ChapterPDF Available

Targeting Inhibitory Chondroitin Sulphate Proteoglycans to Promote Plasticity After Injury

Authors:

Abstract and Figures

Chondroitin sulphate proteoglycans (CSPGs) are one of the major families of inhibitory extracellular matrix molecules in the central nervous system. The expression of various CSPGs is strong during early nervous system development; however, it is downregulated during maturation and up-regulated again after nervous system injury. In vivo injection of an enzyme called chondroitinase ABC, which removes the inhibitory chondroitin sulphate chains on the CSPGs, in the injured area promotes both the regeneration and plasticity of the neurons. Here, we describe the method of in vivo injection of the chondroitinase ABC into the cortex of adult rat brain and the histochemical method to assess the successfulness of the digestion.
Content may be subject to copyright.
127
Andrew J. Murray (ed.), Axon Growth and Regeneration: Methods and Protocols, Methods in Molecular Biology, vol. 1162,
DOI 10.1007/978-1-4939-0777-9_10, © Springer Science+Business Media New York 2014
Chapter 10
Targeting Inhibitory Chondroitin Sulphate Proteoglycans
to Promote Plasticity After Injury
Jessica C. F. Kwok , Janosch P. Heller , Rong-Rong Zhao ,
and James W. Fawcett
Abstract
Chondroitin sulphate proteoglycans (CSPGs) are one of the major families of inhibitory extracellular matrix
molecules in the central nervous system. The expression of various CSPGs is strong during early nervous
system development; however, it is downregulated during maturation and up-regulated again after nervous
system injury. In vivo injection of an enzyme called chondroitinase ABC, which removes the inhibitory
chondroitin sulphate chains on the CSPGs, in the injured area promotes both the regeneration and plastic-
ity of the neurons. Here, we describe the method of in vivo injection of the chondroitinase ABC into the
cortex of adult rat brain and the histochemical method to assess the successfulness of the digestion.
Key words Chondroitin sulphate proteoglycans , Perineuronal nets , Extracellular matrix ,
Chondroitinase ABC , Plasticity , Regeneration , Central nervous system , Spinal cord injury
1 Introduction
The extracellular matrix (ECM) in animals is mainly composed of
proteins (mainly fi brous proteins such as collagen, fi bronectin, and
laminin) and glycosaminoglycans (GAGs)/proteoglycans (PGs).
The general role of the ECM is to provide a scaffold for the struc-
ture of the surrounding tissue. However, the ECM also bears sig-
nalling properties and is involved in tissue development, maturation,
differentiation, cell adhesion and migration, cell survival, and tissue
homeostasis [
1 ]. In the central nervous system (CNS), ECM has
additional functions in regulating neurite extension together with
adhesion and migration of neurons. One major family of the ECM
molecules which controls the abovementioned processes in the
CNS is chondroitin sulphate proteoglycan (CSPG) [
2 , 3 ]. During
early development, CSPGs are involved in the shaping of the neu-
ronal circuitry either by acting directly as an inhibitory molecule or
through binding to different growth factors, therefore limiting
their availability [
4 ]. Apart from controlling circuitry formation
128
during development, CSPGs are also strongly up-regulated after
spinal cord injury [
5 , 6 ]. After CNS injury, adult neurons show
very limited potential to regenerate [
7 9 ]. This regenerative failure
is due to a loss of growth-promoting trophic factors and substrate
molecules [
10 , 11 ] and an up-regulation of growth inhibitory mol-
ecules such as CSPGs, myelin-associated molecules such as Nogo-A,
and myelin-associated glycoprotein [
5 , 8 ]. Digestion of chondroi-
tin sulphate (CS) on the CSPGs using an enzyme called chondroi-
tinase ABC (ChABC) successfully removes the CSPG inhibition
[
5 , 12 ]. Animals treated with the ChABC demonstrate signifi cant
improvement in both structural and functional recovery [
13 , 14 ].
Recently, CSPGs have also been shown to limit plasticity in the
adult nervous system [
15 17 ]. Apart from being found in the loose
ECM, CSPGs are also found surrounding the soma and proximal
dendrites in a structure called perineuronal net (PNN). PNN is a
layer of ECM which wraps on the neuronal surface, where synaptic
connections are made. PNNs are formed at the end of the critical
period when the neuronal system lost the majority of its plastic
ability [
17 , 18 ]. After CNS injury, a signifi cant proportion of func-
tional recovery is achieved through axonal sprouting and dendritic
remodelling [
19 22 ]. However, these pro-recovery events are
hampered by the up-regulated CSPGs around the injury sites and
also on the surface of the spared neurons as PNNs. The axons of
the injured neurons form dystrophic growth cones upon interac-
tion with the CSPGs [
23 ]. Once again, application of ChABC to
the relevant brain regions is able to remove this plasticity brake and
enhance recovery [
16 , 24 , 25 ]. Activation of plasticity is not just
improving recovery after CNS injury; recent studies show that
removing CSs in the PNNs may also benefi t other neurological
conditions such as dementia, memory, and learning [
26 , 27 ]. Thus,
ChABC seems to be a useful tool in enhancing regeneration and
for studying plasticity in different animal models. Here, we describe
the method of in vivo digestion of CSs in adult rat brain by ChABC
injection and the histochemical techniques used to confi rm a
successful enzymatic digestion.
2 Materials
1. Operating isofl urane (e.g., Abbott, 1–2 % in a mixture of 25 %
nitric oxide and 50 % oxygen).
2. Absolute ethanol: Dilute it with different volume of deionised
water (
D -H
2 O) for different experimental steps (such as 70 %
for disinfection).
3. Analgesics (e.g., Carprieve, Norbrook, at 5 mg/kg body
weight) and antibiotics (e.g., Terramycin, Pfi zer, 60 mg/kg).
4. Betadine.
5. Lubricant eye ointment (e.g., Allergan).
2.1 Chemicals
and Solutions
Jessica C.F. Kwok et al.
129
6. Sterile saline (e.g., Aqupharm): Warm up to 37 °C before use.
7. ChABC (protease-free form isolated from Proteus vulgaris ,
Seikagaku, Japan): Dissolve in 0.1 M sterile PBS to a fi nal con-
centration of 100 U/ml. Store the aliquots at −20 °C. On the
day of injection, thaw the aliquot on ice and keep the solution
in the icebox until injection.
8. Pentobarbitone.
1. Animals: 3-month-old adult male Sprague Dawley rats from
Charles River (UK) are used for the experiment ( see Note 1 ).
The animals are housed in groups of up to four rats under a
12-h/12-h light/dark cycle (lights on from 6 a.m. to 6 p.m.)
with room temperature (rtp) at 21 °C. Food and water are
given ad libitum. All procedures are performed in accordance
with the UK Animals (Scientifi c Procedures) Act (1986).
2. Brain atlas for stereotaxic coordinates [
28 ].
3. Stereotaxic apparatus: Stereotaxic frame for small animals
(David Kopf Instruments, UK).
4. Laboratory scale.
5. Electric hair shaver.
6. Temperature-controlled heating mat.
7. Neurosurgery materials: Sterile cotton swabs (Johnson &
Johnson, autoclaved), surgical forceps [e.g., Dumont #5 for-
ceps from Fine Science Tool (FST)], scalpel and scalpel blades,
blunt dissection scissors (FST), skin retractor (FST) or “surgi-
cal hooks” (made from clipped and bent needles), surgical drill
and small bits, Hamilton syringe (Hamilton 65459-01), verti-
cal micro-drive pump with syringe holder which can be fi xed
onto the stereotaxic frame manipulator (e.g., UMP2 pump,
World Precision Instrument), 4-0 suture needles and needle
holder (FST), 27-G needles, 1- and 2.5-mL disposable syringes
(for subcutaneous injection of the analgesic and antibiotic
solution), ligature scissors for removing sutures (FST), surgi-
cal microscope.
1. Buffered saline pre-wash (pH7.4): Dissolve 1.78 g disodium
hydrogen orthophosphate · 2H
2 O, 9 g sodium chloride in 1 l
of
D -H
2 O (pH 7.4).
2. 4 % paraformaldehyde (PFA, pH7.4): Dissolve 8 g of PFA in
200 ml of
D -H
2 O.
3. 1× phosphate-buffered saline (PBS): Dissolve 4 g of sodium
chloride, 0.1 g of potassium chloride, 0.12 g of potassium
dihydrogen phosphate, and 0.72 g of disodium hydrogen
phosphate in 500 ml of
D -H
2 O, adjust the pH to 7.4.
4. 30 % (w/v) sucrose solution: Dissolve 30 g of sucrose in
100 ml of 1× PBS.
2.2 Animals and
Stereotaxic Surgery
2.3 Perfusion
of Animals
CSPGs in Plasticity
130
1. 24-well plates.
2. Fine-hair paintbrush.
3. 1× PBS with 0.05 % sodium azide (PBS-azide): Dissolve
0.05 g of sodium azide in 100 ml of 1× PBS ( see Note 2 ).
4. TS-PBS: Prepare a 10 ml solution by mixing 0.3 ml of normal
goat serum, 0.03 ml of Triton X-100, and 9.67 ml of 1× PBS.
5. PBS-T: Prepare 100 ml of the solution by mixing 0.3 ml of
Triton X-100 with 99.7 ml of 1× PBS.
6. Biotinylated Wisteria fl oribunda agglutinin (Sigma):
Reconstitute the lectin into 1 mg/ml solution in 1× PBS.
Aliquot and store at −20 °C.
7. ABC solution: Mix one volume of solution A with one volume
of solution B in 48 volume of 1× PBS from using the reagents
in the VECTASTAIN ABC kit “Standard” (Vector Labs).
8. Tris buffered non-saline, pH 7.4 (TNS): Dissolve 6 g of
Trizma base in 1 l of
D -H
2 O. Adjust the pH to 7.4.
9. 3,3-Diaminobenzidine (DAB) solution (Sigma): Dissolve the
tablet of DAB substrate and urea/hydrogen peroxide in 5 ml
of
D -H
2 O. Keep the solution in the dark and use within 1 h
( see Note 3 ).
10. Gelatine-coated slides: Slowly dissolve 0.5 g of gelatin in
500 ml of
D -H
2 O at 56 °C in 5–10 min for 0.1 % (w/v) gelatin
solution. Wash the glass slides in 0.1 M sodium hydroxide for
15 min, then rinse with running tap water, and drain. Put the
slides into slide holders and dip into acetone 3–4 times (per-
form this step in a fume hood). Air-dry the slides for a few
minutes in the fume hood. Then dip the slides for 3–4 times
in the 0.1 % gelatin solution. Drain the excess solution, and
leave the slides in a 37 °C oven overnight. Box up and store
the slides in a dry place.
11. Xylene.
12. DPX mounting medium.
13. Microscope.
3 Methods
1. The coordinates are given as three-dimensional ( x, y, and z )
distances (in mm) from the bregma, which is the intersection of
the coronal and sagittal sutures on the surface of the skull. The
x plane represents the medial-lateral (ML) distance, the y plane
represents the anterior–posterior (AP) distance (where a prefi x
“+” indicates the anterior end and “−“as the posterior end),
and the z plane represents the dorsal–ventral (DV) distance
(where a prefi x “+” indicates the ventral side) from the bregma.
2.4 Tissue
Sectioning and
Histochemistry
3.1 Preparation
for the Stereotaxic
Frame and Injection
Coordinates
Jessica C.F. Kwok et al.
131
2. Injection coordinates (mm) for this experiment in reference to
the bregma are:
First injection: AP −0.7, ML −1.5, DV +1.5.
Second injection: AP −0.7, ML −1.4, DV +1.5 ( see Notes
4 and 5 ).
1. Before the start of the surgery, ensure that all surgical tools are
clean and sterile (either by autoclaving or by immersion in
disinfectant) and are ready to use for the operation. Clean the
operating area with 70 % ethanol.
2. Ensure that the respiration and scavenging systems are work-
ing properly and the oxygen level is suffi cient to support the
entire surgery.
3. Ensure that the operating microscope and the lights are working
properly.
4. Connect the drill and the micro-drive pump to the power
supply, and check the devices to ensure proper functions. Set
up the parameters of the pump (speed: 100 nl/min, volume:
1,000 nl).
5. Fix the drill vertically onto the stereotaxic frame probe holder,
which is then secured to the x -axis of the stereotaxic frame
manipulator through the v-shape mounting clamp. Lift the
drill up to allow suffi cient space for placing the rat.
1. All surgical procedures are performed under inhalation anes-
thetic with isofl urane.
2. Inject analgesics subcutaneously 2 h before surgery ( see Note 6 ).
3. Weigh the rat before anesthesia, and check that the animal is
in a good health condition ( see Note 7 ).
4. Anesthetize the rat in an induction box, observe closely, and
wait until the rat stops moving and the breathing rate is stable
at a reduced rate.
5. Move the rat to the shaving area, and place the nose at the
opening of the isofl urane tube ( see Note 8 ).
6. Shave the skull gently, blow away the loose hairs, and clean the
skin with Betadine.
7. Move the rat to the stereotaxic surgical stage, and switch on
the isofl urane.
8. The body temperature should be maintained at 37 °C with a
heating pad during the entire surgery.
9. Apply an eye ointment including vitamin A to protect the eyes.
10. Pinch the feet with nails to make sure that the rat reaches deep
anesthesia and no refl ex is shown. Otherwise wait for longer
time until deep anesthesia is reached.
3.2 Preparation
of Surgical Tools
and Operation Area
3.3 Preparation
of Animal and
Neurosurgery
CSPGs in Plasticity
132
11. To place the rat onto the stereotaxic frame fi x one ear bar in
the apparatus, and gently position the rat’s head so that its ear
canal is securely placed onto the ear bar. Keep the head in
place, and slowly push the second ear bar to the other ear
canal. Stop immediately when feeling that a moderate pressure
is encountered. Screw tight the second ear bar to fi x the head
in position. The head should be horizontal to the ear bars and
should not move laterally.
12. Secure the rat onto an incisor (front teeth) adapter: Gently
open the lower jaw, slowly move the incisor adapter into the
mouth, and fi t the incisors at the opening of the adapter. Pull
back the adaptor slightly, and fi x it in place. Then lower the
nose clamp onto the nose only with a low pressure ( see Note 9 ).
13. Check that the tongue is in a normal position; otherwise, pull
it out gently with small blunt forceps to prevent airway block-
age ( see Note 10 ).
14. Adjust the surgical microscope, and use low magnifi cation to
obtain a clear and complete view of the top of the skull.
15. Next, make a midline skin incision with scalpel blade 10 start-
ing from caudal to the midpoint between the eyes.
16. Retract the skin sideways with surgical hooks or skin retractor
from the incision line ( see Note 11 ).
17. Measure the z coordinates of the bregma and the lambda, and
adjust the head level using the incision adaptor so that the
z coordinates are equal at the two points.
18. Measure the x and y coordinates of the bregma, calculate the
target coordinates for injection, and adjust the manipulator to
the target position accordingly.
19. Lower the drill bit until it is almost touching the skull. Increase
the microscope magnifi cation, and adjust the focus to have a
clear view of the target area.
20. Switch on the drill, and carefully drill a hole in the skull. Stop
drilling when the blood vessels on the dura become visible and
only a very thin layer of the skull is left. Remove the thin skull
layer with a pair of fi ne forceps ( see Note 12 ).
21. Repeat the above procedure for the second coordinate.
22. Prepare for the injection: Remove the probe holder together
with the drill from the manipulator and secure the micro-drive
pump onto the x -axis via the mounting clamp.
23. Load the ChABC solution into the Hamilton syringe until the
reading reaches 2 μl for each animal ( see Note 13 ).
24. Carefully fi x the syringe onto the pump, make sure that the
syringe is secured vertically, and the scale can be seen conve-
niently. Check that the pump parameters are set correctly at
100 nl/min or 6 μl/h for 10 min.
Jessica C.F. Kwok et al.
133
25. Pierce through the dura with a sharp needle tip ( see Note 14 ).
26. Bring the Hamilton needle to the correct x and y coordinates,
and lower it until it touches the dura.
27. Slowly lower the needle to the desired z coordinate, and start
the injection. Allow the pump to run for 10 min ( see Note 15 ).
28. Remove the needle slowly to avoid a backfl ow of the solution.
29. Repeat the injection procedure for the second injection.
30. Clean the opening with moist cotton swabs, and remove all
debris including loose hair or small bone fragments.
31. Suture the skin on the skull, and inject antibiotics subcutane-
ously. Inject warm saline (37 °C) subcutaneously to avoid
dehydration of the animal after surgery.
1. Keep the rat at a warm place (in a cage with blanket or in a
recovery incubator) until fully recovered. When fully recov-
ered, return the rat to a clean cage ( see Note 16 ).
2. Check the animals every day in the following week for signs of
stress including wound infl ammation, pain, wound scratching,
suture opening, and weight loss ( see Note 17 ).
3. Remove the suture 7 days after operation under brief inhala-
tion anesthesia.
1. The post-op rat is terminally anesthetized with pentobarbi-
tone, followed by perfusion with 200 ml of cold PBS pre-wash
through the heart and then 4 % PFA ( see Note 18 ).
2.
The brain is removed and post-fi xed for 24 h at 4 °C.
3. The brain is then transferred to cold 30 % sucrose for cyropro-
tection at least overnight at 4 °C before sectioning.
1. Take out the brain from the sucrose solution, and briefl y blot
it dry on a 3 mm fi lter paper.
2. Freeze down the brain on a sledge microtome platform, and
section the brain into 35 μm thick coronal or parasagittal
sections in PBS-azide ( see Note 20 ).
3. To check if the ChABC has successfully digested the CS on the
PNNs in the brain, we use a lectin called Wisteria fl oribunda
agglutinin (biotinylated, bio-WFA) which labels the CS on
the CSPGs ( see Note 21 ).
4. The brain sections are washed in 1 ml of 1× PBS for 5 min,
three times at rtp.
5. Block the sections with TS-PBS for 1 h at rtp.
6. Remove the blocking solution, add 20 μg/ml of bio-WFA
(in TS-PBS), and incubate overnight at 4 °C.
7. Remove the lectin solution, and wash the sections in PBS-T
for 15 min, three times at rtp.
3.4 Recovery from
Anesthesia
3.5 Perfusion of Rat
3.6 Tissue
Sectioning and
Histochemistry
( See Note 19 )
CSPGs in Plasticity
134
8. While washing the sections, prepare the ABC solution
(VECTASTAIN ABC kit “Standard”) by mixing 1:1:48 of
reagent A:reagent B:1× PBS, and let it stand for at least 1/2 h at
rtp for the formation of ABC complex before use ( see Note 22 ).
9. Add the premixed ABC solution to the sections, and incubate
for 1 h at rtp.
10. Rinse the sections with TNS, three times for 15 min each.
11. Prepare the DAB solution ( see Note 23 ).
12. Incubate the sections in DAB solution, and monitor the color
intensity upon the addition of DAB substrate. Stop the reac-
tion by removing the DAB solution, and wash the sections
twice with TNS for 5 min.
13. Put the sections in a big Petri dish, and mount the sections
onto gelatin-coated slides. Air-dry the sections for 1 h.
14. Dehydrate the sections through a series of ethanol solution of
increasing percentage (50, 70, 90 % and two times of 100 %
EtOH), 5 min per solution. The slides are then incubated
twice in xylene for 5 min at rtp. The slides are coverslipped
with DPX Mounting Medium ( see Note 24 ).
15. The slides are air-dried overnight at rtp in a fume hood before
documentation using a bright-fi eld microscope (Fig.
1 ).
4 Notes
1. Other rat strains and mice can also be used for the injection.
2. Sodium azide is a toxic chemical. Please follow the supplier’s
instruction and local regulation for proper handling and disposal.
3. DAB is a strong mutagen. Please wear gloves when handling
the chemical. All solutions and consumables in contact with
the DAB have to be disposed according to the supplier’s
instruction and local regulations.
4. To obtain targeting coordinates for a specifi c injection region,
subtract the atlas coordinates from the position of the animal’s
bregma in the stereotaxic apparatus. For example, if the breg-
ma’s coordinates are “AP 60, ML 15, DV 30,” then the
desired injection coordinates are “AP 60.7, ML 16.5, DV
28.5” for the fi rst injection and “AP 60.7, ML 16.4, DV 28.5”
for the second injection.
5. The coordinates used in this protocol correspond to the senso-
rimotor cortex of the hindlimb in an adult rat. We chose this
area because hindlimb functions are mostly compromised after
lesion in various spinal cord injury models. Removing inhibitory
PNNs using ChABC opens up a window of plasticity in the
selected region and may enhance the functional recovery of the
Jessica C.F. Kwok et al.
135
Fig. 1 Section collection ( a ) and staining of chondroitin sulphates with bio- WFA ( brown ) in the adult cortex 24 h
after the injection of PBS ( b , d ) or ChABC ( c , e ). ( a ) The brain sections should be collected as a six-section series
to ensure that every stack of sections covers all levels in the brain. ( b ) and ( d ) PBS injection does not remove
the WFA-positive signal in the adult cortex. Subpopulations of neurons are clearly wrapped by perineuronal nets
under high magnifi cation in ( d ). ( c ) and ( e ) 24 h after the injection of the ChABC, a clear area devoid of brown
WFA signal (both in the extracellular matrix and on the cell surface) is observed around the injected area. High-
magnifi cation image in ( e ) shows very faint remnants of perineuronal staining on neurons present in the ChABC
area, while in the area outside the digestion zone, a very clear perineuronal staining is found on the neurons.
The asterisks mark the center of the injection sites. Scale bar in ( b ) and ( c ) is 1 mm ( d ) and ( e ) is 100 μm
CSPGs in Plasticity
136
corresponding body part in response to different treatments
and rehabilitation program.
6. The analgesic is injected subcutaneously 2 h before the sur-
gery to ensure that the optimal level of analgesic effect is
reached at the point of the surgery. It can also be injected
before the start of the surgery.
7. A good health condition is important for a speedy recovery
and to avoid unnecessary complications. If an animal is ill or
under stress before the surgery, either postpone the surgery or
choose a healthier substitute for the operation.
8. It is important to make sure that the tube is fi tted properly to
the nose of the animal. Any leakage from the tube will reduce
the effi ciency of anesthesia and the leaked isofl urane will be
released to the operation area. This poses potential harm to
the researcher.
9. Proper securing of the animal to the stereotaxic frame is criti-
cal for the operation. For detailed descriptions of the proce-
dure, please refer to ref. [
29 ].
10. Closely monitor the breathing of the animal throughout the
surgery. If the animal starts to show abnormal, slow, or labored
breathing, release the animal from the stereotaxic apparatus.
Provide air without isofl urane until the breathing recovers to a
normal pattern. Check that the oxygen is at a suffi cient level.
11. In order to reveal a clear view of the skull, scientist can gently
separate the subcutaneous tissue from the skull using blunt
dissection scissors. One can further clean the connective tissue
on the bregma and lambda areas using sterile surgical blade.
To avoid the skin tissue from drying out, apply several drops
of sterile PBS to the tissue along the incision line using wet
cotton swabs.
12. To avoid drilling through the whole skull and damaging the
cortex, make sure that you only remove a small depth of the
skull each time and stop frequently to check for the depth.
13. The total injection volume is 2 μl (1 μl for each injection
site). However, when loading the solution for the injection
setup, the dead volume of the syringe should also be included
(usually ~1 μl).
14. Hold the needle at a fl at angle to avoid damaging the brain
tissue.
15. When the injection is fi nished, leave the needle in the injection
site for another 3 min before retraction. This is to ensure that the
injected solution is fully absorbed into the surrounding tissues.
16. It is helpful to place a small dish of soaked food pellets in the
cage for easy food access after surgery.
17. The image of the histology performed in this chapter is on an
adult rat brain collected at 24 h after ChABC injection.
Jessica C.F. Kwok et al.
137
Our previous study shows that a single ChABC injection in
the cortex would effectively keep the CS at a low level for at
least 10 days after injection [
30 ].
18. A good perfusion is important for minimizing background
staining from histochemistry and to maintain a nice tissue fi xa-
tion. Perfusion with PBS pre-wash is to rinse out blood cells in
the blood vessels which tend to give strong background after
staining.
19. Histology is only one of the many ways in determining the
success of ChABC digestion. One can also assay for the resid-
ual ChABC activity recovered from the injected tissues if
necessary [
31 , 32 ].
20. The sections are usually collected as six-series stacks in the fi rst
row of a 24-well plate containing PBS-azide (Fig.
1 ). This is to
ensure that each well would have sections covering the differ-
ent levels from the brain. The sections can be stored at 4 °C
for at least 3 months.
21. WFA is a lectin which demonstrates a very strong and specifi c
binding affi nity to CSs. One can also use other antibodies to
detect CS in the area, such as clone CS-56 for CSs (Sigma)
and clone 1 F6 for neurocan (DSHB).
22. If fl uorescence detection is preferred, use fl uorophore-
conjugated streptavidin instead. After incubation for 2 h at rtp,
rinse the sections in 1× PBS three times for 15 min each, before
mounting the section with anti-fading mounting medium.
23. DAB is a toxic and carcinogenic chemical. Please follow the
provider’s instruction and local laboratory practice and regula-
tion for proper handling and disposal.
24. Xylene and DPX are very volatile solution and highly fl am-
mable. Please carry out these steps in a fume hood. Xylene is a
strong organic solvent, so use anticorrosive plasticwares or
glass containers for all the procedures.
References
1. Novak U, Kaye AH (2000) Extracellular
matrix and the brain: components and func-
tion. J Clin Neurosci 7:280–290
2. Rhodes KE, Fawcett JW (2004) Chondroitin
sulphate proteoglycans: preventing plasticity
or protecting the CNS? J Anat 204:33–48
3. Yamaguchi Y (2000) Lecticans: organizers of
the brain extracellular matrix. Cell Mol Life Sci
57:276–289
4. Kwok JC, Dick G, Wang D, Fawcett JW (2011)
Extracellular matrix and perineuronal nets in
CNS repair. Dev Neurobiol 71:1073–1089
5. Lemons ML, Howland DR, Anderson DK
(1999) Chondroitin sulfate proteoglycan
immunoreactivity increases following spinal
cord injury and transplantation. Exp Neurol
160:51–65
6. Zuo J, Neubauer D, Dyess K, Ferguson TA,
Muir D (1998) Degradation of chondroitin
sulfate proteoglycan enhances the neurite-
promoting potential of spinal cord tissue. Exp
Neurol 154:654–662
7. Bravin M, Savio T, Strata P, Rossi F (1997)
Olivocerebellar axon regeneration and tar-
get reinnervation following dissociated
Schwann cell grafts in surgically injured cere-
bella of adult rats. Eur J Neurosci 9:
2634–2649
CSPGs in Plasticity
138
8. Chen MS, Huber AB, van der Haar ME, Frank
M, Schnell L, Spillmann AA, Christ F, Schwab
ME (2000) Nogo-A is a myelin-associated
neurite outgrowth inhibitor and an antigen
for monoclonal antibody IN-1. Nature 403:
434–439
9. Morrow DR, Campbell G, Lieberman AR,
Anderson PN (1993) Differential regenerative
growth of CNS axons into tibial and peroneal
nerve grafts in the thalamus of adult rats. Exp
Neurol 120:60–69
10. Kawaja MD, Gage FH (1991) Reactive astro-
cytes are substrates for the growth of adult
CNS axons in the presence of elevated levels of
nerve growth factor. Neuron 7:1019–1030
11. Smith GM, Jacobberger JW, Miller RH (1993)
Modulation of adhesion molecule expression
on rat cortical astrocytes during maturation.
J Neurochem 60:1453–1466
12. Jones LL, Margolis RU, Tuszynski MH (2003)
The chondroitin sulfate proteoglycans neuro-
can, brevican, phosphacan, and versican are
differentially regulated following spinal cord
injury. Exp Neurol 182:399–411
13. Yick LW, Wu W, So KF, Yip HK, Shum DK
(2000) Chondroitinase ABC promotes axonal
regeneration of Clarke’s neurons after spinal
cord injury. Neuroreport 11:1063–1067
14. Bradbury EJ, Moon LD, Popat RJ, King VR,
Bennett GS, Patel PN, Fawcett JW, McMahon
SB (2002) Chondroitinase ABC promotes
functional recovery after spinal cord injury.
Nature 416:636–640
15. Galtrey CM, Asher RA, Nothias F, Fawcett JW
(2007) Promoting plasticity in the spinal cord
with chondroitinase improves functional
recovery after peripheral nerve repair. Brain
130:926–939
16. Barritt AW, Davies M, Marchand F, Hartley R,
Grist J, Yip P, McMahon SB, Bradbur y
EJ (2006) Chondroitinase ABC promotes
sprouting of intact and injured spinal systems
after spinal cord injury. J Neurosci 26:
10856–10867
17. Pizzorusso T, Medini P, Berardi N, Chierzi S,
Fawcett JW, Maffei L (2002) Reactivation of
ocular dominance plasticity in the adult visual
cortex. Science 298:1248–1251
18. McRae PA, Rocco MM, Kelly G, Brumberg
JC, Matthews RT (2006) Sensory deprivation
alters aggrecan and perineuronal net expres-
sion in the mouse barrel cortex. J Neurosci
27:5405–5413
19. Fawcett JW (2009) Recovery from spinal cord
injury: regeneration, plasticity and rehabilita-
tion. Brain 132:1417–1418
20. Fawcett JW, Curt A (2009) Damage control in
the nervous system: rehabilitation in a plastic
environment. Nat Med 15:735–736
21. Garcia-Alias G, Barkhuysen S, Buckle M,
Fawcett JW (2009) Chondroitinase ABC
treatment opens a window of opportunity for
task- specifi c rehabilitation. Nat Neurosci
12:1145–1151
22. Girgis J, Merrett D, Kirkland S, Metz GA,
Verge V, Fouad K (2007) Reaching training
in rats with spinal cord injury promotes plastic-
ity and task specifi c recovery. Brain 130:
2993–3003
23. Tom VJ, Steinmetz MP, Miller JH, Doller
CM, Silver J (2004) Studies on the develop-
ment and behavior of the dystrophic growth
cone, the hallmark of regeneration failure, in
an in vitro model of the glial scar and after spi-
nal cord injury. J Neurosci 24:6531–6539
24. Cafferty WB, Bradbury EJ, Lidierth M, Jones
M, Duffy PJ, Pezet S, McMahon SB (2008)
Chondroitinase ABC-mediated plasticity of
spinal sensory function. J Neurosci 28:
11998–12009
25. Fawcett J (2009) Molecular control of brain
plasticity and repair. Prog Brain Res 175:
501–509
26. Romberg C, Yang S, Melani R, Andrews MR,
Horner AE, Spillantini MG, Bussey TJ,
Fawcett JW, Pizzorusso T, Saksida LM (2013)
Depletion of perineuronal nets enhances
recognition memory and long-term depres-
sion in the perirhinal cortex. J Neurosci 33:
7057–7065
27. Gogolla N, Caroni P, Lüthi A, Herry C (2009)
Perineuronal nets protect fear memories from
erasure. Science 325:1258–1261
28. Paxinos G, Watson C (1986) The rat brain in
stereotaxic coordinates, 2nd edn. Academic,
San Diego
29. Cetin A, Komai S, Eliava M, Seeburg PH,
Osten P (2007) Stereotaxic gene delivery in
the rodent brain. Nat Protoc 1:3166–3173
30. Lin R, Kwok JC, Crespo D, Fawcett JW
(2008) Chondroitinase ABC has a long-lasting
effect on chondroitin sulphate glycosamino-
glycan content in the injured rat brain.
J Neurochem 104:400–408
31. Hyatt AJ, Wang D, Kwok JC, Fawcett JW,
Martin KR (2010) Controlled release of chon-
droitinase ABC from fi brin gel reduces the
level of inhibitory glycosaminoglycan chains in
lesioned spinal cord. J Control Release 147:
24–29
32. Zhao RR, Muir EM, Alves JN, Rickman H,
Allan AY, Kwok JC, Roet KC, Verhaagen J,
Schneider BL, Bensadoun JC, Ahmed SG,
Yáñez-Muñoz RJ, Keynes RJ, Fawcett JW,
Rogers JH (2011) Lentiviral vectors express
chondroitinase ABC in cortical projections
and promote sprouting of injured corticospi-
nal axons. J Neurosci Methods 201:228–238
Jessica C.F. Kwok et al.
... For example, MHCI molecules, such as PirB receptor and ligands Kb and Db, which are expressed by neurons, induce neuronal regeneration failure following stroke by a SHP-2 dependent pathway (Adelson et al., 2012). Moreover, numerous molecules that are known to inhibit tissue regeneration are present in glial scar (Kwok et al., 2014a; Silver and Miller, 2004). Among these molecules, proteoglycans overexpressed by reactive astrocytes turn the glial scar into a dense barrier impeding the sprouting of axonal cones (Kwok et al., 2012Kwok et al., , 2014a). ...
... Moreover, numerous molecules that are known to inhibit tissue regeneration are present in glial scar (Kwok et al., 2014a; Silver and Miller, 2004). Among these molecules, proteoglycans overexpressed by reactive astrocytes turn the glial scar into a dense barrier impeding the sprouting of axonal cones (Kwok et al., 2012Kwok et al., , 2014a). In this context, previous studies have shown that inhibition of astrogliosis stimulates axonal regeneration and neurites growth, not only by reducing glial scar density, but also decreasing the production of inhibitory molecules by activated astrocytes (Fawcett and Asher, 1999; Gates and Dunnett, 2001; Sandvig et al., 2004; Silver and Miller, 2004). ...
... Selective GAG chain disruption by chondroitinase ABC administration reduced the inhibitory effects of CSPGs on active growth cones, leading to greater axonal sprouting from injured and spared neurons after SCIs [204][205][206][207]. It also modulated inflammatory processes that led to neuroprotection in the injured spinal cord, including the promotion of anti-inflammatory macrophage polarization over pro-inflammatory polarization [208]. ...
Article
Full-text available
High spinal cord injuries (SCIs) lead to permanent functional deficits, including respiratory dysfunction. Patients living with such conditions often rely on ventilatory assistance to survive, and even those that can be weaned continue to suffer life-threatening impairments. There is currently no treatment for SCI that is capable of providing complete recovery of diaphragm activity and respiratory function. The diaphragm is the main inspiratory muscle, and its activity is controlled by phrenic motoneurons (phMNs) located in the cervical (C3–C5) spinal cord. Preserving and/or restoring phMN activity following a high SCI is essential for achieving voluntary control of breathing. In this review, we will highlight (1) the current knowledge of inflammatory and spontaneous pro-regenerative processes occurring after SCI, (2) key therapeutics developed to date, and (3) how these can be harnessed to drive respiratory recovery following SCIs. These therapeutic approaches are typically first developed and tested in relevant preclinical models, with some of them having been translated into clinical studies. A better understanding of inflammatory and pro-regenerative processes, as well as how they can be therapeutically manipulated, will be the key to achieving optimal functional recovery following SCIs.
... The impact of this inhibition on the inflammatory and plasticity processes that take place following injury represents a worthy research topic. This could be achieved by studying the effect of ET inhibitors on the overexpression of chondroitin sulphate proteoglycan (known to inhibit plasticity processes) (Bartus et al., 2012;Kwok et al., 2014) or on axonal sprouting or rewiring (Ballermann & Fouad, 2006;Darlot et al., 2012) and the expression of neural growth factors (Keefe et al., 2017). The potential beneficial effects of ETs should also be explored, such as ET facilitation of myelin debris clearing following injury. ...
Article
Full-text available
Spinal cord injuries involve a primary injury that can lead to permanent loss of function and a secondary injury associated with pathologic and inflammatory processes. Extracellular traps are extracellular structures expressed by immune cells that are primarily composed of chromatin, granular enzymes and histones. Extracellular traps are known to induce tissue damage when overexpressed, and could be associated in the occurrence of secondary damage. In the present study, we used flow cytometry to demonstrate that at 1‐day following a C2 spinal cord lateral hemisection in male Swiss mice, resident microglia form vital microglia extracellular traps, and infiltrating neutrophils form vital neutrophil extracellular traps. We also used immunolabelling to show that microglia near the lesion area are most likely to form these microglia extracellular traps. As expected, infiltrating neutrophils are located at the site of injury, though only some of them engage in post‐injury extracellular traps formation. We also observed the formation of microglia and neutrophil extracellular traps in our sham animal models of durotomy, but formation was less frequent than following the C2 hemisection. Our results demonstrate for the first time that microglia form extracellular traps in the spinal cord following injury and durotomy. It remains however to determine the mechanisms and kinetics of neutrophil and microglia extracellular traps formation following spinal cord injury. This information would allow to better mitigate this inflammatory process that may contribute to secondary injury, and to effectively target extracellular traps to improve functional outcomes following spinal cord injury.
... This response may be mediated by cooperation between canonical Smad3 and mTORC1 signaling. Although digesting CSPG GAG chains with the bacterial enzyme chondroitinase ABC (ChABC) or rhASB has been shown to promote axonal extension in the CNS (22,27), ChABC has failed to reach clinical trials. Therefore, the present study tried to link the synthesis of C4S with the transdifferentiation of cardiac fibroblasts. ...
Article
Full-text available
Ischemia-reperfusion infarct-derived chondroitin sulfate proteoglycans (CSPGs) are important for sustaining denervation of the infarct. Sympathetic denervation within the heart after myocardial infarction (MI) predicts the probability of a higher risk for serious ventricular arrhythmias. Chondroitin-4-sulfate (C4S) is the predominant chondroitin sulfate component in the heart. However, the mechanisms that induce CSPG expression in fibroblasts following MI remain to be elucidated. The present study found that oxygen-glucose deprivation (OGD) and TGFβ1 stimulation induced myofibroblast transformation and C4S synthesis in vitro by using reverse transcription-quantitative PCR, western blotting and immunofluorescence. MTT assay was used to detect cell viability following OGD or OGD + TGF lotreatment. Using the PI3K inhibitor ZSTK474, the Akt inhibitor MK2206, or the mTOR inhibitor AZD8055, it was observed that OGD and TGFβ1 stimulation induced myofibroblast transformation and that C4S synthesis was mTOR-dependent, whereas the upstream canonical PI3K/Akt axis was dispensable by using western blotting and immunofluorescence. siRNA knockdown of Smad3, Raptor, or Rictor, indicated that mTORC1 was critical for promoting OGD- and TGFβ1-induced myofibroblast transformation and C4S synthesis by using western blotting and immunofluorescence. This response, may be mediated via cooperation between canonical Smad3 and mTORC1 signaling. These data suggested that inhibiting myofibroblast transformation may reduce C4S synthesis. Target mTORC1 may provide additional insight into the regeneration of sympathetic nerves and the reduction of fibrosis after MI at the cellular level. These findings may contribute to the understanding of the mechanism by which C4S overproduction in the hearts of patients with MI is associated with myocardial fibrosis.
... Memory acquisition is associated with formation of new inhibitory inputs to GABAergic parvalbumin + (PV + ) neurons (2)(3)(4). The somata and dendrites of most of these PV+ neurons are surrounded by perineuronal nets (PNNs), condensed extracellular matrix structures that surround synapses and are involved in the control of developmental and adult plasticity (5)(6)(7)(8). Digestion of inhibitory CSPGs (the main effectors of PNNs) with chondroitinase ABC (ChABC), enables increased formation of inhibitory synapses on PV + neurons (2,9), and enhances memory formation and duration in young animals, as does transgenic attenuation of PNNs (10,11). Moreover in animals with defective memory due to pathology from mutant tau or amyloid-beta, CSPG digestion or antibody treatment to block inhibitory 4-sulphated CS-GAGs restores memory to normal levels (12)(13)(14). ...
Preprint
Full-text available
Memory loss is a usual consequence of ageing. In aged brains, perineuronal nets (PNNs), which limit neuroplasticity and are implicated in memory, become inhibitory due to decreased 6-sulphation of their glycan chains (C6S). Aged mice show progressive deficits in memory tasks, but removal of PNNs or digestion of their glycans rescued age-related memory loss. Reduction of permissive C6S by transgenic deletion of 6-sulfotransferase led to very early memory loss. However, restoring C6S levels in aged animals by AAV delivery or transgenic expression of 6-sulfotransferase restored memory. Low C6S levels caused loss of cortical long-term potentiation, which was restored by AAV-mediated 6-sulfotransferase delivery. The study shows that loss of C6S in the aged brain leads to declining memory and cognition, which can be restored by C6S replacement.
... Important impediments that form the basis of this phenomenon are proteins expressed in CNS myelin (e.g. Nogo-A), which inhibit neurite growth, and the formation of a glial scar that contains extracellular matrix molecules such as chondroitin sulphate proteoglycans (Kwok et al. , 2014). Apart from the failure to regenerate, plastic 'hardware' changes in the adult CNS of mammals and humans are restricted. ...
... 14 Conversely, mice lacking the CS-rich proteoglycan (CSPG) receptor PTP-σ displayed reduced arrhythmia susceptibility. 15,16 In the central nervous system, CS degradation in vivo mediated by the bacterial enzyme chondroitinase ABC [17][18][19] or rhASB 20 critically promoted postinjury functional recovery. We therefore hypothesized a role for CS in nonmucopolysaccharidosis pathological cardiac remodeling and tested the utility of CS targeting in heart failure treatment. ...
Article
Background -Heart failure (HF) is a leading cause of mortality and morbidity, and the search for novel therapeutic approaches continues. In the monogenic disease mucopolysaccharidosis (MPS) VI, loss of function mutations in arylsulfatase B (ASB) leads to myocardial accumulation of chondroitin sulfate (CS) glycosaminoglycans (GAGs), manifesting as a myriad of cardiac symptoms. Here, we studied changes in myocardial CS in non-MPS failing hearts, and assessed its generic role in pathological cardiac remodeling. Methods -Healthy and diseased human and rat left ventricles were subjected to histological and immuno-staining methods to analyze for GAG distribution. GAGs were extracted and analyzed for quantitative and compositional changes using Alcian Blue assay and liquid chromatography mass spectrometry. Expression changes in 20 CS-related genes were studied in three primary human cardiac cell types and THP-1 derived macrophages under each of 9 in vitro stimulatory conditions. In two rat models of pathological remodeling induced by transverse aortic constriction (TAC) or isoprenaline infusion, recombinant human arylsulfatase B (rhASB), clinically used as enzyme replacement therapy (ERT) in MPS VI, was administered intravenously for 7 or 5 weeks respectively. Cardiac function, myocardial fibrosis and inflammation were assessed by echocardiography and histology. CS-interacting molecules were assessed using surface plasmon resonance and a mechanism of action was verified in vitro Results -Failing human hearts displayed significant perivascular and interstitial CS accumulation, particularly in regions of intense fibrosis. Relative composition of CS disaccharides remained unchanged. Transforming growth factor β (TGFβ) induced CS upregulation in cardiac fibroblasts. CS accumulation was also observed in both the pressure-overload and the isoprenaline models of pathological remodeling in rats. Early treatment with rhASB in the TAC model, and delayed treatment in the isoprenaline model, proved rhASB to be effective at preventing cardiac deterioration and augmenting functional recovery. Functional improvement was accompanied by reduced myocardial inflammation and overall fibrosis. Tumor necrosis factor α (TNFα) was identified as a direct binding partner of CS GAG chains, and rhASB reduced TNFα-induced inflammatory gene activation in vitro in endothelial cells and macrophages. Conclusions -CS GAGs accumulate during cardiac pathological remodeling, and mediate myocardial inflammation and fibrosis. RhASB targets CS effectively as a novel therapeutic approach for the treatment of heart failure.
... The physiological mechanisms behind the decline in CNS function with ageing are still incompletely understood. Our previous work has demonstrated a link between CS-GAGs, PNNs, plasticity and memory [18,19,46]. Thus digesting brain CS-GAGs or removing PNNs can extend memory in normal rodents, or restore memory in models of Alzheimer's disease and ageing [18,19] (Yang et al. unpublished observations). ...
Article
Full-text available
Chondroitin sulfate (CS) proteoglycans in perineuronal nets (PNNs) from the central nervous system (CNS) are involved in the control of plasticity and memory. Removing PNNs reactivates plasticity and restores memory in models of Alzheimer's disease and ageing. Their actions depend on the glycosaminoglycan (GAG) chains of CS proteoglycans, which are mainly sulfated in the 4 (C4S) or 6 (C6S) positions. While C4S is inhibitory, C6S is more permissive to axon growth, regeneration and plasticity. C6S decreases during critical period closure. We asked whether there is a late change in CS-GAG sulfation associated with memory loss in aged rats. Immunohistochemistry revealed a progressive increase in C4S and decrease in C6S from 3 to 18 months. GAGs extracted from brain PNNs showed a large reduction in C6S at 12 and 18 months, increasing the C4S/C6S ratio. There was no significant change in mRNA levels of the chondroitin sulfotransferases. PNN GAGs were more inhibitory to axon growth than those from the diffuse extracellular matrix. The 18-month PNN GAGs were more inhibitory than 3-month PNN GAGs. We suggest that the change in PNN GAG sulfation in aged brains renders the PNNs more inhibitory, which lead to a decrease in plasticity and adversely affect memory.
Chapter
The spinal cord has a limited potential to regenerate after traumatic injury, due to extensive tissue destruction and a natural healing process, which creates physical, cellular, and molecular barriers in an attempt to preserve residual function. In this chapter, we focus on tissue-engineering approaches that use biomaterials as combinatorial therapies for spinal cord injury (SCI) repair. SCI is first described as the disruption of anatomic organization, along with its clinical and epidemiological consequences. Key tissue-engineering principles to facilitate spinal cord regeneration using biomaterial platforms in multimodal approaches are discussed. Bioengineering considerations for material fabrication in macro- and microarchitectures, and for integrated biocompatibility in animal models of SCI are addressed. The applications of natural and synthetic polymer scaffolds with cellular and molecular functionalization, and their outcomes in regenerating spinal cord tissue after injury, are presented in detail. The chapter concludes with perspectives on the clinical translation of these technologies to improve neurologic function in patients with SCI.
Article
Full-text available
Spinal cord injury (SCI) typically causes devastating neurological deficits, particularly through damage to fibers descending from the brain to the spinal cord. A major current area of research is focused on the mechanisms of adaptive plasticity that underlie spontaneous or induced functional recovery following SCI. Spontaneous functional recovery is reported to be greater early in life, raising interesting questions about how adaptive plasticity changes as the spinal cord develops. To facilitate investigation of this dynamic, we have developed a SCI model in the neonatal mouse. The model has relevance for pediatric SCI, which is too little studied. Because neural plasticity in the adult involves some of the same mechanisms as neural plasticity in early life(1), this model may potentially have some relevance also for adult SCI. Here we describe the entire procedure for generating a reproducible spinal cord compression (SCC) injury in the neonatal mouse as early as postnatal (P) day 1. SCC is achieved by performing a laminectomy at a given spinal level (here described at thoracic levels 9-11) and then using a modified Yasargil aneurysm mini-clip to rapidly compress and decompress the spinal cord. As previously described, the injured neonatal mice can be tested for behavioral deficits or sacrificed for ex vivo physiological analysis of synaptic connectivity using electrophysiological and high-throughput optical recording techniques(1). Earlier and ongoing studies using behavioral and physiological assessment have demonstrated a dramatic, acute impairment of hindlimb motility followed by a complete functional recovery within 2 weeks, and the first evidence of changes in functional circuitry at the level of identified descending synaptic connections(1).
Article
Full-text available
Perineuronal nets (PNNs) are extracellular matrix structures surrounding cortical neuronal cell bodies and proximal dendrites and are involved in the control of brain plasticity and the closure of critical periods. Expression of the link protein Crtl1/Hapln1 in neurons has recently been identified as the key event triggering the formation of PNNs. Here we show that the genetic attenuation of PNNs in adult brain Crtl1 knock-out mice enhances long-term object recognition memory and facilitates long-term depression in the perirhinal cortex, a neural correlate of object recognition memory. Identical prolongation of memory follows localized digestion of PNNs with chondroitinase ABC, an enzyme that degrades the chondroitin sulfate proteoglycan components of PNNs. The memory-enhancing effect of chondroitinase ABC treatment attenuated over time, suggesting that the regeneration of PNNs gradually restored control plasticity levels. Our findings indicate that PNNs regulate both memory and experience-driven synaptic plasticity in adulthood.
Article
Full-text available
A perineuronal net (PNN) is a layer of lattice-like matrix which enwraps the surface of the soma and dendrites, and in some cases the axon initial segments, in sub-populations of neurons in the central nervous system (CNS). First reported by Camillo Golgi more than a century ago, the molecular structure and the potential role of this matrix have only been unraveled in the last few decades. PNNs are mainly composed of hyaluronan, chondroitin sulfate proteoglycans, link proteins, and tenascin R. The interactions between these molecules allow the formation of a stable pericellular complex surrounding synapses on the neuronal surface. PNNs appear late in development co-incident with the closure of critical periods for plasticity. They play a direct role in the control of CNS plasticity, and their removal is one way in which plasticity can be re-activated in the adult CNS. In this review, we examine the molecular components and formation of PNNs, their role in maturation and synaptic plasticity after CNS injury, and the possible mechanisms of PNN action.
Article
Full-text available
Several diseases and injuries of the central nervous system could potentially be treated by delivery of an enzyme, which might most effectively be achieved by gene therapy. In particular, the bacterial enzyme chondroitinase ABC is beneficial in animal models of spinal cord injury. We have adapted the chondroitinase gene so that it can direct secretion of active chondroitinase from mammalian cells, and inserted it into lentiviral vectors. When injected into adult rat brain, these vectors lead to extensive secretion of chondroitinase, both locally and from long-distance axon projections, with activity persisting for more than 4 weeks. In animals which received a simultaneous lesion of the corticospinal tract, the vector reduced axonal die-back and promoted sprouting and short-range regeneration of corticospinal axons. The same beneficial effects on damaged corticospinal axons were observed in animals which received the chondroitinase lentiviral vector directly into the vicinity of a spinal cord lesion.
Article
Full-text available
Adult Fears Why are fear memories almost impossible to get rid of—even with extensive extinction training? Animal studies have shown that the efficacy of extinction learning depends on age. Fear memories in young animals can be permanently erased, but in adults they can be easily recovered after extinction training. Perineuronal nets, the highly organized form of extracellular matrix around inhibitory neurons, mediate the shift from juvenile to adult forms of learning in sensory systems. Gogolla et al. (p. 1258 ; see the Perspective by Pizzorusso ) have discovered that the formation of perineuronal nets in the amygdala coincides with the developmental shift in the ability to erase fear memories by extinction. Removal of perineuronal nets in adult animals re-enabled the erasure of fear memories. Thus, in adults it appears that fear memories are actively protected from erasure by the perineuronal nets.
Article
The ability of Schwann cells to induce the regeneration of severed olivocerebellar and Purkinje cell axons across an injury up to their deafferented targets was tested by transplanting freshly dissociated cells from newborn rat sciatic nerves into surgically lesioned adult cerebella. The grafted glial cells consistently filled the lesion gap and migrated into the host parenchyma. Transected olivocerebellar axons vigorously regenerated into the graft, where their growth pattern and direction followed the arrangement of Schwann cell bundles. Although some of these axons terminated within the transplant, many of them rejoined the cerebellar parenchyma beyond the lesion. Here, their fate depended on the territory encountered. No growth occurred in the white matter. Numerous fibres penetrated into the granular layer and formed terminal branches that remained confined within this layer. A few of them, however, regenerated up to the molecular layer and formed climbing fibres on Purkinje cell dendrites. By contrast, the growth of transected Purkinje cell axons into the grafts was very poor. These results underscore the different intrinsic responsiveness of Purkinje cell and olivocerebellar axons to the growth-promoting action of Schwann cells, and show that the development and outcome of the regenerative phenomena is strongly conditioned by the spatial organization and specific features of the environmental cues encountered by the outgrowing axons along the course they follow. However, Schwann cells effectively bridge the lesion gap, induce the regeneration of olivocerebellar axons, and direct their growth up to the deafferented host cortex, where some of them succeed in reinnervating their natural targets.
Article
Chondroitinase ABC (ChABC) is a bacterial enzyme that can enhance plasticity following injury to the central nervous system (CNS) by degrading the glycosaminoglycan (GAG) side chains of proteoglycans. CNS lesions treated with ChABC often show enhanced axonal sprouting and improved functional recovery and there is therefore much interest in the potential use of ChABC as a clinical treatment in humans. When highly concentrated fibrin gel containing ChABC was implanted adjacent to a spinal cord lesion, bioactive ChABC was detectable in the spinal cord for at least three weeks. Nearly six times more bioactive ChABC was detected in the spinal cord 3 weeks after injury when the fibrin delivery system was used vs. an intraspinal injection of ChABC (61+/-30 mU vs. 11+/-4 mU). Furthermore, 3 weeks after injury the level of inhibitory GAG found in injured spinal cord treated with the delivery system was 37% lower than the level of GAG in spinal cord treated with an injection of ChABC. When using the delivery system, 24.4% of the initial ChABC dose could still be detected in the lesion after 3 weeks, compared to just 4.4% when using an intraspinal injection of ChABC.
Article
Chondroitinase ABC treatment promotes spinal cord plasticity. We investigated whether chondroitinase-induced plasticity combined with physical rehabilitation promotes recovery of manual dexterity in rats with cervical spinal cord injuries. Rats received a C4 dorsal funiculus cut followed by chondroitinase ABC or penicillinase as a control. They were assigned to two alternative rehabilitation procedures, the first reinforcing skilled reaching and the second reinforcing general locomotion. Chondroitinase treatment enhanced sprouting of corticospinal axons independently of the rehabilitation regime. Only the rats receiving the combination of chondroitinase and specific rehabilitation showed improved manual dexterity. Rats that received general locomotor rehabilitation were better at ladder walking, but had worse skilled-reaching abilities than rats that received no treatment. Our results indicate that chondroitinase treatment opens a window during which rehabilitation can promote recovery. However, only the trained skills are improved and other functions may be negatively affected.