrocytes  regulate  formation  of  nodes  of  

Ranvier  via  the  recognition  molecule  OMgp.  Neuron  Glia  Biol.  2,  151–64  

(2006).  

51. Nikulina,  E.,  Tidwell,  J.  L.,  Dai,  H.  N.,  Bregman,  B.  S.  &  Filbin,  M.  T.  The  

phosphodiesterase  inhibitor  rolipram  delivered  after  a  spinal  cord  lesion  

promotes  axonal  regeneration  and  functional  recovery.  Proc.  Natl.  Acad.  Sci.  

U.  S.  A.  101,  8786–90  (2004).  

Page 34: Mini Disseration Kabir Nigam

 34  

52. Park,  J.,  Yiu,  G.,  Kaneko,  S.,  Wang,  J.,  Chang,  J.,  He,  X.,  Garcia,  K.  &  He,  Z.  A  TNF  

receptor  family  member,  TROY,  is  a  coreceptor  with  Nogo  receptor  in  

mediating  the  inhibitory  activity  of  myelin  inhibitors.  Neuron  45,  345–51  

(2005).  

53. Pasterkamp,  R.  Emerging  roles  for  semaphorins  in  neural  regeneration.  Brain  

Res.  Rev.  35,  36–54  (2001).  

54. Pasterkamp,  R.,  Anderson,  P.  &  Verhaagen,  J.  Peripheral  nerve  injury  fails  to  

induce  growth  of  lesioned  ascending  dorsal  column  axons  into  spinal  cord  

scar  tissue  expressing  the  axon  repellent  Semaphorin3A.  Eur.  J.  Neurosci.  13,  

457-­‐71  (2001).  

55. Pasterkamp,  R.  J.  &  Verhaagen,  J.  Semaphorins  in  axon  regeneration:  

developmental  guidance  molecules  gone  wrong?  Philos.  Trans.  R.  Soc.  Lond.  B.  

Biol.  Sci.  361,  1499–511  (2006).  

56. Popovich,  P.  G.  &  Longbrake,  E.  E.  Can  the  immune  system  be  harnessed  to  

repair  the  CNS?  Nat.  Rev.  Neurosci.  9,  481–93  (2008).  

57. Puttagunta,  R.,  Schmandke,  A.,  Floriddia,  E.,  Gaub,  P.,  Fomin,  N.,  Ghyselinck,  N.  

&  Di  Giovanni,  S.  RA-­‐RAR-­‐β  counteracts  myelin-­‐dependent  inhibition  of  

neurite  outgrowth  via  Lingo-­‐1  repression.  J.  Cell  Biol.  193,  1147–56  (2011).  

58. Puttagunta,  R.,  Tedeschi,  A.,  Sória,  M.,  Hervera,  A.,  Lindner,  R.,  Rathore,  K.,  

Gaub,  P.,  Joshi,  Y.,  Nguyen,  T.,  Schmandke,  A.,  Laskowski,  C.,  Boutillier,  A.,  

Bradke,  F.  &  Di  Giovanni,  S.  PCAF-­‐dependent  epigenetic  changes  promote  

axonal  regeneration  in  the  central  nervous  system.  Nat.  Commun.  5,  3527  

(2014).  

Page 35: Mini Disseration Kabir Nigam

 35  

59. Ramón-­‐Cueto,  A.  &  Nieto-­‐Sampedro,  M.  Regeneration  into  the  spinal  cord  of  

transected  dorsal  root  axons  is  promoted  by  ensheathing  glia  transplants.  

Exp.  Neurol.  127,  232–44  (1994).  

60. Ramón-­‐Cueto,  A.,  Plant,  G.  W.,  Avila,  J.  &  Bunge,  M.  B.  Long-­‐Distance  Axonal  

Regeneration  in  the  Transected  Adult  Rat  Spinal  Cord  Is  Promoted  by  

Olfactory  Ensheathing  Glia  Transplants.  J.  Neurosci.  18,  3803–3815  (1998).  

61. Rolls,  A.,  Shechter,  R.  &  Schwartz,  M.  The  bright  side  of  the  glial  scar  in  CNS  

repair.  Nat.  Rev.  Neurosci.  10,  235-­‐41  (2009).  

62. Rudge,  J.  &  Silver,  J.  Inhibition  of  neurite  outgrowth  on  astroglial  scars  in  

vitro.  J.  Neurosci.  10,  3594–3603  (1990).  

63. Sharma,  K.,  Selzer,  M.  E.  &  Li,  S.  Scar-­‐mediated  inhibition  and  CSPG  receptors  

in  the  CNS.  Exp.  Neurol.  237,  370–8  (2012).  

64. Shen,  Y.,  Tenney,  A.,  Busch,  S.,  Horn,  K.,  Cuascut,  F.,  Liu,  K.,  He,  Z.,  Silver,  J.  &  

Flanagan,  J.  PTPsigma  is  a  receptor  for  chondroitin  sulfate  proteoglycan,  an  

inhibitor  of  neural  regeneration.  Science  326,  592–6  (2009).  

65. Silver,  J.  &  Miller,  J.  Regeneration  beyond  the  glial  scar.  Nat.  Rev.  Neurosci.  5,  

146–156  (2003)  

66. Snow,  D.,  Lemmon,  V.,  Carrino,  D.,  Caplan,  A.  &  Silver,  J.  Sulfated  

proteoglycans  in  astroglial  barriers  inhibit  neurite  outgrowth  in  vitro.  Exp.  

Neurol.  109,  111-­‐30  (1990).  

67. Sofroniew,  M.  Reactive  astrocytes  in  neural  repair  and  protection.  

Neuroscientist.  11,  400-­‐7  (2005).  

Page 36: Mini Disseration Kabir Nigam

 36  

68. Stichel,  C.  C.  &  Müller,  H.  W.  The  CNS  lesion  scar:  new  vistas  on  an  old  

regeneration  barrier.  Cell  Tissue  Res.  294,  1–9  (1998).  

69. Trapp,  B.  Distribution  of  the  myelin-­‐associated  glycoprotein  and  P0  protein  

during  myelin  compaction  in  quaking  mouse  peripheral  nerve.  J.  Cell  

Biol.  107,  675–685  (1988).  

70. Turnley,  A.  &  Bartlett,  P.  MAG  and  MOG  enhance  neurite  outgrowth  of  

embryonic  mouse  spinal  cord  neurons.  Neuroreport  9,  1987–1990  (1998).  

71. Wang,  X.,  Chun,  S.,  Treloar,  H.,  Vartanian,  T.,  Greer,  C.  &  Strittmatter,  S.  

Localization  of  Nogo-­‐A  and  Nogo-­‐66  receptor  proteins  at  sites  of  axon-­‐myelin  

and  synaptic  contact.  J.  Neurosci.  22,  5505-­‐15  (2002).  

72. Wang,  K.,  Koprivica,  V.,  Kim,  J.,  Sivasankaran,  R.,  Guo,  Y.,  Neve,  R.  &  He,  Z.  

Oligodendrocyte-­‐myelin  glycoprotein  is  a  Nogo  receptor  ligand  that  inhibits  

neurite  outgrowth.  Nature  417,  941–944  (2002).  

73. Wilhelmsson,  U.,  Bushong,  E.,  Price,  D.,  Smarr,  B.,  Phung,  V.,  Terada,  M.,  

Ellisman,  M.  &  Pekny,  M.  Redefining  the  concept  of  reactive  astrocytes  as  cells  

that  remain  within  their  unique  domains  upon  reaction  to  injury.  Proc.  Natl.  

Acad.  Sci.  103,  17513–17518  (2006).  

74. Yamaguchi,  W.,  Tamai,  R.,  Kageura,  M.,  Furuyama,  T.  &  Inagaki,  S.  Sema4D  as  

an  inhibitory  regulator  in  oligodendrocyte  development.  Mol.  Cell.  Neurosci.  

49,  290-­‐99  (2012).  

75. Yamashita,  T.,  Higuchi,  H.  &  Tohyama,  M.  The  p75  receptor  transduces  the  

signal  from  myelin-­‐associated  glycoprotein  to  Rho.  J.  Cell  Biol.  157,  565–70  

(2002).  

Page 37: Mini Disseration Kabir Nigam

 37  

76. Yokoyama,  N.,  Romero,  M.,  Cowan,  C.,  Galvan,  P.,  Helmbacher,  F.,  Charnay,  P.,  

Parada,  L.  &  Henkemeyer,  M.    Forward  signaling  mediated  by  ephrin-­‐B3  

prevents  contralateral  corticospinal  axons  from  recrossing  the  spinal  cord  

midline.  Neuron.  29,  85-­‐97  (2001).