Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Received: 31 March 2019 Revised: 24 July 2019 Accepted: 1 August 2019

DOI: 10.1002/glia.23706

REVIEW ARTICLE

The fate and function of oligodendrocyte progenitor cells


after traumatic spinal cord injury

Greg J. Duncan1 | Sohrab B. Manesh2 | Brett J. Hilton3 | Peggy Assinck4 |


Jason R. Plemel5 | Wolfram Tetzlaff2,6

1
Department of Neurology, Jungers Center for
Neurosciences Research, Oregon Health and Abstract
Science University, Portland, Oregon Oligodendrocyte progenitor cells (OPCs) are the most proliferative and dispersed
2
Graduate Program in Neuroscience,
population of progenitor cells in the adult central nervous system, which allows these
International Collaboration on Repair
Discoveries (ICORD), University of British cells to rapidly respond to damage. Oligodendrocytes and myelin are lost after trau-
Columbia (UBC), Vancouver, British Columbia,
matic spinal cord injury (SCI), compromising efficient conduction and, potentially, the
Canada
3
Deutsches Zentrum für Neurodegenerative long-term health of axons. In response, OPCs proliferate and then differentiate into
Erkrankungen (DZNE), Bonn, Germany new oligodendrocytes and Schwann cells to remyelinate axons. This culminates in
4
MRC Centre for Regenerative Medicine,
highly efficient remyelination following experimental SCI in which nearly all intact
University of Edinburgh, Edinburgh, UK
5
Department of Medicine, Division of
demyelinated axons are remyelinated in rodent models. However, myelin regenera-
Neurology, Neuroscience and Mental Health tion comprises only one role of OPCs following SCI. OPCs contribute to scar forma-
Institute, University of Alberta, Calgary,
Alberta, Canada
tion after SCI and restrict the regeneration of injured axons. Moreover, OPCs alter
6
Departments of Zoology and Surgery, their gene expression following demyelination, express cytokines and perpetuate the
University of British Columbia, Vancouver, immune response. Here, we review the functional contribution of myelin regenera-
British Columbia, Canada
tion and other recently uncovered roles of OPCs and their progeny to repair
Correspondence following SCI.
Wolfram Tetzlaff, International Collaboration
on Repair Discoveries (ICORD), University of
British Columbia, Blusson Spinal Cord Centre, KEYWORDS
818 West 10th Avenue, Vancouver, BC V5Z oligodendrocyte, oligodendrocyte progenitor cell, remyelination, Schwann cell, spinal cord
1M9, Canada. injury
Email: tetzlaff@icord.org

1 | I N T RO D UC T I O N et al., 2018; Plemel et al., 2014). Preclinical data show that as the milieu
of the spinal cord becomes less toxic (Crowe et al., 1997; Williams et al.,
Traumatic spinal cord injuries (SCIs) often leave patients with permanent 2014), a period of remyelination follows (Blight, 1985; Hesp, Goldstein,
sensory and motor disabilities alongside long-term health problems. The Miranda, Kaspar, & McTigue, 2015; Powers et al., 2012). Even when
pathobiology of SCI can be split into the initial trauma (primary damage) motor and sensory function is completely lost below the site of injury in
followed by a secondary phase. The secondary phase of injury involves a humans, the majority of injuries are anatomically incomplete (Kakulas,
cascade of ischemia, inflammation, and cell death including, and of 1988) and up to a hundred thousand axons may be spared in humans
importance in this review, oligodendrocytes (Crowe, Bresnahan, Shu-
(Kakulas & Kaelan, 2015). Many of these surviving axons are stripped of
man, Masters, & Beattie, 1997; Hilton, Moulson, & Tetzlaff, 2017; Kwon,
their myelin (demyelinated; Bunge, Puckett, Becerra, Marcillo, &
Tetzlaff, Grauer, Beiner, & Vaccaro, 2004; Norenberg, Smith, & Marcillo,
Quencer, 1993; Guest, Hiester, & Bunge, 2005; Norenberg et al., 2004)
2004). Oligodendrocytes are susceptive to damage from reactive oxygen
likely due to the loss of oligodendrocytes. Accordingly, many have
species, excitotoxicity, extracellular ATP, and pro-inflammatory cyto-
hypothesized that myelin regeneration is a plausible therapeutic target
kines all of which are present following SCI (Crowe et al., 1997; Giacci
to promote signal conduction and protect axons from secondary damage
Greg J. Duncan and Sohrab B. Manesh authors contributed equally to this work. (Alizadeh, Dyck, & Karimi-Abdolrezaee, 2015; Myers, Bankston, Burke,

Glia. 2019;1–19. wileyonlinelibrary.com/journal/glia © 2019 Wiley Periodicals, Inc. 1


2 DUNCAN ET AL.

Ohri, & Whittemore, 2016; Papastefanaki & Matsas, 2015; Plemel et al., the administration of tetrodotoxin (a voltage-gated sodium channel
2014). However, there is no direct evidence that oligodendrocyte blocker) to block neuronal activity (Gautier et al., 2015). While few
remyelination contributes to the partial spontaneous functional recovery studies have modulated these synapse-like connections between
after SCI. Despite this, transplantation of myelin-forming cells is/has neurons and OPCs in vivo during traumatic injury or demyelination,
been used in several clinical trials of SCI (https://clinicaltrials.gov/). administration of tetrodotoxin or α-amino-3-hydroxy-5-methyl-
Moreover, oligodendrocyte lineage cells have diverse roles follow- 4-isoxazolepropionic acid receptor (AMPAR) inhibitors impairs
ing traumatic SCI beyond oligodendrocyte remyelination. They pro- remyelination (Gautier et al., 2015). Synaptic connections also form
duce remyelinating Schwann cells, contribute to glial scaring where between NG2+ glia and dystrophic axons after SCI, and might restrict
they alter the growth of axons and may even regulate immune axon growth (Filous et al., 2014) and is thoroughly discussed below.
responses. Here, we describe the diverse roles that oligodendrocyte OPCs are persistently monitoring their environment and are highly
lineage cells play in the lesion environment with a focus on how these responsive to changes in neurons in health and disease.
roles may contribute to the spontaneous locomotor recovery OPCs form oligodendrocytes not only during development, but
observed after SCI. also throughout adulthood (Hill, Li, & Grutzendler, 2018; Hughes,
Orthmann-Murphy, Langseth, & Bergles, 2018; Young et al., 2013). In
addition to de-novo oligodendrocytes, changes to established myelin
1.1 | OPCs during development and in adulthood
occurs throughout life and can modulate axonal conductance. In vivo
During development, oligodendrocyte progenitor cells (OPCs; expre- imaging using label-free spectral confocal reflectance microscopy
ssing the markers PDGFRα and NG2) are first observed in the ventricular showed ~19% of established myelin sheaths either retract or extend
germinal zones of the brain and spinal cord at embryonic day 12.5 during adulthood in mice (Hill et al., 2018). The thickness of individual
(E12.5) in mice (Huang, Guo, Bai, Scheller, & Kirchhoff, 2019; Pringle & myelin sheaths can change as well: overactivation of mTOR signaling
Richardson, 1993; Timsit et al., 1995; Warf, Fok-Seang, & Miller, 1991) in adult oligodendrocytes increases myelin thickness and wrap num-
while a second wave of these cells originate from the dorsal cord at ber (Snaidero et al., 2014), optogenetic or chemogenetic activation of
E15.5 (Cai et al., 2005; Fogarty, Richardson, & Kessaris, 2005; Vallstedt, neurons enhances the thickness of their myelin (Gibson et al., 2014;
Klos, & Ericson, 2005). Ultimately, these two populations of OPCs dis- Mitew et al., 2018), whereas social isolation reduces adult myelination
play similar properties (Marques et al., 2018; Tripathi et al., 2011) and of the prefrontal cortex (Liu et al., 2012; Makinodan, Rosen, Ito, &
both differentiate to form the myelinating cell of the central nervous sys- Corfas, 2012). However, whether myelin sheaths can become thinner
tem (CNS), the oligodendrocyte. However, there is some heterogeneity is not well-established. Cleavage of neurofascin 155 by thrombin at
during adulthood in the electrophysiological properties of OPCs the paranode reduces myelin thickness and increases nodal size and
depending on the region of the CNS in which they are located (Foerster,
may be blocked by astrocytic exocytosis of thrombin protease inhibi-
Hill, & Franklin, 2019; Spitzer et al., 2019; van Bruggen, Agirre, &
tors (Dutta et al., 2018). The idea that perinodal astrocytes can modu-
Castelo-Branco, 2017).
late myelin dynamics is very novel, but whether such a mechanism is
In the adult, OPCs tile throughout the entire CNS through a mech-
critical for adaptive changes in myelin is still yet to be determined. An
anism of self-repulsion (Hughes, Kang, Fukaya, & Bergles, 2013).
additional possibility is that breakdown of the blood brain barrier
Branching and highly motile filopodia survey the environment to allow
(BBB) during SCI or in MS lesions, may result in vascular thrombin
OPCs to quickly detect and replace lost or differentiated OPCs (Hill,
entering the parenchyma and inducing cleavage of NF155 breaking
Patel, Goncalves, Grutzendler, & Nishiyama, 2014; Hughes et al.,
the connection of the paranodal loops to the axon, and possibly
2013). Additionally, OPCs also form close connections with neurons
resulting in demyelination. Whether such a mechanism drives demye-
which allow them to monitor changes in neuronal activity (Bergles,
lination following SCI, requires explicit validation. Nevertheless, there
Roberts, Somogyi, & Jahr, 2000; Gallo, Mangin, Kukley, & Dietrich,
is now compelling data that myelination is highly dynamic throughout
2008; Hill et al., 2014; Hughes et al., 2013). These connections have
adulthood and can be changed in response to environment stimuli and
properties similar to traditional neuron–neuron synapses, including
learning paradigms.
accumulation of vesicles in presynaptic terminals apposed from OPCs,
expression of neurotransmitter receptors on OPCs akin to postsynap-
tic neurons, release of vesicles into the intersynaptic space, and sub- 1.2 | Oligodendrocytes support axonal health and
sequent fast-kinetic currents in OPCs (Bergles et al., 2000; Gallo et al., function
1996; Hamilton et al., 2017; Karadottir, Hamilton, Bakiri, & Attwell,
1.2.1 | Myelin and conduction
2008; Kukley, Capetillo-Zarate, & Dietrich, 2007; Ziskin, Nishiyama,
Rubio, Fukaya, & Bergles, 2007), and has been thoroughly reviewed Myelin facilitates rapid signal conduction in axons. As a multilamellar
elsewhere (Almeida & Lyons, 2014; Bergles, Jabs, & Steinhauser, membrane, myelin increases the resistance and lowers the capaci-
2010; de Faria Jr., Pama, Evans, Luzhynskaya, & Karadottir, 2018). In tance of the nerve fiber. These insulating properties allow for rapid
parallel to development where OPCs have synaptic input from unmy- impulse propagation, with current being reinitiated at the unmyelin-
elinated axons (Kukley et al., 2007; Ziskin et al., 2007), demyelination ated nodes of Ranvier in a process called “saltatory conductance”
increases excitatory currents within OPCs, which are suppressed by (Lillie, 1925). This permits the rapid propagation of action potentials in
DUNCAN ET AL. 3

even relatively small caliber axons, minimizing the space necessary for demands to the axons rather than spreading the ATP usage between
fast conduction of signals. Subtle myelin changes can also fine tune separate cells.
impulse propagation. Factors such as internodal length, myelin thick- The increased energy demands of the demyelinated axon may cause
ness, or partial myelination of single axons can be modulated to alter an energy crisis ultimately leading to their degeneration. If ATP demands
conduction velocity throughout life (Etxeberria et al., 2016; Ford are not met, Na+/K+ ATPases will not effectively remove Na+ ions thus
et al., 2015; Hill et al., 2018; Hughes et al., 2018). After SCI, demyelin- reversing the Na/Ca exchanger and causing an influx of Ca2+ ions (Stys,
ation may silence or slow action potential propagation (Blight, 1983a; Waxman, & Ransom, 1991). This overload of intracellular Ca2+ ions acti-
Hains, Saab, Lo, & Waxman, 2004; James et al., 2011; Nashmi & Feh- vates calcium dependent systems such as calpains, which in turn can lead
lings, 2001) and potentially contribute to functional deficits. to degradation of the cytoskeleton, compromising axonal structure
It is still unclear the extent to which conduction failure after SCI stems (Figure 2; Liu, Yin, Zhang, & Qian, 2014; Posmantur et al., 1997). Increases
from demyelination or other factors. Acutely after SCI, even when axons to intracellular Ca2+ can also trigger a mitochondrial permeability transi-
are spared, conduction within the spinal cord is severely diminished and tion that leads to depolarization of mitochondrial membranes, impairing

correlates with temporary loss of spinal cord reflexes in humans, a poorly mitochondrial function, and contributing to axonal degeneration

understood mechanism termed clinically as spinal shock (Atkinson & (Bernardi, 1992; Lehninger, Vercesi, & Bababunmi, 1978). As many white

Atkinson, 1996; James et al., 2011; Smith & Jeffery, 2005). Within the matter axons are frequently active, being forced to maintain and replenish

first 2 weeks, electrophysiological recordings of the spinal cord show at an ion gradient over such a large surface area can lead to ATP depletion
and a state of “virtual hypoxia” that eventually leads to axonal death (see
least some conduction is restored (Beaumont, Onifer, Reed, & Magnuson,
review by Trapp & Stys, 2009). Given the slow movement of molecules
2006; Hains et al., 2004; Hubscher & Johnson, 2006; James et al., 2011;
within the cytoplasm (Oblinger, Foe, Kwiatkowska, & Kemp, 1988) and
Metz et al., 2000). This is, however, temporally discordant with oligoden-
that oligodendrocyte myelin physically separates axons from vasculature,
drocyte remyelination, which largely occurs after the first 2 weeks in
they must receive metabolites locally, at least during times of high-energy
experimental models (Figure 1; Duncan, Manesh, et al., 2018; James et al.,
demand (Nave, 2010). Although astrocytes physically contact myelinated
2011), though it is plausible that later improvements in the speed of con-
neurons, these interactions are limited to cell bodies, synapses, and nodes
duction are derived from remyelination. However, conduction block can
of Ranvier, whereas oligodendrocytes make considerably more contact
also be mediated by nitric oxide and extracellular molecules like NG2
with white matter axons along the entire internode. This puts oligoden-
released in the wake of trauma (Hunanyan et al., 2010; Redford, Kapoor, &
drocytes in a favorable position to provide metabolic support to axons.
Smith, 1997; Shrager, Custer, Kazarinova, Rasband, & Mattson, 1998).
MCT1, a transporter of lactate, is highly enriched in oligodendrocytes and
Intraspinal injections of anti-NG2 antibodies resulted in improved axonal
its disruption is sufficient to cause axon loss (Lee et al., 2012). Metabolic
conductance in a chronic hemisection injury model (Petrosyan et al.,
support from oligodendrocytes to axons is integral to their function and
2013). Compounded with the ability of demyelinated axons to eventually
merely impairing axon-oligodendrocyte interactions, even without out-
conduct without myelin (Felts, Baker, & Smith, 1997), and rather efficient
right demyelination, is often enough to induce axon loss (Edgar et al.,
remyelination of axons in rodents (Duncan, Radcliff, et al., 2018; James
2009; Griffiths et al., 1998; Snaidero et al., 2017). Collectively these data
et al., 2011), it remains unclear whether demyelination is causing persis-
make it clear that axons depend on oligodendrocytes for their function
tent conduction deficits after SCI.
and health, and oligodendrocyte deficiency may leave axons vulnerable
to loss.
1.2.2 | Oligodendrocytes and trophic support However, it remains unclear if demyelination and oligodendrocyte
loss is driving axonal degeneration after SCI. Whether or not demye-
Axons project long distances from the neuronal cell body, which pre-
lination persists for long enough after SCI or whether the extent of
sents logistical challenges for energy production. Although myelin demyelination crosses this threshold to induce axon loss is unclear. If
reduces the energy requirements on the axon for repeated action demyelination persists in more chronic settings, mitochondrial DNA
potentials, the overall metabolic requirements necessary to establish damage and free radical production creates a vicious cycle increasing
and maintain myelinated axons is higher than developmentally unmy- the likelihood of axonal death due to a metabolic deficit (Campbell,
elinated axons when factoring in myelin genesis and maintenance of Worrall, & Mahad, 2014; Criste, Trapp, & Dutta, 2014). Through these
oligodendrocyte resting potential (Harris & Attwell, 2012). Demyelin- mechanisms, remyelination could prevent long-term damage to axons
ation results in the redistribution and upregulation of ion channels following SCI.
along the axon (England, Gamboni, & Levinson, 1991; England,
Gamboni, Levinson, & Finger, 1990). More specifically, Nav1.2, Nav
1.3 | OPCs as multipotent progenitors after SCI
1.6, Na/Ca2+ exchanger, Kv1.1, and Kv1.2 expression is upregulated
and redistributed on demyelinated axons (Craner et al., 2004; Rasband In adulthood, OPCs remain, as their name would indicate, as dedicated
et al., 1998). Although normally confined to nodes of Ranvier and progenitors of oligodendrocytes (Kang, Fukaya, Yang, Rothstein, &
juxtaparanode, these ion channels spread along the axon length fol- Bergles, 2010; Young et al., 2013). However, after injury or a demye-
lowing demyelination and increase the energy expenditure necessary linating lesion, OPCs are activated and display extensive lineage plas-
to repolarize the axolemma. Demyelination therefore shifts the energy ticity. Traditionally, OPCs have been identified by their expression of
4 DUNCAN ET AL.

(a) Healthy Adult CNS (b) Spinal Cord Injury

OPC/NG2+
OPC/NG2+ Myelin debris Reactive astrocyte

Axon Myelinating cells Scar formation

Oligodendrocyte New oligodendrocyte New myelinating


and associated Schwann cells Reactive
and associated
myelin sheaths astrocyte
myelin sheaths

(c) Uninjured (d) Acute SCI (e) Chronic SCI


(days-weeks post-injury) (months p
post-injury)

Rostral
Ros
Rostra
rall Rostral

Caudal Caudal
Rostral
Rostral

Rostral

Developmental Developmental
Developmental oligodendrocyte oligodendrocyte
oligodendrocyte myelin Ne myelin
w
myelin oli
g
epicentre
epicentre

epicentre

Injury

od r
Injury

Injury

Demyelinated Schwann
end

Axons cell myelin


oc
yte
mye
lin
Caudal
Caudal

Caudal

Myelin content on intact axons (>1mm) Myelin content on intact axons (>1mm) Myelin content on intact axons (>1mm)

F I G U R E 1 Following traumatic injury to the spinal cord, OPCs are multipotent and differentiate into cells capable of remyelination and contributing to
the glial scar. (a) Schematic indicating that OPCs are proliferative in the healthy adult CNS and are restricted to producing new oligodendrocytes.
(b) Following SCI, OPCs become multipotent differentiating into astrocytes, Schwann cells and oligodendrocytes. Both the proliferative rate and the number
of new oligodendrocytes increases after SCI. (c) Diagram of the uninjured spinal cord with oligodendrocyte lineage cells indicated within the spinal cord.
Axons represent descending supraspinal myelinated axons. Graphs below the spinal cord reveal the relative amount of myelin on intact axons that are
greater than 1 μm in diameter, which are typically myelinated, as well as its cellular source. (d) Acute injury (defined as days to several weeks post-SCI) results
in the transection of some axons, which form endbulbs with the distal segment beginning to undergo Wallerian-like degeneration. During the acute
timeframe, OPCs become activated and begin proliferating in response to injury but have not yet began to generate large numbers of new oligodendrocytes
or Schwann cells. (e) Chronically after SCI (defined as months following SCI), OPCs have formed new remyelinating oligodendrocytes along with Schwann
cells which effectively regenerate myelin along intact denuded axons. OPCs also contribute to glial scarring by forming some astrocytes
DUNCAN ET AL. 5

NG2 and PDGFRα (Nishiyama, Lin, Giese, Heldin, & Stallcup, 1996; labeling showed elevated numbers throughout the spinal cord in the
Pringle, Mudhar, Collarini, & Richardson, 1992; Raff, Lillien, Richard- first 4 weeks postinjury (McTigue, Wei, & Stokes, 2001). Similar stud-
son, Burne, & Noble, 1988; Richardson, Pringle, Mosley, ies also revealed BrdU co-labeling with CC1 (a marker of mature oligo-
Westermark, & Dubois-Dalcq, 1988), although these markers can be dendrocytes) and glial fibrillary acidic protein (GFAP, a marker of
expressed on other cell types, like perineurial fibroblasts and pericytes astrocytes) suggesting these cells constitute the majority of newly
(Assinck, Duncan, Plemel, et al., 2017; Kang et al., 2010). Assessment formed cells after injury (Zai & Wrathall, 2005). Indeed, within the first
of proliferation of NG2+ cells using bromodeoxyuridine (BrdU) 24 hr after injury, retrovirally labeled NG2+ progenitors contribute to
scar forming astrocytes, while those born after 7 days postinjury seem
to be mostly contributing to myelin-producing oligodendrocytes
(a) (Sellers, Maris, & Horner, 2009). More recently, fate mapping of
PDGFRα+ cells confirmed that most newly formed oligodendrocytes
after SCI are derived from OPCs (Assinck, Duncan, Plemel, et al.,
Uninjured
MCT1
2017; Hesp et al., 2015). Other groups have also found that 25% of
MCT1
ATP ATP P MCT2
CT2
ATP
AT
ATP ATP MCT2

Kvv
Nav
Kv
P
PP P
PP
PP P PP
PP PP
PP P
PP
PP P PP
PP
Nav
ATP NG2+ fate mapped cells form astrocytes by 4 weeks post-SCI com-
Kvv Kvv Kvv
Kvv PP PP PP PP PP PP Kvv Kv
Nav
PP P PP P PP PP P PP P PP Nav
PP
PP P
PP
PP P PP
PP pared to 9% seen in experimental autoimmune encephalomyelitis
duction (EAE; Hackett et al., 2016, 2018). OPC are less likely to become astro-
Con
cytes in other traumatic models of injury like dorsal column transec-
tions (Barnabe-Heider et al., 2010), and cortical stab injuries (Dimou,
(b) SCI Simon, Kirchhoff, Takebayashi, & Gotz, 2008). Together, these studies
indicate that the injury environment of contusive SCI may be more
conducive to diversity in OPC fate.
Demyelination
The myelin constituency of the spinal cord changes following injury.
OPCs have been shown to produce myelinating Schwann cells after
MCT1
ATP MCT2 chemical demyelination (Zawadzka et al., 2010) as well as following trau-
Kv
ATP
Kv Kv Kv Kv
Nav Kvv
Nav Nav Nav
ATP
Nav Kvv
ATP matic injuries (Assinck, Duncan, Plemel, et al., 2017; Bartus et al., 2019). A
duction population of protein zero (P0) positive Schwann cells derived from OPCs
Con
Myeli epair

appear to dominate the dorsal regions of the injured cord after dorsal con-
Sustained Myelin

tusion injuries (Assinck, Duncan, Plemel, et al., 2017). While, the majority
(c)
nR

of these Schwann cells are derived from PDGFRα-positive OPCs and are
myelinating axons in the spinal cord, both FOXJ1-expressing

Remyelination
MCT1 MCT1
1
MCT2 MCT1
MCT2
CT2
2 MCT2
ATP PP
PP P
PP
PP P
P
P PP
PP P
PP
PP P
PP

F I G U R E 2 Remyelination restores rapid conductance and


PP PP
ATP
AT
TP PP PP PP
Loss

Nav Nav
Kvv Kv Kvv Kv PP P PP P PP
Kvv Kvv Kvv Kvv
Nav PP PP PP Nav
ATP PP P PP P PP
ATP oligodendrocyte-derived trophic support to the axon. (a) Schematic of
an axon myelinated by oligodendrocytes. Voltage-gated sodium
nduction
Co
channels (Nav) and voltage-gated potassium (Kv) channels are located
at the node of Ranvier and juxtaparanode, respectively. The action
(d) potential is reinitiated at the node of Ranvier allowing for saltatory
conductance. Oligodendrocytes provide trophic support to axons by
shuttling lactate and/or pyruvate through MCT1 to the axon, which
Chronic Demyelination expresses MCT2. Myelinated axons are capable of plentiful ATP
production via their mitochondria, and use lower amounts of ATP to
ATP
propagate action potentials relative to unmyelinated axons.
Nav Kv Nav Kv Nav Kv Nav Kv Nav (b) Following SCI, demyelination occurs causing sodium and
potassium channels to spread along the denuded axolemma to
Condu
ction maintain conductance. Trophic support from oligodendrocytes is lost
and the axon requires more energy to propagate action potentials.
(c) Remyelination, which has characteristically short internodes and
thin myelin, restores rapid conduction to the axon. Nav and Kv
Nav voltage-gated sodium channel relocate back to the node and juxtaparanode, respectively. Trophic
Kv voltage-gated potassium channel support is restored to the axon, but mitochondrial number is
mitochondrion increased over healthy axons. (d) Chronic demyelination causes
swelling of the axon and the accumulation of nonphosphorylated
PP
PP P
PP
PP P PP
PP
phosphorylated neurofilaments neurofilaments. Mitochondria drastically increase in size and
unphosphorylated neurofilaments number to meet higher energetic demands for action potential
propagation
6 DUNCAN ET AL.

nonmyelinating and P0+ myelinating Schwann cells from the peripheral axons, which may restore conductance to intact circuitry distal to the
nervous system (PNS) also form remyelinating Schwann cells within the lesion, is likely of much more functional relevance. To determine if intact
CNS (Assinck, Duncan, Plemel, et al., 2017; Ma et al., 2018; Zawadzka axons from descending locomotor pathways are demyelinated, tracers
et al., 2010). Interestingly, Schwann cell myelination appears in the spinal were injected into the red nuclei and motor cortex to trace rubrospinal
cord as early as 2 weeks postinjury when oligodendrocyte myelin has and corticospinal axons along their length (Powers et al., 2012). The pres-
only just begun, suggesting a potentially faster response (Duncan, Man- ence of punctate Caspr staining, found exclusively at paranodes (Peles
esh, et al., 2018). Although the mechanism guiding a fate-switch in OPCs et al., 1997), was used as a surrogate measure of intact myelin (Arancibia-
to differentiate into Schwann cells is still not fully understood, Schwann Carcamo et al., 2017; Etxeberria et al., 2016; Pedraza, Huang, & Colman,
cells are almost exclusively found where there are fewer astrocytic pro- 2009). Shorter internodes, characteristic of adult-derived myelin and
cesses. Accordingly, astrocyte-specific deletion of Stat3 decreases remyelination (Young et al., 2013), were found on 53% of rubrospinal
astrogliosis around the lesion and promotes Schwann cell myelination axons in close proximity (±1 mm) to the lesion epicenter suggesting
after chemical demyelination (Monteiro de Castro, Deja, Ma, Zhao, & extensive remyelination of these tracts (Powers et al., 2012). No intact
Franklin, 2015). This BMP/Wnt rich environment devoid of astrocytes is axons showed evidence of Caspr or potassium channel spreading, typical
permissive for OPCs to adopt a Schwann cell fate, while at the same time of demyelination, and contrasted with transected axons (Powers et al.,
inhibiting them from forming oligodendrocytes (Ulanska-Poutanen et al., 2012). Together, these data indicate that spared axons are not chronically
2018). Schwann cell differentiation from OPCs is therefore dependent on demyelinated and implies they are instead effectively remyelinated.
the injury environment with contusive injuries being very conducive to Further evidence for efficient remyelination after SCI in rodents
damaging astrocytes and producing an environment capable of comes from genetic fate mapping approaches, which specifically label
supporting considerable Schwann cell myelination. new myelin. Inducing the expression of a membrane-tethered fluores-
cent reporter in OPCs, prior to demyelination, can be used to label de
novo myelination when these OPCs differentiate into oligodendro-
1.4 | Remyelination after SCI cytes (Kang et al., 2010). This technique can then be used to distin-
1.4.1 | Endogenous remyelination is an efficient guish new myelin from surviving myelin following SCI (Assinck,
regenerative process after experimental SCI Duncan, Plemel, et al., 2017; Duncan, Manesh, et al., 2018; Hesp
et al., 2015; Powers et al., 2013). Doing so reveals persistent and sub-
Primary demyelination of axons is seen soon after SCI, but whether stantial remyelination by OPC-derived oligodendrocytes and Schwann
demyelination persists chronically in experimental models has been the cells beginning 3 weeks following SCI and culminating with upward of
focus of a series of studies, often with contradictory findings depending 28% of the axons at the lesion epicenter being wrapped by new mye-
on the technique used to measure demyelination (Lasiene, Shupe, Per- lin at 12 weeks post-SCI (Assinck, Duncan, Hilton, et al., 2017). These
lmutter, & Horner, 2008; Totoiu & Keirstead, 2005). Chronic demyelin- approaches validate the profound amount of remyelination observed
ation after SCI has been reported in cats and rodents (Blight, 1983b, by measuring internodal length on spared fibers (Powers et al., 2012).
1985; Totoiu & Keirstead, 2005) and in some human autopsy samples, While considerable progress has been made in understanding the
particularly those with residual spinal cord compression (Bunge et al., levels of de- and remyelination in experimental SCI, less is known about
1993; Guest et al., 2005). However, these observations of demyelinated what occurs clinically. Humans often have poor remyelination efficiency
axons are taken from cross sections at the lesion epicenter and do not dis- in diseases like multiple sclerosis (MS; Goldschmidt, Antel, Konig, Bruck, &
tinguish whether these axons had been transected above or below the Kuhlmann, 2009; Patrikios et al., 2006; Yeung et al., 2019), and thus it is
epicenter or whether these axons are spared. Dystrophic tips from trans- plausible that remyelination efficiency might be decreased after SCI rela-
ected axons persist near the lesion (Guest et al., 2005; Kadoya et al., tive to rodents. However, a limited number of pathological studies dem-
2009; Li & Raisman, 1995; Powers et al., 2012), and can spread potassium onstrate there are relatively few demyelinated axons in most cases
channels along their axolemma, consistent with chronic demyelination (Guest et al., 2005; Kakulas, 2004; Kakulas & Kaelan, 2015; Norenberg
(Karimi-Abdolrezaee, Eftekharpour, & Fehlings, 2004; Lasiene et al., et al., 2004). The absence of extensive demyelination chronically could be
2008; Powers et al., 2012). Transected corticospinal fibers undergo retro- the result of demyelinated axons being vulnerable to degeneration (Buss
grade axonal degeneration (dieback) from the lesion (Pallini, Fernandez, & et al., 2004; Nave & Werner, 2014), or as a consequence of efficient
Sbriccoli, 1988; Seif, Nomura, & Tator, 2007), where they stabilize several remyelination. Immunohistochemical staining of neurofilaments for axons
millimeters away from the lesion in rodent models (Pallini et al., 1988; Seif and Schwann- and oligodendrocyte-specific myelin markers indicate
et al., 2007; Stirling et al., 2004). These axons are likely unable to contrib- remyelination occurs, but that the majority of cases have some level of
ute to functional recovery (Guizar-Sahagun et al., 2004; Kadoya et al., acute demyelination within the first year after SCI (Guest et al., 2005;
2009; Silver, Schwab, & Popovich, 2014). Hypothetically, Kakulas, 2004; Kakulas & Kaelan, 2015), and is more persistent when the
oligodendrocyte-support of transected axons might alleviate dieback and spinal cord is chronically compressed (Bunge et al., 1993). Remyelination
improve the chance that axons can undergo regenerative growth, but is effective in a spontaneous canine model of SCI (Jeffery et al., 2006),
myelination has not been explicitly demonstrated to reduce dieback, nor and in nonhuman primates as long as astrocytes are preserved
if such preservation makes transected axons more prone to reforming (Lachapelle et al., 2005). The presence of demyelinated axons for over a
functional connections. Instead, the myelin status of spared descending year after human SCI is in accordance with the relatively slow
DUNCAN ET AL. 7

OPC-axon interactions Cytokine production


by activated OPC

Blocked conductance
in intact axon
IL-1β
Distrophic Ccl2
endbulb
Csf2
Cxcl1
Cxcl2
IL-6

OPC-derived astrocytes Inflammatory cells enter


contribute to the glial scar the paranchyma

Endo
Tight junction
alteration

thelia
l ce
Basal lamina

lls
alteration

Macrophage MMP-9

F I G U R E 3 Hypothetical functions of oligodendrocyte progenitor cells in the injured CNS beyond remyelination. Schematic of the injured
spinal cord illustrating functions of OPCs beyond generating new oligodendrocytes in response to tissue damage and demyelination. Inlays
highlight the role of OPCs in constraining axon growth, producing astrocytes that contribute to the glial scar, cytokine production, and altering
blood–brain-barrier permeability

proliferative and differentiation rate of human OPCs in both culture and circuitry above and below the site of injury. More specifically, spared
in vivo relative to rodent OPCs (Ehrlich et al., 2017; Wang et al., 2013; descending input from the brain stem can mediate locomotor recov-
Windrem et al., 2004). Taken together, there is little evidence from human ery or, alternatively, “relay” circuits can form between brain stem neu-
samples that the relative extent of chronic demyelination differs dramati- rons and propriospinal neurons to generate the supraspinal input
cally from rodent models and remyelination is observed, albeit the kinet- necessary to induce locomotion (Courtine et al., 2008; Hilton &
ics are likely drastically delayed and could leave axons at risk for Tetzlaff, 2018; Raineteau, Fouad, Bareyre, & Schwab, 2002).
degeneration. That being said, our current knowledge of the extent of Myelin regeneration is a plausible mechanism to contribute to
demyelination is based on a very small number of studies and there is still these forms of plasticity. Descending axons that are spared by the
a great need to understand the frequency of primary demyelination and injury may be demyelinated and thus require new myelin for effective
effectiveness of remyelination after traumatic CNS injury in humans. conductance or to prevent their degeneration. Similarly, the function-
ality of new circuitry, including between propriospinal neurons and
intraspinal networks mediating locomotion, may require myelination.
1.4.2 | Spontaneous oligodendrocyte remyelination
Given the extent of remyelination of descending motor axons
does not drive locomotor recovery after SCI
after contusive SCI in rodents, we hypothesized that it is a critical
Despite the limited ability of the adult mammalian CNS to regenerate driver of locomotor recovery. We, therefore, blocked OPC differentia-
new connections, spontaneous recovery of motor function can occur tion and subsequent remyelination to determine if oligodendrocyte
after human SCI as well as in rodents (Curt, Van Hedel, Klaus, Dietz, & remyelination is necessary for the recovery of locomotion (Duncan,
Group, 2008; Hilton et al., 2016; Torres-Espin, Beaudry, Fenrich, & Manesh, et al., 2018). The transcription factor Myrf, essential for oligo-
Fouad, 2018). In rodent models of SCI, the re-establishment of dendrocyte myelin gene expression (Bujalka et al., 2013; Emery et al.,
stepping after thoracic injury has provided a model for understanding 2009), was deleted from OPCs prior to SCI (Duncan, Manesh, et al.,
the anatomical and mechanistic basis of spontaneous recovery of 2018). Removing Myrf in OPCs does not alter undamaged myelin or
motor function (Basso, Beattie, & Bresnahan, 1995; Courtine et al., OPC recruitment to the lesion (Duncan et al., 2017). Following contu-
2008; Filli, Zorner, Weinmann, & Schwab, 2011; Murray et al., 2010; sion SCI, Myrf knockout from OPCs results in a near complete block
Rossignol & Frigon, 2011; Schucht, Raineteau, Schwab, & Fouad, of oligodendrocyte remyelination, culminating in chronic demyelin-
2002). Locomotor recovery is induced, in part, by plasticity in spared ation and 44% fewer axons myelinated at the lesion epicenter
8 DUNCAN ET AL.

(Duncan, Manesh, et al., 2018). Remarkably, spontaneous locomotor test, but has no effect on gross open field locomotion (Basso mouse
recovery proceeded unencumbered when assessed in an open field score), like previously reported with the Nrg1 deletion (Bartus et al., 2016;
locomotor test (BMS), by gait analysis (Catwalk) or by horizontal lad- Bartus et al., 2019). However, glial scarring, axon growth, and inflamma-
der crossings. tion were not examined, all of which are impacted by Nrg1-ErbB signal-
Clearly potential compensatory mechanisms must be further ing. It also remains unclear why inhibiting the remyelination of a relatively
explored, although we found that preventing oligodendrocyte remyeli- low number of sensory axons in the dorsal column by Schwann cells
nation did not spur additional Schwann cell myelination (Duncan, Manesh, would impair motor recovery after contusive SCI when a much larger
et al., 2018), which seemed confined to injured areas with sparse astrocyte number of demyelinated axons in the ventrolateral columns after oligo-
coverage in the dorsal column of the spinal cord. It is plausible that the rel- dendrocyte ablation does not (Duncan, Manesh, et al., 2018). Neverthe-
ative extent of demyelination along axons in the rodent contusion model, less, these studies point to a role of NRG-1-ErbB3/4 in OPC-derived
which is primarily within 2 mm of the lesion (Powers et al., 2012), is not Schwann cell differentiation and locomotor recovery, though given the
sufficient to block conduction. Indeed, axons can conduct through seg- multifaceted nature of Neuregulin signaling and Schwann cells in repair,
ments of demyelination of exceeding 2.5 mm in vivo (Felts et al., 1997) the precise mechanism(s) by which it confers functional benefits remains
presumably by upregulating and spreading voltage-gated sodium channels unclear.
along the axolemma (England et al., 1990, 1991). Additionally, there is a
temporal discordance between oligodendrocyte remyelination and func-
1.4.4 | Does the acceleration of remyelination
tional recovery. The vast majority of spontaneous locomotor recovery in
constitute a therapeutic target for SCI?
mice has already occurred within the first two weeks after injury, prior to
nearly all oligodendrocyte remyelination (Duncan, Manesh, et al., 2018). Endogenous OPCs differentiate into oligodendrocytes that regenerate
Overall, we found that OPCs are a critical source of remyelinating oligo- myelin around the majority of spared demyelinated axons following
dendrocytes, and that preventing OPC differentiation results in sustained experimental SCI, but remyelination is notably delayed relative to toxin
demyelination. To our surprise, we found that endogenous oligodendro- models of demyelination. Remyelination commences within a week fol-
cyte remyelination of spared axons is not a key contributor to spontane- lowing lysolecithin demyelination (Jeffery & Blakemore, 1995), whereas
ous locomotor recovery following contusive SCI. oligodendrocyte remyelination is scant until at least the third week after
SCI (Assinck, Duncan, Plemel, et al., 2017; Duncan, Manesh, et al., 2018;
James et al., 2011). The speed of remyelination following SCI is likely
1.4.3 | Is Schwann cell myelination linked to
affected by injury-induced changes in the microenvironment, which have
functional recovery?
been recently reviewed elsewhere (Alizadeh & Karimi-Abdolrezaee,
Schwann cell myelination is observed within the spinal cord following 2016; Plemel et al., 2014; Pu, Stephenson, & Yong, 2018; Pukos, Goodus,
experimental contusive SCI, can restore conductance to CNS axons Sahinkaya, & McTigue, 2019). Notably, remyelination may be slowed by
(Felts & Smith, 1987; Honmou, Felts, Waxman, & Kocsis, 1996) and could cytotoxicity near the lesion, insufficient tissue oxygenation (Tsai et al.,
be a mediator of recovery. The deletion of Neuregulin-1 (Nrg-1), which is 2016; Yuen et al., 2014), or inhibitory factors in the glial scar and myelin
instructive for Schwann cell myelination (Michailov et al., 2004), could be debris (Buss & Schwab, 2003; Church, Milich, Lerch, Popovich, &
used to determine its role in locomotor recovery after SCI. Indeed, Bartus McTigue, 2017; Dyck et al., 2018; Dyck, Kataria, Akbari-Kelachayeh, Sil-
et al. (2016) found the inducible deletion of Nrg-1 is associated with ver, & Karimi-Abdolrezaee, 2019; Keough et al., 2016; Plemel, Manesh,
impaired functional locomotor recovery after SCI. One limitation of this Sparling, & Tetzlaff, 2013). Accelerated remyelination is sufficient to pre-
approach is that the tamoxifen inducible Cre mice (CAG Cre-ERT2) used serve axons following T-cell mediated demyelination (Mei et al., 2016).
have expression in virtually every cell type (Hayashi & McMahon, 2002), Rapid remyelination may restore metabolic support of the axon to the oli-
and therefore loss of NRG-1 may affect locomotor recovery by mecha- godendrocyte (Philips & Rothstein, 2017; Saab & Nave, 2017), protect
nisms other than by inhibiting Schwann cell myelination. NRG-1 signaling against inflammatory mediators like nitric oxide (Redford et al., 1997), and
promotes oligodendrocyte differentiation (Gauthier, Kosciuczyk, calcium accumulation within the axon (Witte et al., 2019). It is therefore
Tapley, & Karimi-Abdolrezaee, 2013; Kataria et al., 2018), regulates possible that demyelinated axons after SCI may be more vulnerable to
astrogliosis (Alizadeh et al., 2017; Gauthier et al., 2013) and alters inflam- loss given the relatively sluggish remyelination. However, following SCI
mation by reducing the expression of key inflammatory cytokines after primary demyelination is confined to the proximity of the lesion epicenter
SCI (Alizadeh et al., 2017; Alizadeh, Santhosh, Kataria, Gounni, & Karimi- (Powers et al., 2012), and no study has directly ascertained if the loss of
Abdolrezaee, 2018; Gauthier et al., 2013). Additionally, the deletion of only a small number of myelin sheaths along the length of an axon leaves
just the NRG-1 Type I and II isotypes—which are dispensable for Schwann an axon sufficiently metabolically compromised to be vulnerable to sub-
cell myelination—is sufficient to inhibit open field locomotion (Bartus sequent loss. Still, if hastened remyelination is successful at preserving
et al., 2016). To address these concerns, Bartus et al. (2019) knocked out axons it may be a viable therapeutic target for SCI (Plemel et al., 2014).
the ErbB3/4 receptors from PDGFRα+ cells following SCI and similarly Given the longer timeframe necessary for OPC differentiation in humans
observed impaired Schwann cell myelination from endogenous OPCs, relative to rodents (Goldman & Kuypers, 2015; Windrem et al., 2004),
with no reported effect on oligodendrocyte remyelination. Functionally, accelerating remyelination may be helpful clinically as axons may be
the deletion of ErbB3/4 results in deficits on an inclined beam walking stressed longer in humans than in experimental rodent models.
DUNCAN ET AL. 9

Two strategies have been used to speed remyelination: (a) the may be causal in locomotor improvements. It will be critical to determine
transplantation of exogenous cells to directly remyelinate, or less fre- which cell types elicit recovery, and the mechanisms by which these cells
quently, (b) enhancing endogenous precursor-derived remyelination. drive recovery in order to optimize them for use in the clinic.
The following section will examine the efficacy of these strategies to
promote remyelination, the evidence of whether this increases axo-
1.5 | Manipulation of endogenous oligodendrocyte
n/tissue sparing and its relationship to improvements in functional
lineage cells to accelerate remyelination
recovery.
OPCs are the most proliferative cell population and the primary source of
new oligodendrocytes after SCI (Barnabe-Heider et al., 2010; Duncan,
1.4.5 | Cellular transplantation to accelerate
Manesh, et al., 2018). These attributes make OPCs an ideal cell population
remyelination
to target to enhance their differentiation into new oligodendrocytes to
The transplantation of cells into the injured spinal cord with the aim of accelerate remyelination. Approaches that have improved endogenous
producing new oligodendrocytes or Schwann cells has been frequently OPC differentiation after SCI include attenuating inhibitors of OPC differ-
pursued to improve remyelination, and is the topic of several thorough entiation like BMP signaling (Sellers et al., 2009; Wang et al., 2011) or
reviews (Assinck, Duncan, Hilton, et al., 2017; Kanno, Pearse, Ozawa, myelin debris (Church et al., 2017; Plemel et al., 2013). Alternatively,
Itoi, & Bunge, 2015; Myers et al., 2016). Transplanted cells will be in com- expressing known promoters of oligodendrocyte differentiation like T3
petition with endogenous OPC-derived remyelinating cells, which are (Shultz, Wang, Nong, Zhang, & Zhong, 2017), or growth factors like NT-3,
capable of efficient, if somewhat delayed, remyelination. Transplanted BDNF (McTigue, Horner, Stokes, & Gage, 1998) have also been success-
cells, therefore, only have a limited time window of efficacy that is ful in promoting OPC differentiation into new oligodendrocytes. None of
supported by the preponderance of studies: myelinating cell transplants these experiments examined axon sparing or functional recovery in the
can improve remyelination, white matter content, and function when context of more rapid oligodendrocyte differentiation, but a series of
transplanted within the first several weeks after injury (Biernaskie et al., experiments increasing Nrg1 signaling have done so (Alizadeh et al.,
2007; Cao et al., 2010; Karimi-Abdolrezaee, Eftekharpour, Wang, 2017; Gauthier et al., 2013; Whittaker et al., 2012). Administration of
Morshead, & Fehlings, 2006; Keirstead et al., 2005; Pearse et al., 2004). recombinant Nrg1-β1, promotes remyelination in a chemical model of
While more chronic transplantations of remyelinating cells have in some demyelination (Kataria et al., 2018), increases white matter sparing, and
cases shown to be beneficial in increasing locomotor recovery (Barakat improves open field locomotion following SCI (Alizadeh et al., 2017;
et al., 2005; Karimi-Abdolrezaee, Eftekharpour, Wang, Schut, & Fehlings, Gauthier et al., 2013). However, Nrg1-β1 also alters the inflammatory
2010; Nori et al., 2018; Okubo et al., 2018), none of these chronic studies milieu and reduces glial scaring (Alizadeh et al., 2018), making it difficult to
demonstrate that the extent of remyelination is improved using gold- ascertain if remyelination is driving these functional benefits. Application
standard measurements of electron microscopy or fluorescently-labelled of Nrg1-β3 (GGF2) following SCI increases oligodendrocyte differentia-
new myelin. Therefore, myelinating cell transplants likely need to be tion and the quality of stepping in rats and mice (Whittaker et al., 2012).
transplanted early after SCI to outcompete endogenous progenitors and In rats, Nrg1-β3 administration does not alter spared white matter at
accelerate remyelination. 7 days postinjury but increases it by 42 days post-SCI (Whittaker et al.,
It is also possible that benefits following myelinating cell transplanta- 2012). This delay in white matter sparing suggests that Nrg1-β3 does not
tion are conferred by mechanisms other than remyelination (Assinck, drastically alter early secondary damage to myelinated axons, but may be
Duncan, Hilton, et al., 2017). Transplanted neural progenitor cells (NPCs) reflective of improved regenerative processes, such as remyelination, by
and glial progenitors differentiate into many cell types following SCI, all of 42 days postinjury. Collectively, these studies are a proof of principle that
which may influence repair (Tetzlaff et al., 2011). To determine if the ben- speeding endogenous oligodendrocyte differentiation is possible after
eficial effect is incurred by more rapid remyelination, transplantation of SCI and has in some cases been associated with improved white matter
NPCs from shiverer mice, which lack compact myelin (Dupouey et al., volume and functional benefits. However, further cell-specific manipula-
1979; Kirschner & Ganser, 1980), were compared to transplanted myelin- tions will be needed to determine if locomotor recovery is being derived
competent NPCs (Hawryluk et al., 2014; Yasuda et al., 2011). They found by remyelination or other factors.
that locomotor improvements were abolished when transplanted cells
were rendered unable to myelinate (Hawryluk et al., 2014; Yasuda et al.,
1.6 | Beyond Remyelination: Alternative functions of
2011) supporting a role of myelinogenesis by transplanted cells in driving
oligodendrocyte lineage cells after SCI
improved recovery. Likewise, transplantations with enhanced oligoden-
drocyte differentiation over naïve NPCs have been associated with Local tissue damage, including that caused by mechanical injury, results in
improved myelination and locomotor function, indicating forming oligo- the activation of glial cells, which produce a barrier or scar to segregate
dendrocyte lineage cells may be particularly advantageous (Cao et al., inflammatory cells to areas of tissue damage and protect residual tissue
2010; Hofstetter et al., 2005; Hwang et al., 2009; Piltti et al., 2017). (Burda & Sofroniew, 2014; Faulkner et al., 2004; Silver et al., 2014;
Whether transplanted oligodendrocytes enhance spared white matter via Sofroniew, 2015). The scar contains numerous cell types, prominently
direct trophic support of axons remains uncertain. Transplanted cells may astrocytes fibroblasts, but also OPCs (Busch et al., 2010; Cregg et al.,
also spur endogenous cell remyelination (Biernaskie et al., 2007), which 2014; Silver et al., 2014). Recent evidence indicates that OPCs local to
10 DUNCAN ET AL.

the scar may be major regulators of tissue repair, extending their func- preventing DRG regeneration following crush. Collectively, NG2+ cells
tions beyond generating new oligodendrocytes for remyelination are associated with entrapped growth cones from both damaged CNS or
(Fernandez-Castaneda & Gaultier, 2016; Hamanaka, Ohtomo, Takase, PNS axons growing into the CNS parenchyma, and may be serving to
Lok, & Arai, 2018; Silver et al., 2014). The following sections will discuss entrap and diminish axon growth.
potential functions of OPCs in CNS injury focusing on their roles in regu- NG2 is often used to identify OPCs in uninjured tissue (Nishiyama
lating axonal growth and the immune response to injury. et al., 1996) and is also strongly expressed by OPCs after SCI (McTigue
et al., 2001). It should be cautioned that other cells including blood-
derived macrophages express NG2 near the lesion (McTigue, Tripathi, &
1.6.1 | OPCs contribute to scarring and may regulate
Wei, 2006), so it remains unclear whether these entrapping cells in vivo
axonal growth
are in fact oligodendrocyte lineage cells. The NG2+ cells associated with
OPCs, often identified by the expression of NG2 (Nishiyama et al., 1996), entrapped axons express both vimentin and nestin (Filous et al., 2014).
proliferate near the lesion in response to SCI (Horky, Galimi, Gage, & These are progenitor markers not typically associated with adult OPCs
Horner, 2006; McTigue et al., 2001; Tripathi & McTigue, 2007; Zai & (Barnabe-Heider et al., 2010; Zhang et al., 2014), and are only expressed
Wrathall, 2005) and contribute to the scar (Filous et al., 2014) (Figure 3). transiently in OPCs following dissociation for cell culture (Behar,
This leaves OPCs in the ideal location to regulate axon growth and die- McMorris, Novotny, Barker, & Dubois-Dalcq, 1988; Tang, Tokumoto, &
back, potentially through the expression of extracellular proteins. Extra- Raff, 2000).Whether activated OPCs have adapted a more progenitor-
cellular molecules like chondroitin sulfate proteoglycans (CSPG)s inhibit like phenotype in vivo in response to injury (Moyon et al., 2015) or if these
axon regeneration (Cregg et al., 2014; Siebert, Conta Steencken, & NG2-expressing cells are derived from other cell types remains unclear.
Osterhout, 2014). Dystrophic axons, which persist near the lesion epicen- Future studies will need to specifically determine if axon-entrapping NG2
ter after SCI (Kadoya et al., 2009; Li & Raisman, 1995; Powers et al., + cells are derived from the oligodendrocyte lineage by genetically label-
2012), are often associated with cells expressing the CSPG NG2 (Busch ing OPCs prior to injury.
et al., 2010; Tan, Colletti, Rorai, Skene, & Levine, 2006). Function blocking Scar formation has beneficial aspects: it constrains blood-derived leu-
NG2 antibodies (Tan et al., 2006), as well as the depletion of proliferative kocytes and protects tissue from additional damage (Faulkner et al., 2004;
NG2 cells (Hesp et al., 2018), increase axon growth following SCI, Goritz et al., 2011; Herrmann et al., 2008; Sabelstrom et al., 2013). How-
supporting a role for NG2 in constraining axon regeneration. However, in ever, parenchymal astrocytes are limited in their capacity to migrate and
apparent contrast to a function of NG2+ cells in reducing axon growth, proliferate (Barnabe-Heider et al., 2010; Tsai et al., 2012). In cases of sub-
NG2 gene ablation decreases the number of serotonergic axons within stantial tissue disruption, OPCs could be an additional source of astro-
the lesion (de Castro Jr, Tajrishi, Claros, & Stallcup, 2005). One possibility cytes to further supplement their numbers. There has been some debate
to rectify these findings is that NG2+ cells also prevent axonal dieback by about the capacity of OPCs to differentiate into astrocytes following SCI,
associating with dystrophic axons, which is supported by greater sensory but two groups have recently provided compelling evidence they do so
axon dieback in NG2 knockout mice, and the capacity of OPCs to stabilize using genetic fate mapping approaches (Hackett et al., 2016; Huang et al.,
axons endbulbs in vitro (Busch et al., 2010). These data suggest a model in 2018). Attenuating the OPC response to injury also diminishes
which axon dieback is mitigated by NG2+ cells following SCI, but then astrogliosis, suggesting a correlative relationship between these pro-
NG2+ cells subsequently entrap growth cones and diminish axon growth cesses (Hesp et al., 2018; Rodriguez et al., 2014). For example, inhibiting
(Silver et al., 2014; Son, 2015). NG2+ cell proliferation by using mice expressing thymidine kinase in NG2
Likewise, NG2+ cells may impair regeneration of peripheral dorsal + cells and treating these mice with ganciclovir (GCV), results in discontin-
root ganglion (DRG) axons into the spinal cord. Crushing peripheral roots uous expression of GFAP around the lesion, worsened edema, and
has been traditionally used as a model to understand the intrinsic inhibi- impaired recovery after SCI (Hesp et al., 2018). Likewise, inhibition of
tory environment of the CNS on axon growth, as DRG axons tend to stall OPC proliferation by inducible knockout of β-catenin from OPCs reduces
at the PNS/CNS boundary (Di Maio et al., 2011; Hanna, Son, & Dempsey, GFAP expression around the lesion (Rodriguez et al., 2014). Interestingly,
2011; Ramer, Priestley, & McMahon, 2000; Son, 2015). Reports from the in both approaches (Tk/GCV or β-catenin knockout) increased axon num-
1980's suggested that these immobilized axons can form “axo-glial ber was observed within the scar. Collectively, these data suggest that
terminals,” which resemble presynapses apposed from glial processes, OPCs are critical for the appropriate scar formation, reducing edema, and
which were thought to be astrocytes (Carlstedt, 1985; Liuzzi & Lasek, provide further support that NG2 cells or their progeny restrict axon
1987). However, by using wholemount immunostaining of the DRG and growth. However, it is still unclear if OPCs are acting directly, or simply
spinal cord following a dorsal root crush, it was shown that DRG axons altering astrocyte function. Likewise, current conditional and inducible
are often immobilized on NG2+ processes rather than GFAP+ astrocytes Cre lines (e.g., PDGFRα CreERT and NG2 CreERT) used to manipulate
or myelin (Di Maio et al., 2011; Hanna et al., 2011; Son, 2015). Ultrastruc- gene expression in OPCs share expression of these genes with
tural examination confirmed that DRG axons have vesicle accumulation vasculature-associated cells including pericytes (Assinck, Duncan, Plemel,
resembling an active zone proximal to glial processes, though these glial et al., 2017; Kang et al., 2010). Pericytes contribute to the accumulation
processes were not explicitly identified to be NG2+ by immunoelectron of fibrotic stromal cells within the scar, which inhibits corticospinal axon
microscopy. Genetic knockout or antibody-mediated blocking of NG2 growth (Dias et al., 2018). Caution is therefore necessary when inter-
proteoglycan will be needed to confirm if NG2 cells are causative in preting these studies as an OPC-specific effect. However, the
DUNCAN ET AL. 11

accumulation of OPCs at the lesion, their capacity to differentiate into Hesp et al., 2018). Given the rapid proliferation of OPCs and presence
astrocytes, their prominent expression of NG2, and the functional studies within the scar where they could affect immune infiltration and lesion
where their proliferation is inhibited paint a cumulative picture that these development, it will be crucial to determine to what extent they sculpt
cells are likely crucial for the scarring response and regulate axon growth. the inflammatory response after SCI.

1.6.2 | OPCs rapidly respond to tissue injury and 2 | C O N CL U S I O N


may regulate the subsequent immune response
OPCs are the largest source of endogenous progenitors in the adult spinal
Live-imaging studies have determined that OPCs can elongate pro-
cord. Following traumatic injury, OPCs become multipotent contributing
cesses toward tissue injury within hours and begin proliferating rap-
to the replacement of lost oligodendrocytes and astrocytes (Assinck,
idly thereafter (Hill et al., 2014; Hughes et al., 2013). While not as
Duncan, Plemel, et al., 2017; Hackett et al., 2018). OPCs produce virtually
rapid as local microglial responses to tissue damage (Davalos et al.,
all new remyelinating oligodendrocytes and the majority of myelinating
2005), OPCs respond quickly to damage. Several new studies provide
Schwann cells (Assinck, Duncan, Plemel, et al., 2017; Barnabe-Heider
compelling evidence that this quick response to demyelination helps
et al., 2010), which together are effective in remyelinating nearly all intact
regulate both adaptive and innate immune responses (Figure 3).
demyelinated axons following experimental SCI in rodents (Duncan, Man-
OPCs regulate adaptive immune responses by acting as antigen pre-
esh, et al., 2018; Lasiene et al., 2008; Powers et al., 2012). While
senting cells during chemical and autoimmune-mediated demyelination
remyelination is sufficient to improve conductance along axons (Smith,
(Falcao et al., 2018; Kirby et al., 2018). In vivo gene profiling confirms that
Blakemore, & Mcdonald, 1979) and may help preserve axons over the
oligodendrocyte-lineage cells express the major histocompatibility com-
long-term (Lee et al., 2012), spontaneous oligodendrocyte remyelination
plex Genes I and II in response to cytokines like IFNγ (Falcao et al., 2018;
is not a major driver of the limited locomotor recovery observed early
Kirby et al., 2018). Additionally, OPCs phagocytose antigens in reaction
after contusive SCI (Duncan, Manesh, et al., 2018). However,
to IFNγ stimulation, and subsequently present these antigens to CD4+ T-
helper cells to support their survival and proliferation (Falcao et al., 2018). remyelination is somewhat delayed in the rodent spinal cord relative to

OPCs also present antigens to activate CD8+ cytotoxic T-cells (Kirby chemical demyelination, and remyelinating cell transplants and strategies

et al., 2018). Interestingly, transfer of IFNγ-expressing astrocytes or effec- to improve endogenous remyelination has been shown to facilitate loco-

tor T-cells results in impaired differentiation of OPCs into oligodendro- motor recovery, especially when instituted shortly after SCI, though it

cytes and slowed remyelination (Kirby et al., 2018). These data indicate remains unclear precisely how these interventions are driving recovery

that OPCs are co-opted by the immune system to be involved in antigen (Assinck, Duncan, Hilton, et al., 2017). Given the multifaceted nature of

presentation serving as an intermediate to drive further immune the OPC responses to traumatic injury, it is plausible that OPCs have roles

responses at the cost of their capacity to generate new oligodendrocytes. in repair not related to the generation of new myelinating cells. Accord-
OPCs are thus influenced by their local environment during tissue dam- ingly, the inhibition of OPC proliferation in response to SCI impairs recov-
age and can encourage subsequent adaptive immune responses. ery, without affecting oligodendrocyte density (Hesp et al., 2018). OPCs
OPCs also produce cytokines to regulate the innate immune response contribute to the glial scar and through their expression of the NG2 pro-
following tissue damage (Figure 3). For example, during autoimmune teoglycan may regulate axonal growth and dieback (Filous et al., 2014;
demyelination, the deletion of the key transducer of IL-17 signaling, Act1, Hackett et al., 2018). They may also regulate inflammatory responses
in OPCs and pericytes reduces expression of inflammatory cytokines, leu- after SCI and are capable of antigen presentation and cytokine production
kocyte infiltration, and clinical scores (Kang et al., 2013). OPCs alter their (Falcao et al., 2018; Kang et al., 2013; Moyon et al., 2015). These potential
transcriptome in response to demyelination and prominently express the novel functions of OPCs in tissue remodeling after injury could be critical
inflammatory cytokines IL1β and CCL2 (Moyon et al., 2015), a potent reg- mediators of functional recovery and may provide new therapeutic ave-
ulator of monocyte infiltration (Deshmane, Kremlev, Amini, & Sawaya, nues to promote recovery.
2009). Following ischemia (Seo et al., 2013), or demyelination (Pham
et al., 2012), OPCs may regulate BBB permeability by rapidly expressing
ACKNOWLEDG MENTS
MMP9, which is associated with weakening of the BBB, and subsequent
immune infiltration (Hamanaka et al., 2018; Rempe, Hartz, & Bauer, The authors would like to thank the generous support from the Canadian
2016). Indeed, conditioned media from OPCs reduces the expression of Institute of Health Research and the Multiple Sclerosis Society of Canada
tight junction proteins in endothelial cells, which may permit increased to W.T. We also thank all current and past members of the Tetzlaff labo-
leukocyte infiltration (Seo et al., 2013). In traumatic SCI, inhibiting OPC ratory for their thought-provoking discussion on myelin regeneration and
proliferation via the inducible knockout of β-catenin reduces the number drivers of locomotor recovery following SCI. G.J.D. is supported by a
of microglia/macrophages within the lesion, suggesting a role of OPCs in National Multiple Sclerosis Society fellowship (FG-1808-32238). B.J.H. is
potentiating the innate immune response after SCI (Rodriguez et al., supported by a Wings for Life (WfL) Aguayo-Tator Mentoring Fellowship
2014). As before, given that the knockout would also affect pericytes, it and a nonstipendiary European Molecular Biology Organization (EMBO)
remains unclear whether this phenotype is derived from diminishing the Long-Term Fellowship (ALTF 28-2017). P.A. is supported by an Multiple
OPC or pericyte responses to injury (Duncan, Assinck, & Hilton, 2014; Sclerosis Society of Canada Postdoctoral Fellowship and Marie Curie
12 DUNCAN ET AL.

(MSCA) Individual Fellowship. W.T. holds the John and Penny Ryan Brit- remyelination after spinal cord injury. Glia, 67, 1036–1046. https://
ish Columbia Leadership Chair in Spinal Cord Research. doi.org/10.1002/glia.23586
Bartus, K., Galino, J., James, N. D., Hernandez-Miranda, L. R., Dawes, J. M.,
Fricker, F. R., … Bradbury, E. J. (2016). Neuregulin-1 controls an
CONF LICT OF IN TE RE ST endogenous repair mechanism after spinal cord injury. Brain, 139(Pt 5),
1394–1416. https://doi.org/10.1093/brain/aww039
Authors have no conflict of interest to report. Basso, D. M., Beattie, M. S., & Bresnahan, J. C. (1995). A sensitive and reli-
able locomotor rating scale for open field testing in rats. Journal of
Neurotrauma, 12(1), 1–21. https://doi.org/10.1089/neu.1995.12.1
ORCID Beaumont, E., Onifer, S. M., Reed, W. R., & Magnuson, D. S. (2006). Mag-
netically evoked inter-enlargement response: An assessment of
Brett J. Hilton https://orcid.org/0000-0003-2813-2294 ascending propriospinal fibers following spinal cord injury. Experimen-
Jason R. Plemel https://orcid.org/0000-0003-1385-1464 tal Neurology, 201(2), 428–440. https://doi.org/10.1016/j.expneurol.
2006.04.032
Wolfram Tetzlaff https://orcid.org/0000-0003-3462-1676
Behar, T., McMorris, F. A., Novotny, E. A., Barker, J. L., & Dubois-Dalcq, M.
(1988). Growth and differentiation properties of O-2A progenitors
purified from rat cerebral hemispheres. Journal of Neuroscience
RE FE R ENC E S
Research, 21(2–4), 168–180. https://doi.org/10.1002/jnr.490210209
Bergles, D. E., Jabs, R., & Steinhauser, C. (2010). Neuron-glia synapses in
Alizadeh, A., Dyck, S. M., & Karimi-Abdolrezaee, S. (2015). Myelin damage
the brain. Brain Research Reviews, 63(1–2), 130–137. https://doi.org/
and repair in pathologic CNS: Challenges and prospects. Frontiers in
Molecular Neuroscience, 8, 35. https://doi.org/10.3389/fnmol.2015. 10.1016/j.brainresrev.2009.12.003
Bergles, D. E., Roberts, J. D., Somogyi, P., & Jahr, C. E. (2000). Glutamatergic
00035
Alizadeh, A., Dyck, S. M., Kataria, H., Shahriary, G. M., Nguyen, D. H., synapses on oligodendrocyte precursor cells in the hippocampus. Nature,
Santhosh, K. T., & Karimi-Abdolrezaee, S. (2017). Neuregulin-1 posi- 405(6783), 187–191. https://doi.org/10.1038/35012083
tively modulates glial response and improves neurological recovery Bernardi, P. (1992). Modulation of the mitochondrial cyclosporin A-
following traumatic spinal cord injury. Glia, 65(7), 1152–1175. https:// sensitive permeability transition pore by the proton electrochemical
doi.org/10.1002/glia.23150 gradient. Evidence that the pore can be opened by membrane depolar-
Alizadeh, A., & Karimi-Abdolrezaee, S. (2016). Microenvironmental regula- ization. The Journal of Biological Chemistry, 267(13), 8834–8839.
tion of oligodendrocyte replacement and remyelination in spinal cord Biernaskie, J., Sparling, J. S., Liu, J., Shannon, C. P., Plemel, J. R., Xie, Y., …
injury. The Journal of Physiology, 594(13), 3539–3552. https://doi.org/ Tetzlaff, W. (2007). Skin-derived precursors generate myelinating
10.1113/JP270895 Schwann cells that promote remyelination and functional recovery
Alizadeh, A., Santhosh, K. T., Kataria, H., Gounni, A. S., & Karimi- after contusion spinal cord injury. Journal of Neuroscience, 27(36),
Abdolrezaee, S. (2018). Neuregulin-1 elicits a regulatory immune response 9545–9559. https://doi.org/10.1523/Jneurosci.1930-07.2007
following traumatic spinal cord injury. Journal of Neuroinflammation, 15(1), Blight, A. R. (1983a). Axonal physiology of chronic spinal cord injury in the
53. https://doi.org/10.1186/s12974-018-1093-9 cat: Intracellular recording in vitro. Neuroscience, 10(4), 1471–1486.
Almeida, R. G., & Lyons, D. A. (2014). On the resemblance of synapse for- Blight, A. R. (1983b). Cellular morphology of chronic spinal cord injury in
mation and CNS myelination. Neuroscience, 276, 98–108. https://doi. the cat: Analysis of myelinated axons by line-sampling. Neuroscience,
org/10.1016/j.neuroscience.2013.08.062 10(2), 521–543.
Arancibia-Carcamo, I. L., Ford, M. C., Cossell, L., Ishida, K., Tohyama, K., & Blight, A. R. (1985). Delayed demyelination and macrophage invasion: A
Attwell, D. (2017). Node of Ranvier length as a potential regulator of candidate for secondary cell damage in spinal cord injury. Central Ner-
myelinated axon conduction speed. eLife, 6, 6. https://doi.org/10. vous System Trauma, 2(4), 299–315. https://doi.org/10.1089/cns.
7554/eLife.23329 1985.2.299
Assinck, P., Duncan, G. J., Hilton, B. J., Plemel, J. R., & Tetzlaff, W. (2017). Bujalka, H., Koenning, M., Jackson, S., Perreau, V. M., Pope, B., Hay, C. M.,
Cell transplantation therapy for spinal cord injury. Nature Neuroscience, … Emery, B. (2013). MYRF is a membrane-associated transcription fac-
20(5), 637–647. https://doi.org/10.1038/nn.4541 tor that autoproteolytically cleaves to directly activate myelin genes.
Assinck, P., Duncan, G. J., Plemel, J. R., Lee, M. J., Stratton, J. S., PLoS Biology, 11(8), e1001625. https://doi.org/10.1371/journal.pbio.
Manesh, S. B., … Tetzlaff, W. (2017). Myelinogenic plasticity of oligo- 1001625
dendrocyte precursor cells following spinal cord contusion injury. The Bunge, R. P., Puckett, W. R., Becerra, J. L., Marcillo, A., & Quencer, R. M.
Journal of Neuroscience, 37, 8635–8654. https://doi.org/10.1523/ (1993). Observations on the pathology of human spinal cord injury. A
JNEUROSCI.2409-16.2017 review and classification of 22 new cases with details from a case of
Atkinson, P. P., & Atkinson, J. L. (1996). Spinal shock. Mayo Clinic Proceedings, chronic cord compression with extensive focal demyelination.
71(4), 384–389. https://doi.org/10.1016/S0025-6196(11)64067-6 Advances in Neurology, 59, 75–89.
Barakat, D. J., Gaglani, S. M., Neravetla, S. R., Sanchez, A. R., Burda, J. E., & Sofroniew, M. V. (2014). Reactive gliosis and the mul-
Andrade, C. M., Pressman, Y., … Pearse, D. D. (2005). Survival, integra- ticellular response to CNS damage and disease. Neuron, 81(2),
tion, and axon growth support of glia transplanted into the chronically 229–248. https://doi.org/10.1016/j.neuron.2013.12.034
contused spinal cord. Cell Transplantation, 14(4), 225–240. https://doi. Busch, S. A., Horn, K. P., Cuascut, F. X., Hawthorne, A. L., Bai, L. H.,
org/10.3727/000000005783983106 Miller, R. H., & Silver, J. (2010). Adult NG2+cells are permissive to neu-
Barnabe-Heider, F., Goritz, C., Sabelstrom, H., Takebayashi, H., rite outgrowth and stabilize sensory axons during macrophage-
Pfrieger, F. W., Meletis, K., & Frisen, J. (2010). Origin of new glial cells induced axonal dieback after spinal cord injury. Journal of Neuroscience,
in intact and injured adult spinal cord. Cell Stem Cell, 7(4), 470–482. 30(1), 255–265. https://doi.org/10.1523/Jneurosci.3705-09.2010
https://doi.org/10.1016/j.stem.2010.07.014 Buss, A., Brook, G. A., Kakulas, B., Martin, D., Franzen, R., Schoenen, J., …
Bartus, K., Burnside, E. R., Galino, J., James, N. D., Bennett, D. L. H., & Schmitt, A. B. (2004). Gradual loss of myelin and formation of an astro-
Bradbury, E. J. (2019). ErbB receptor signaling directly controls oligo- cytic scar during Wallerian degeneration in the human spinal cord.
dendrocyte progenitor cell transformation and spontaneous Brain, 127(Pt 1), 34–44. https://doi.org/10.1093/brain/awh001
DUNCAN ET AL. 13

Buss, A., & Schwab, M. E. (2003). Sequential loss of myelin proteins during border. The Journal of Neuroscience, 31(12), 4569–4582. https://doi.
Wallerian degeneration in the rat spinal cord. Glia, 42(4), 424–432. org/10.1523/JNEUROSCI.4638-10.2011
https://doi.org/10.1002/glia.10220 Dias, D. O., Kim, H., Holl, D., Werne Solnestam, B., Lundeberg, J.,
Cai, J., Qi, Y., Hu, X., Tan, M., Liu, Z., Zhang, J., … Qiu, M. (2005). Genera- Carlen, M., … Frisen, J. (2018). Reducing pericyte-derived scarring pro-
tion of oligodendrocyte precursor cells from mouse dorsal spinal cord motes recovery after spinal cord injury. Cell, 173(1), 153–165 e122.
independent of Nkx6 regulation and Shh signaling. Neuron, 45(1), https://doi.org/10.1016/j.cell.2018.02.004
41–53. https://doi.org/10.1016/j.neuron.2004.12.028 Dimou, L., Simon, C., Kirchhoff, F., Takebayashi, H., & Gotz, M. (2008).
Campbell, G. R., Worrall, J. T., & Mahad, D. J. (2014). The central role of mito- Progeny of Olig2-expressing progenitors in the gray and white matter
chondria in axonal degeneration in multiple sclerosis. Multiple Sclerosis, 20 of the adult mouse cerebral cortex. The Journal of Neuroscience, 28(41),
(14), 1806–1813. https://doi.org/10.1177/1352458514544537 10434–10442. https://doi.org/10.1523/JNEUROSCI.2831-08.2008
Cao, Q., He, Q., Wang, Y., Cheng, X., Howard, R. M., Zhang, Y., … Duncan, G. J., Assinck, P., & Hilton, B. J. (2014). Canonical Wnt signalling
Whittemore, S. R. (2010). Transplantation of ciliary neurotrophic in PDGFRalpha-expressing cells is a critical regulator of astrogliosis
factor-expressing adult oligodendrocyte precursor cells promotes and axon regeneration following CNS injury. The Journal of Neurosci-
remyelination and functional recovery after spinal cord injury. The ence, 34(49), 16163–16165. https://doi.org/10.1523/JNEUROSCI.
Journal of Neuroscience, 30(8), 2989–3001. https://doi.org/10.1523/ 4052-14.2014
JNEUROSCI.3174-09.2010 Duncan, G. J., Manesh, S. B., Hilton, B. J., Assinck, P., Liu, J., Moulson, A.,
Carlstedt, T. (1985). Regenerating axons form nerve terminals at astro- … Tetzlaff, W. (2018). Locomotor recovery following contusive spinal
cytes. Brain Research, 347(1), 188–191. cord injury does not require oligodendrocyte remyelination. Nature
Church, J. S., Milich, L. M., Lerch, J. K., Popovich, P. G., & McTigue, D. M. Communications, 9(1), 3066. https://doi.org/10.1038/s41467-018-
(2017). E6020, a synthetic TLR4 agonist, accelerates myelin debris 05473-1
clearance, Schwann cell infiltration, and remyelination in the rat spinal Duncan, G. J., Plemel, J. R., Assinck, P., Manesh, S. B., Muir, F. G. W.,
cord. Glia, 65(6), 883–899. https://doi.org/10.1002/glia.23132 Hirata, R., … Tetzlaff, W. (2017). Myelin regulatory factor drives
Courtine, G., Song, B., Roy, R. R., Zhong, H., Herrmann, J. E., Ao, Y., … remyelination in multiple sclerosis. Acta Neuropathologica, 134,
Sofroniew, M. V. (2008). Recovery of supraspinal control of stepping 403–422. https://doi.org/10.1007/s00401-017-1741-7
via indirect propriospinal relay connections after spinal cord injury. Duncan, I. D., Radcliff, A. B., Heidari, M., Kidd, G., August, B. K., &
Nature Medicine, 14(1), 69–74. https://doi.org/10.1038/nm1682 Wierenga, L. A. (2018). The adult oligodendrocyte can participate in
Craner, M. J., Newcombe, J., Black, J. A., Hartle, C., Cuzner, M. L., & remyelination. Proceedings of the National Academy of Sciences of the
Waxman, S. G. (2004). Molecular changes in neurons in multiple scle- United States of America, 115, E11807–E11816. https://doi.org/10.
rosis: Altered axonal expression of Nav1.2 and Nav1.6 sodium chan- 1073/pnas.1808064115
nels and Na+/Ca2+ exchanger. Proceedings of the National Academy of Dupouey, P., Jacque, C., Bourre, J. M., Cesselin, F., Privat, A., &
Sciences of the United States of America, 101(21), 8168–8173. https:// Baumann, N. (1979). Immunochemical studies of myelin basic protein
doi.org/10.1073/pnas.0402765101 in shiverer mouse devoid of major dense line of myelin. Neuroscience
Cregg, J. M., DePaul, M. A., Filous, A. R., Lang, B. T., Tran, A., & Silver, J. Letters, 12(1), 113–118.
(2014). Functional regeneration beyond the glial scar. Experimental Dutta, D. J., Woo, D. H., Lee, P. R., Pajevic, S., Bukalo, O., Huffman, W. C.,
Neurology, 253, 197–207. https://doi.org/10.1016/j.expneurol.2013. … Fields, R. D. (2018). Regulation of myelin structure and conduction
12.024 velocity by perinodal astrocytes. Proceedings of the National Academy
Criste, G., Trapp, B., & Dutta, R. (2014). Axonal loss in multiple sclerosis: of Sciences of the United States of America, 115(46), 11832–11837.
Causes and mechanisms. Handbook of Clinical Neurology, 122, https://doi.org/10.1073/pnas.1811013115
101–113. https://doi.org/10.1016/B978-0-444-52001-2.00005-4 Dyck, S., Kataria, H., Akbari-Kelachayeh, K., Silver, J., & Karimi-
Crowe, M. J., Bresnahan, J. C., Shuman, S. L., Masters, J. N., & Abdolrezaee, S. (2019). LAR and PTPsigma receptors are negative reg-
Beattie, M. S. (1997). Apoptosis and delayed degeneration after spinal ulators of oligodendrogenesis and oligodendrocyte integrity in spinal
cord injury in rats and monkeys. Nature Medicine, 3(1), 73–76. cord injury. Glia, 67(1), 125–145. https://doi.org/10.1002/glia.23533
Curt, A., Van Hedel, H. J., Klaus, D., Dietz, V., & Group, E.-S. S. (2008). Dyck, S., Kataria, H., Alizadeh, A., Santhosh, K. T., Lang, B., Silver, J., &
Recovery from a spinal cord injury: Significance of compensation, neu- Karimi-Abdolrezaee, S. (2018). Perturbing chondroitin sulfate proteo-
ral plasticity, and repair. Journal of Neurotrauma, 25(6), 677–685. glycan signaling through LAR and PTPsigma receptors promotes a ben-
https://doi.org/10.1089/neu.2007.0468 eficial inflammatory response following spinal cord injury. Journal of
Davalos, D., Grutzendler, J., Yang, G., Kim, J. V., Zuo, Y., Jung, S., … Neuroinflammation, 15(1), 90. https://doi.org/10.1186/s12974-018-
Gan, W. B. (2005). ATP mediates rapid microglial response to local 1128-2
brain injury in vivo. Nature Neuroscience, 8(6), 752–758. https://doi. Edgar, J. M., McLaughlin, M., Werner, H. B., McCulloch, M. C., Barrie, J. A.,
org/10.1038/nn1472 Brown, A., … Griffiths, I. R. (2009). Early ultrastructural defects of
de Castro, R., Jr., Tajrishi, R., Claros, J., & Stallcup, W. B. (2005). Differen- axons and axon-glia junctions in mice lacking expression of Cnp1. Glia,
tial responses of spinal axons to transection: Influence of the NG2 pro- 57(16), 1815–1824. https://doi.org/10.1002/glia.20893
teoglycan. Experimental Neurology, 192(2), 299–309. https://doi.org/ Ehrlich, M., Mozafari, S., Glatza, M., Starost, L., Velychko, S.,
10.1016/j.expneurol.2004.11.027 Hallmann, A. L., … Kuhlmann, T. (2017). Rapid and efficient generation
de Faria, O., Jr., Pama, E. A. C., Evans, K., Luzhynskaya, A., & of oligodendrocytes from human induced pluripotent stem cells using
Karadottir, R. T. (2018). Neuroglial interactions underpinning myelin transcription factors. Proceedings of the National Academy of Sciences
plasticity. Developmental Neurobiology, 78(2), 93–107. https://doi.org/ of the United States of America, 114(11), E2243–E2252. https://doi.
10.1002/dneu.22539 org/10.1073/pnas.1614412114
Deshmane, S. L., Kremlev, S., Amini, S., & Sawaya, B. E. (2009). Monocyte Emery, B., Agalliu, D., Cahoy, J. D., Watkins, T. A., Dugas, J. C.,
chemoattractant protein-1 (MCP-1): An overview. Journal of Inter- Mulinyawe, S. B., … Barres, B. A. (2009). Myelin gene regulatory factor is a
feron & Cytokine Research, 29(6), 313–326. https://doi.org/10.1089/ critical transcriptional regulator required for CNS myelination. Cell, 138(1),
jir.2008.0027 172–185. https://doi.org/10.1016/j.cell.2009.04.031
Di Maio, A., Skuba, A., Himes, B. T., Bhagat, S. L., Hyun, J. K., Tessler, A., … England, J. D., Gamboni, F., & Levinson, S. R. (1991). Increased numbers of
Son, Y. J. (2011). In vivo imaging of dorsal root regeneration: Rapid sodium channels form along demyelinated axons. Brain Research, 548
immobilization and presynaptic differentiation at the CNS/PNS (1–2), 334–337.
14 DUNCAN ET AL.

England, J. D., Gamboni, F., Levinson, S. R., & Finger, T. E. (1990). Changed Giacci, M. K., Bartlett, C. A., Smith, N. M., Iyer, K. S., Toomey, L. M., Jiang, H., …
distribution of sodium channels along demyelinated axons. Proceedings Fitzgerald, M. (2018). Oligodendroglia are particularly vulnerable to oxida-
of the National Academy of Sciences of the United States of America, 87 tive damage after Neurotrauma in vivo. The Journal of Neuroscience, 38
(17), 6777–6780. (29), 6491–6504. https://doi.org/10.1523/JNEUROSCI.1898-17.2018
Etxeberria, A., Hokanson, K. C., Dao, D. Q., Mayoral, S. R., Mei, F., Gibson, E. M., Purger, D., Mount, C. W., Goldstein, A. K., Lin, G. L., Wood, L. S.,
Redmond, S. A., … Chan, J. R. (2016). Dynamic modulation of mye- … Monje, M. (2014). Neuronal activity promotes oligodendrogenesis and
lination in response to visual stimuli alters optic nerve conduction adaptive myelination in the mammalian brain. Science, 344(6183),
velocity. The Journal of Neuroscience, 36(26), 6937–6948. https://doi. 1252304. https://doi.org/10.1126/science.1252304
org/10.1523/JNEUROSCI.0908-16.2016 Goldman, S. A., & Kuypers, N. J. (2015). How to make an oligodendrocyte.
Falcao, A. M., van Bruggen, D., Marques, S., Meijer, M., Jakel, S., Agirre, E., Development, 142(23), 3983–3995. https://doi.org/10.1242/dev.
… Castelo-Branco, G. (2018). Disease-specific oligodendrocyte lineage 126409
cells arise in multiple sclerosis. Nature Medicine, 24(12), 1837–1844. Goldschmidt, T., Antel, J., Konig, F. B., Bruck, W., & Kuhlmann, T. (2009).
https://doi.org/10.1038/s41591-018-0236-y Remyelination capacity of the MS brain decreases with disease chro-
Faulkner, J. R., Herrmann, J. E., Woo, M. J., Tansey, K. E., Doan, N. B., & nicity. Neurology, 72(22), 1914–1921. https://doi.org/10.1212/WNL.
Sofroniew, M. V. (2004). Reactive astrocytes protect tissue and pre- 0b013e3181a8260a
serve function after spinal cord injury. The Journal of Neuroscience, 24 Goritz, C., Dias, D. O., Tomilin, N., Barbacid, M., Shupliakov, O., & Frisen, J.
(9), 2143–2155. https://doi.org/10.1523/JNEUROSCI.3547-03.2004 (2011). A pericyte origin of spinal cord scar tissue. Science, 333(6039),
Felts, P. A., Baker, T. A., & Smith, K. J. (1997). Conduction in segmentally 238–242. https://doi.org/10.1126/science.1203165
demyelinated mammalian central axons. The Journal of Neuroscience, Griffiths, I., Klugmann, M., Anderson, T., Yool, D., Thomson, C.,
17(19), 7267–7277. Schwab, M. H., … Nave, K. A. (1998). Axonal swellings and degenera-
Felts, P. A., & Smith, K. J. (1987). Schwann-cell Remyelination of tion in mice lacking the major proteolipid of myelin. Science, 280
demyelinated central nerve-fibers restores secure conduction. Neuro- (5369), 1610–1613.
pathology and Applied Neurobiology, 13(6), 494–494. Guest, J. D., Hiester, E. D., & Bunge, R. P. (2005). Demyelination and
Fernandez-Castaneda, A., & Gaultier, A. (2016). Adult oligodendrocyte Schwann cell responses adjacent to injury epicenter cavities following
progenitor cells––Multifaceted regulators of the CNS in health and chronic human spinal cord injury. Experimental Neurology, 192(2),
disease. Brain, Behavior, and Immunity, 57, 1–7. https://doi.org/10. 384–393. https://doi.org/10.1016/j.expneurol.2004.11.033
1016/j.bbi.2016.01.005 Guizar-Sahagun, G., Grijalva, I., Salgado-Ceballos, H., Espitia, A., Orozco, S.,
Filli, L., Zorner, B., Weinmann, O., & Schwab, M. E. (2011). Motor deficits Ibarra, A., … Madrazo, I. (2004). Spontaneous and induced aberrant
and recovery in rats with unilateral spinal cord hemisection mimic the sprouting at the site of injury is irrelevant to motor function outcome
Brown-Sequard syndrome. Brain, 134(Pt 8), 2261–2273. https://doi. in rats with spinal cord injury. Brain Research, 1013(2), 143–151.
org/10.1093/brain/awr167 https://doi.org/10.1016/j.brainres.2004.03.062
Filous, A. R., Tran, A., Howell, C. J., Busch, S. A., Evans, T. A., Hackett, A. R., Lee, D. H., Dawood, A., Rodriguez, M., Funk, L.,
Stallcup, W. B., … Silver, J. (2014). Entrapment via synaptic-like con- Tsoulfas, P., & Lee, J. K. (2016). STAT3 and SOCS3 regulate NG2 cell
nections between NG2 proteoglycan+ cells and dystrophic axons in proliferation and differentiation after contusive spinal cord injury. Neu-
the lesion plays a role in regeneration failure after spinal cord injury. robiology of Disease, 89, 10–22. https://doi.org/10.1016/j.nbd.2016.
The Journal of Neuroscience, 34(49), 16369–16384. https://doi.org/10. 01.017
1523/JNEUROSCI.1309-14.2014 Hackett, A. R., Yahn, S. L., Lyapichev, K., Dajnoki, A., Lee, D. H.,
Foerster, S., Hill, M. F. E., & Franklin, R. J. M. (2019). Diversity in the oligo- Rodriguez, M., … Lee, J. K. (2018). Injury type-dependent differentia-
dendrocyte lineage: Plasticity or heterogeneity? Glia., 67(10), 1797– tion of NG2 glia into heterogeneous astrocytes. Experimental Neurol-
1805. https://doi.org/10.1002/glia.23607 ogy, 308, 72–79. https://doi.org/10.1016/j.expneurol.2018.07.001
Fogarty, M., Richardson, W. D., & Kessaris, N. (2005). A subset of oligo- Hains, B. C., Saab, C. Y., Lo, A. C., & Waxman, S. G. (2004). Sodium channel
dendrocytes generated from radial glia in the dorsal spinal cord. Devel- blockade with phenytoin protects spinal cord axons, enhances axonal
opment, 132(8), 1951–1959. https://doi.org/10.1242/dev.01777 conduction, and improves functional motor recovery after contusion
Ford, M. C., Alexandrova, O., Cossell, L., Stange-Marten, A., Sinclair, J., SCI. Experimental Neurology, 188(2), 365–377. https://doi.org/10.
Kopp-Scheinpflug, C., … Grothe, B. (2015). Tuning of Ranvier node and 1016/j.expneurol.2004.04.001
internode properties in myelinated axons to adjust action potential Hamanaka, G., Ohtomo, R., Takase, H., Lok, J., & Arai, K. (2018). Role of
timing. Nature Communications, 6, 8073. https://doi.org/10.1038/ oligodendrocyte-neurovascular unit in white matter repair. Neurosci-
ncomms9073 ence Letters, 684, 175–180. https://doi.org/10.1016/j.neulet.2018.
Gallo, V., Mangin, J. M., Kukley, M., & Dietrich, D. (2008). Synapses on 07.016
NG2-expressing progenitors in the brain: Multiple functions? The Jour- Hamilton, N. B., Clarke, L. E., Arancibia-Carcamo, I. L., Kougioumtzidou, E.,
nal of Physiology, 586(16), 3767–3781. https://doi.org/10.1113/ Matthey, M., Karadottir, R., … Attwell, D. (2017). Endogenous GABA
jphysiol.2008.158436 controls oligodendrocyte lineage cell number, myelination, and CNS
Gallo, V., Zhou, J. M., McBain, C. J., Wright, P., Knutson, P. L., & internode length. Glia, 65(2), 309–321. https://doi.org/10.1002/glia.
Armstrong, R. C. (1996). Oligodendrocyte progenitor cell proliferation 23093
and lineage progression are regulated by glutamate receptor-mediated Hanna, A. S., Son, Y. J., & Dempsey, R. (2011). Live imaging of dorsal root
K+ channel block. The Journal of Neuroscience, 16(8), 2659–2670. regeneration and the resurgence of a forgotten idea. Neurosurgery, 69(2),
Gauthier, M. K., Kosciuczyk, K., Tapley, L., & Karimi-Abdolrezaee, S. N18–N21. https://doi.org/10.1227/01.neu.0000400019.16401.c3
(2013). Dysregulation of the neuregulin-1-ErbB network modulates Harris, J. J., & Attwell, D. (2012). The energetics of CNS white matter. The
endogenous oligodendrocyte differentiation and preservation after Journal of Neuroscience, 32(1), 356–371. https://doi.org/10.1523/
spinal cord injury. The European Journal of Neuroscience, 38(5), JNEUROSCI.3430-11.2012
2693–2715. https://doi.org/10.1111/ejn.12268 Hawryluk, G. W. J., Spano, S., Chew, D., Wang, S., Erwin, M.,
Gautier, H. O., Evans, K. A., Volbracht, K., James, R., Sitnikov, S., Lundgaard, I., Chamankhah, M., … Fehlings, M. G. (2014). An examination of the
… Karadottir, R. T. (2015). Neuronal activity regulates remyelination via mechanisms by which neural precursors augment recovery following
glutamate signalling to oligodendrocyte progenitors. Nature Communica- spinal cord injury: A key Role for Remyelination. Cell Transplantation,
tions, 6, 8518. https://doi.org/10.1038/ncomms9518 23(3), 365–380. https://doi.org/10.3727/096368912x662408
DUNCAN ET AL. 15

Hayashi, S., & McMahon, A. P. (2002). Efficient recombination in diverse Hughes, E. G., Kang, S. H., Fukaya, M., & Bergles, D. E. (2013). Oligoden-
tissues by a tamoxifen-inducible form of Cre: A tool for temporally drocyte progenitors balance growth with self-repulsion to achieve
regulated gene activation/inactivation in the mouse. Developmental homeostasis in the adult brain. Nature Neuroscience, 16(6), 668–676.
Biology, 244(2), 305–318. https://doi.org/10.1006/dbio.2002.0597 https://doi.org/10.1038/nn.3390
Herrmann, J. E., Imura, T., Song, B. B., Qi, J. W., Ao, Y., Nguyen, T. K., … Hughes, E. G., Orthmann-Murphy, J. L., Langseth, A. J., & Bergles, D. E. (2018).
Sofroniew, M. V. (2008). STAT3 is a critical regulator of astrogliosis Myelin remodeling through experience-dependent oligodendrogenesis in
and scar formation after spinal cord injury. Journal of Neuroscience, 28 the adult somatosensory cortex. Nature Neuroscience, 21(5), 696–706.
(28), 7231–7243. https://doi.org/10.1523/Jneurosci.1709-08.2008 https://doi.org/10.1038/s41593-018-0121-5
Hesp, Z. C., Goldstein, E. Z., Miranda, C. J., Kaspar, B. K., & McTigue, D. M. Hunanyan, A. S., Garcia-Alias, G., Alessi, V., Levine, J. M., Fawcett, J. W.,
(2015). Chronic oligodendrogenesis and remyelination after spinal cord Mendell, L. M., & Arvanian, V. L. (2010). Role of chondroitin sulfate
injury in mice and rats. The Journal of Neuroscience, 35(3), 1274–1290. proteoglycans in axonal conduction in mammalian spinal cord. The
https://doi.org/10.1523/JNEUROSCI.2568-14.2015 Journal of Neuroscience, 30(23), 7761–7769. https://doi.org/10.1523/
Hesp, Z. C., Yoseph, R. Y., Suzuki, R., Jukkola, P., Wilson, C., JNEUROSCI.4659-09.2010
Nishiyama, A., & McTigue, D. M. (2018). Proliferating NG2-cell- Hwang, D. H., Kim, B. G., Kim, E. J., Lee, S. I., Joo, I. S., Suh-Kim, H., …
dependent angiogenesis and scar formation Alter axon growth and Kim, S. U. (2009). Transplantation of human neural stem cells trans-
functional recovery after spinal cord injury in mice. The Journal of Neu- duced with Olig2 transcription factor improves locomotor recovery
roscience, 38(6), 1366–1382. https://doi.org/10.1523/JNEUROSCI. and enhances myelination in the white matter of rat spinal cord fol-
3953-16.2017 lowing contusive injury. BMC Neuroscience, 10. doi:Artn 117, 117.
Hill, R. A., Li, A. M., & Grutzendler, J. (2018). Lifelong cortical myelin plas- https://doi.org/10.1186/1471-2202-10-117
ticity and age-related degeneration in the live mammalian brain. Nature James, N. D., Bartus, K., Grist, J., Bennett, D. L., McMahon, S. B., &
Neuroscience, 21, 683–695. https://doi.org/10.1038/s41593-018- Bradbury, E. J. (2011). Conduction failure following spinal cord injury:
0120-6 Functional and anatomical changes from acute to chronic stages. The
Hill, R. A., Patel, K. D., Goncalves, C. M., Grutzendler, J., & Nishiyama, A. Journal of Neuroscience, 31(50), 18543–18555. https://doi.org/10.
(2014). Modulation of oligodendrocyte generation during a critical 1523/JNEUROSCI.4306-11.2011
temporal window after NG2 cell division. Nature Neuroscience, 17(11), Jeffery, N. D., & Blakemore, W. F. (1995). Remyelination of mouse spinal
1518–1527. https://doi.org/10.1038/nn.3815 cord axons demyelinated by local injection of lysolecithin. Journal of
Hilton, B. J., Anenberg, E., Harrison, T. C., Boyd, J. D., Murphy, T. H., & Neurocytology, 24(10), 775–781.
Tetzlaff, W. (2016). Re-establishment of cortical motor output maps Jeffery, N. D., Smith, P. M., Lakatos, A., Ibanez, C., Ito, D., & Franklin, R. J.
and spontaneous functional recovery via spared Dorsolaterally (2006). Clinical canine spinal cord injury provides an opportunity to
projecting Corticospinal neurons after dorsal column spinal cord injury examine the issues in translating laboratory techniques into practical
in adult mice. The Journal of Neuroscience, 36(14), 4080–4092. https:// therapy. Spinal Cord, 44(10), 584–593. https://doi.org/10.1038/sj.sc.
doi.org/10.1523/JNEUROSCI.3386-15.2016 3101912
Hilton, B. J., Moulson, A. J., & Tetzlaff, W. (2017). Neuroprotection and Kadoya, K., Tsukada, S., Lu, P., Coppola, G., Geschwind, D., Filbin, M. T., …
secondary damage following spinal cord injury: Concepts and Tuszynski, M. H. (2009). Combined intrinsic and extrinsic neuronal
methods. Neuroscience Letters, 652, 3–10. https://doi.org/10.1016/j. mechanisms facilitate bridging axonal regeneration one year after spi-
neulet.2016.12.004 nal cord injury. Neuron, 64(2), 165–172. https://doi.org/10.1016/j.
Hilton, B. J., & Tetzlaff, W. (2018). A brainstem bypass for spinal cord neuron.2009.09.016
injury. Nature Neuroscience, 21(4), 457–458. https://doi.org/10.1038/ Kakulas, A. (1988). The applied neurobiology of human spinal cord injury:
s41593-018-0099-z A review. Paraplegia, 26(6), 371–379. https://doi.org/10.1038/sc.
Hofstetter, C. P., Holmstrom, N. A. V., Lilja, J. A., Schweinhardt, P., 1988.57
Hao, J. X., Spenger, C., … Olson, L. (2005). Allodynia limits the useful- Kakulas, B. A. (2004). Neuropathology: The foundation for new treatments
ness of intraspinal neural stem cell grafts; directed differentiation in spinal cord injury. Spinal Cord, 42(10), 549–563. https://doi.org/10.
improves outcome. Nature Neuroscience, 8(3), 346–353. https://doi. 1038/sj.sc.3101670
org/10.1038/Nn1405 Kakulas, B. A., & Kaelan, C. (2015). The neuropathological foundations for
Honmou, O., Felts, P. A., Waxman, S. G., & Kocsis, J. D. (1996). Restoration the restorative neurology of spinal cord injury. Clinical Neurology and
of normal conduction properties in demyelinated spinal cord axons in Neurosurgery, 129(Suppl 1), S1–S7. https://doi.org/10.1016/j.clineuro.
the adult rat by transplantation of exogenous Schwann cells. Journal of 2015.01.012
Neuroscience, 16(10), 3199–3208. Kang, S. H., Fukaya, M., Yang, J. K., Rothstein, J. D., & Bergles, D. E.
Horky, L. L., Galimi, F., Gage, F. H., & Horner, P. J. (2006). Fate of endoge- (2010). NG2+ CNS glial progenitors remain committed to the oligo-
nous stem/progenitor cells following spinal cord injury. The Journal of dendrocyte lineage in postnatal life and following neurodegeneration.
Comparative Neurology, 498(4), 525–538. https://doi.org/10.1002/ Neuron, 68(4), 668–681. https://doi.org/10.1016/j.neuron.2010.
cne.21065 09.009
Huang, W., Bai, X., Stopper, L., Catalin, B., Cartarozzi, L. P., Scheller, A., & Kang, Z., Wang, C., Zepp, J., Wu, L., Sun, K., Zhao, J., … Li, X. (2013). Act1
Kirchhoff, F. (2018). During development NG2 glial cells of the spinal mediates IL-17-induced EAE pathogenesis selectively in NG2+ glial
cord are restricted to the Oligodendrocyte lineage, but generate astro- cells. Nature Neuroscience, 16(10), 1401–1408. https://doi.org/10.
cytes upon acute injury. Neuroscience, 385, 154–165. https://doi.org/ 1038/nn.3505
10.1016/j.neuroscience.2018.06.015 Kanno, H., Pearse, D. D., Ozawa, H., Itoi, E., & Bunge, M. B. (2015).
Huang, W., Guo, Q., Bai, X., Scheller, A., & Kirchhoff, F. (2019). Early Schwann cell transplantation for spinal cord injury repair: Its significant
embryonic NG2 glia are exclusively gliogenic and do not generate neu- therapeutic potential and prospectus. Reviews in the Neurosciences, 26
rons in the brain. Glia, 67(6), 1094–1103. https://doi.org/10.1002/ (2), 121–128. https://doi.org/10.1515/revneuro-2014-0068
glia.23590 Karadottir, R., Hamilton, N. B., Bakiri, Y., & Attwell, D. (2008). Spiking and non-
Hubscher, C. H., & Johnson, R. D. (2006). Chronic spinal cord injury spiking classes of oligodendrocyte precursor glia in CNS white matter.
induced changes in the responses of thalamic neurons. Experimental Nature Neuroscience, 11(4), 450–456. https://doi.org/10.1038/nn2060
Neurology, 197(1), 177–188. https://doi.org/10.1016/j.expneurol. Karimi-Abdolrezaee, S., Eftekharpour, E., & Fehlings, M. G. (2004). Tempo-
2005.09.007 ral and spatial patterns of Kv1.1 and Kv1.2 protein and gene
16 DUNCAN ET AL.

expression in spinal cord white matter after acute and chronic spinal Lillie, R. S. (1925). Factors affecting transmission and recovery in the pas-
cord injury in rats: Implications for axonal pathophysiology after sive iron nerve model. Journal of General Physiology, 7(4), 473–507.
neurotrauma. The European Journal of Neuroscience, 19(3), 577–589. https://doi.org/10.1085/Jgp.7.4.473
Karimi-Abdolrezaee, S., Eftekharpour, E., Wang, J., Morshead, C. M., & Liu, J., Dietz, K., DeLoyht, J. M., Pedre, X., Kelkar, D., Kaur, J., …
Fehlings, M. G. (2006). Delayed transplantation of adult neural precur- Casaccia, P. (2012). Impaired adult myelination in the prefrontal cortex
sor cells promotes remyelination and functional neurological recovery of socially isolated mice. Nature Neuroscience, 15(12), 1621–1623.
after spinal cord injury. Journal of Neuroscience, 26(13), 3377–3389. https://doi.org/10.1038/nn.3263
https://doi.org/10.1523/Jneurosci.4184-05.2006 Liu, S., Yin, F., Zhang, J., & Qian, Y. (2014). The role of calpains in traumatic
Karimi-Abdolrezaee, S., Eftekharpour, E., Wang, J., Schut, D., & brain injury. Brain Injury, 28(2), 133–137. https://doi.org/10.3109/
Fehlings, M. G. (2010). Synergistic effects of transplanted adult neural 02699052.2013.860479
stem/progenitor cells, chondroitinase, and growth factors promote Liuzzi, F. J., & Lasek, R. J. (1987). Astrocytes block axonal regeneration in
functional repair and plasticity of the chronically injured spinal cord. mammals by activating the physiological stop pathway. Science, 237
The Journal of Neuroscience, 30(5), 1657–1676. https://doi.org/10. (4815), 642–645.
1523/JNEUROSCI.3111-09.2010 Ma, D., Wang, B., Zawadzka, M., Gonzalez, G., Wu, Z., Yu, B., … Zhao, C.
Kataria, H., Alizadeh, A., Shahriary, G. M., Saboktakin Rizi, S., Henrie, R., (2018). A subpopulation of Foxj1-expressing, Nonmyelinating
Santhosh, K. T., … Karimi-Abdolrezaee, S. (2018). Neuregulin-1 pro- Schwann cells of the peripheral nervous system contribute to
motes remyelination and fosters a pro-regenerative inflammatory Schwann cell Remyelination in the central nervous system. The Journal
response in focal demyelinating lesions of the spinal cord. Glia, 66(3), of Neuroscience, 38(43), 9228–9239. https://doi.org/10.1523/
538–561. https://doi.org/10.1002/glia.23264 JNEUROSCI.0585-18.2018
Keirstead, H. S., Nistor, G., Bernal, G., Totoiu, M., Cloutier, F., Sharp, K., & Makinodan, M., Rosen, K. M., Ito, S., & Corfas, G. (2012). A critical period
Steward, O. (2005). Human embryonic stem cell-derived oligodendro- for social experience-dependent oligodendrocyte maturation and mye-
cyte progenitor cell transplants remyelinate and restore locomotion lination. Science, 337(6100), 1357–1360. https://doi.org/10.1126/
after spinal cord injury. The Journal of Neuroscience, 25(19), science.1220845
Marques, S., van Bruggen, D., Vanichkina, D. P., Floriddia, E. M.,
4694–4705. https://doi.org/10.1523/JNEUROSCI.0311-05.2005
Munguba, H., Varemo, L., … Castelo-Branco, G. (2018). Transcriptional
Keough, M. B., Rogers, J. A., Zhang, P., Jensen, S. K., Stephenson, E. L.,
convergence of Oligodendrocyte lineage progenitors during develop-
Chen, T., … Yong, V. W. (2016). An inhibitor of chondroitin sulfate pro-
ment. Developmental Cell, 46(4), 504–517 e507. https://doi.org/10.
teoglycan synthesis promotes central nervous system remyelination.
1016/j.devcel.2018.07.005
Nature Communications, 7, 11312. https://doi.org/10.1038/
McTigue, D. M., Horner, P. J., Stokes, B. T., & Gage, F. H. (1998). Neuro-
ncomms11312
trophin-3 and brain-derived neurotrophic factor induce oligodendro-
Kirby, L., Jin, J., Gonzalez-Cardona, J., Smith, M., Martin, K., Wang, J., …
cyte proliferation and myelination of regenerating axons in the
Calabresi, P. A. (2018). Oligodendrocyte precursor cells are co-opted
contused adult rat spinal cord. The Journal of Neuroscience, 18(14),
by the immune system to cross-present antigen and mediate cytotox-
5354–5365.
icity. bioRxiv, 461434. https://doi.org/10.1101/461434
McTigue, D. M., Tripathi, R., & Wei, P. (2006). NG2 colocalizes with axons
Kirschner, D. A., & Ganser, A. L. (1980). Compact myelin exists in the
and is expressed by a mixed cell population in spinal cord lesions. Jour-
absence of basic protein in the shiverer mutant mouse. Nature, 283
nal of Neuropathology and Experimental Neurology, 65(4), 406–420.
(5743), 207–210.
https://doi.org/10.1097/01.jnen.0000218447.32320.52
Kukley, M., Capetillo-Zarate, E., & Dietrich, D. (2007). Vesicular glutamate
McTigue, D. M., Wei, P., & Stokes, B. T. (2001). Proliferation of
release from axons in white matter. Nature Neuroscience, 10(3),
NG2-positive cells and altered oligodendrocyte numbers in the con-
311–320. https://doi.org/10.1038/nn1850
tused rat spinal cord. The Journal of Neuroscience, 21(10), 3392–3400.
Kwon, B. K., Tetzlaff, W., Grauer, J. N., Beiner, J., & Vaccaro, A. R. (2004).
Mei, F., Lehmann-Horn, K., Shen, Y. A., Rankin, K. A., Stebbins, K. J.,
Pathophysiology and pharmacologic treatment of acute spinal cord
Lorrain, D. S., … Chan, J. R. (2016). Accelerated remyelination during
injury. The Spine Journal, 4(4), 451–464. https://doi.org/10.1016/j.
inflammatory demyelination prevents axonal loss and improves func-
spinee.2003.07.007
tional recovery. eLife, 5, 5. https://doi.org/10.7554/eLife.18246
Lachapelle, F., Bachelin, C., Moissonnier, P., Nait-Oumesmar, B.,
Metz, G. A., Curt, A., van de Meent, H., Klusman, I., Schwab, M. E., &
Hidalgo, A., Fontaine, D., & Baron-Van Evercooren, A. (2005). Failure
Dietz, V. (2000). Validation of the weight-drop contusion model in
of remyelination in the nonhuman primate optic nerve. Brain Pathol-
rats: A comparative study of human spinal cord injury. Journal of
ogy, 15(3), 198–207. Neurotrauma, 17(1), 1–17. https://doi.org/10.1089/neu.2000.17.1
Lasiene, J., Shupe, L., Perlmutter, S., & Horner, P. (2008). No evidence for Michailov, G. V., Sereda, M. W., Brinkmann, B. G., Fischer, T. M., Haug, B.,
chronic demyelination in spared axons after spinal cord injury in a Birchmeier, C., … Nave, K. A. (2004). Axonal neuregulin-1 regulates
mouse. The Journal of Neuroscience, 28(15), 3887–3896. https://doi. myelin sheath thickness. Science, 304(5671), 700–703. https://doi.
org/10.1523/JNEUROSCI.4756-07.2008 org/10.1126/science.1095862
Lee, Y., Morrison, B. M., Li, Y., Lengacher, S., Farah, M. H., Hoffman, P. N., Mitew, S., Gobius, I., Fenlon, L. R., McDougall, S. J., Hawkes, D., Xing, Y. L.,
… Rothstein, J. D. (2012). Oligodendroglia metabolically support axons … Emery, B. (2018). Pharmacogenetic stimulation of neuronal activity
and contribute to neurodegeneration. Nature, 487(7408), 443–448. increases myelination in an axon-specific manner. Nature Communica-
https://doi.org/10.1038/nature11314 tions, 9(1), 306. https://doi.org/10.1038/s41467-017-02719-2
Lehninger, A. L., Vercesi, A., & Bababunmi, E. A. (1978). Regulation of Ca2 Monteiro de Castro, G., Deja, N. A., Ma, D., Zhao, C., & Franklin, R. J.
+ release from mitochondria by the oxidation-reduction state of pyri- (2015). Astrocyte activation via Stat3 signaling determines the balance
dine nucleotides. Proceedings of the National Academy of Sciences of of Oligodendrocyte versus Schwann cell Remyelination. The American
the United States of America, 75(4), 1690–1694. Journal of Pathology, 185(9), 2431–2440. https://doi.org/10.1016/j.
Li, Y., & Raisman, G. (1995). Sprouts from cut corticospinal axons persist in ajpath.2015.05.011
the presence of astrocytic scarring in long-term lesions of the adult rat Moyon, S., Dubessy, A. L., Aigrot, M. S., Trotter, M., Huang, J. K.,
spinal cord. Experimental Neurology, 134(1), 102–111. https://doi.org/ Dauphinot, L., … Lubetzki, C. (2015). Demyelination causes adult CNS
10.1006/exnr.1995.1041 progenitors to revert to an immature state and express immune cues
DUNCAN ET AL. 17

that support their migration. The Journal of Neuroscience, 35(1), 4–20. transmembrane receptor with multiple domains implicated in protein-
https://doi.org/10.1523/JNEUROSCI.0849-14.2015 protein interactions. The EMBO Journal, 16(5), 978–988. https://doi.
Murray, K. C., Nakae, A., Stephens, M. J., Rank, M., D'Amico, J., org/10.1093/emboj/16.5.978
Harvey, P. J., … Fouad, K. (2010). Recovery of motoneuron and loco- Petrosyan, H. A., Hunanyan, A. S., Alessi, V., Schnell, L., Levine, J., &
motor function after spinal cord injury depends on constitutive activity Arvanian, V. L. (2013). Neutralization of inhibitory molecule NG2
in 5-HT2C receptors. Nature Medicine, 16(6), 694–700. https://doi. improves synaptic transmission, retrograde transport, and locomotor
org/10.1038/nm.2160 function after spinal cord injury in adult rats. The Journal of Neuroscience,
Myers, S. A., Bankston, A. N., Burke, D. A., Ohri, S. S., & Whittemore, S. R. 33(9), 4032–4043. https://doi.org/10.1523/JNEUROSCI.4702-12.2013
(2016). Does the preclinical evidence for functional remyelination fol- Pham, L. D., Hayakawa, K., Seo, J. H., Nguyen, M. N., Som, A. T., Lee, B. J.,
lowing myelinating cell engraftment into the injured spinal cord sup- … Arai, K. (2012). Crosstalk between oligodendrocytes and cerebral
port progression to clinical trials? Experimental Neurology, 283(Pt B), endothelium contributes to vascular remodeling after white matter
560–572. https://doi.org/10.1016/j.expneurol.2016.04.009 injury. Glia, 60(6), 875–881. https://doi.org/10.1002/glia.22320
Nashmi, R., & Fehlings, M. G. (2001). Mechanisms of axonal dysfunction Philips, T., & Rothstein, J. D. (2017). Oligodendroglia: Metabolic supporters
after spinal cord injury: With an emphasis on the role of voltage-gated of neurons. The Journal of Clinical Investigation, 127(9), 3271–3280.
potassium channels. Brain Research. Brain Research Reviews, 38(1–2), https://doi.org/10.1172/JCI90610
165–191. Piltti, K. M., Funes, G. M., Avakian, S. N., Salibian, A. A., Huang, K. I.,
Nave, K. A. (2010). Myelination and support of axonal integrity by glia. Carta, K., … Anderson, A. J. (2017). Increasing human neural stem cell
Nature, 468(7321), 244–252. https://doi.org/10.1038/nature09614 transplantation dose alters Oligodendroglial and neuronal differentia-
Nave, K. A., & Werner, H. B. (2014). Myelination of the nervous system: tion after spinal cord injury. Stem Cell Reports, 8(6), 1534–1548.
Mechanisms and functions. Annual Review of Cell and Developmental https://doi.org/10.1016/j.stemcr.2017.04.009
Biology, 30, 503–533. https://doi.org/10.1146/annurev-cellbio- Plemel, J. R., Keough, M. B., Duncan, G. J., Sparling, J. S., Yong, V. W.,
100913-013101 Stys, P. K., & Tetzlaff, W. (2014). Remyelination after spinal cord
Nishiyama, A., Lin, X. H., Giese, N., Heldin, C. H., & Stallcup, W. B. (1996). injury: Is it a target for repair? Progress in Neurobiology, 117, 54–72.
Co-localization of NG2 proteoglycan and PDGF alpha-receptor on https://doi.org/10.1016/j.pneurobio.2014.02.006
O2A progenitor cells in the developing rat brain. Journal of Neurosci- Plemel, J. R., Manesh, S. B., Sparling, J. S., & Tetzlaff, W. (2013). Myelin
ence Research, 43(3), 299–314. inhibits oligodendroglial maturation and regulates oligodendrocytic
Norenberg, M. D., Smith, J., & Marcillo, A. (2004). The pathology of human transcription factor expression. Glia, 61(9), 1471–1487. https://doi.
spinal cord injury: Defining the problems. Journal of Neurotrauma, 21 org/10.1002/Glia.22535
(4), 429–440. https://doi.org/10.1089/089771504323004575 Posmantur, R., Kampfl, A., Siman, R., Liu, J., Zhao, X., Clifton, G. L., &
Nori, S., Khazaei, M., Ahuja, C. S., Yokota, K., Ahlfors, J. E., Liu, Y., … Hayes, R. L. (1997). A calpain inhibitor attenuates cortical cytoskeletal
Fehlings, M. G. (2018). Human Oligodendrogenic neural progenitor protein loss after experimental traumatic brain injury in the rat. Neuro-
cells delivered with Chondroitinase ABC facilitate functional repair of science, 77(3), 875–888.
chronic spinal cord injury. Stem Cell Reports, 11(6), 1433–1448. Powers, B. E., Lasiene, J., Plemel, J. R., Shupe, L., Perlmutter, S. I.,
https://doi.org/10.1016/j.stemcr.2018.10.017 Tetzlaff, W., & Horner, P. J. (2012). Axonal thinning and extensive
Oblinger, M. M., Foe, L. G., Kwiatkowska, D., & Kemp, R. G. (1988). Phos- remyelination without chronic demyelination in spinal injured rats. The
phofructokinase in the rat nervous system: Regional differences in Journal of Neuroscience, 32(15), 5120–5125. https://doi.org/10.1523/
activity and characteristics of axonal transport. Journal of Neuroscience JNEUROSCI.0002-12.2012
Research, 21(1), 25–34. https://doi.org/10.1002/jnr.490210105 Powers, B. E., Sellers, D. L., Lovelett, E. A., Cheung, W., Aalami, S. P.,
Okubo, T., Nagoshi, N., Kohyama, J., Tsuji, O., Shinozaki, M., Shibata, S., … Zapertov, N., … Horner, P. J. (2013). Remyelination reporter reveals
Okano, H. (2018). Treatment with a gamma-Secretase inhibitor pro- prolonged refinement of spontaneously regenerated myelin. Proceed-
motes functional recovery in human iPSC- derived transplants for ings of the National Academy of Sciences of the United States of America,
chronic spinal cord injury. Stem Cell Reports, 11(6), 1416–1432. 110(10), 4075–4080. https://doi.org/10.1073/pnas.1210293110
https://doi.org/10.1016/j.stemcr.2018.10.022 Pringle, N. P., Mudhar, H. S., Collarini, E. J., & Richardson, W. D. (1992).
Pallini, R., Fernandez, E., & Sbriccoli, A. (1988). Retrograde degeneration of PDGF receptors in the rat CNS: During late neurogenesis, PDGF
corticospinal axons following transection of the spinal cord in rats. A alpha-receptor expression appears to be restricted to glial cells of the
quantitative study with anterogradely transported horseradish peroxi- oligodendrocyte lineage. Development, 115(2), 535–551.
dase. Journal of Neurosurgery, 68(1), 124–128. https://doi.org/10. Pringle, N. P., & Richardson, W. D. (1993). A singularity of PDGF alpha-
3171/jns.1988.68.1.0124 receptor expression in the dorsoventral axis of the neural tube may define
Papastefanaki, F., & Matsas, R. (2015). From demyelination to the origin of the oligodendrocyte lineage. Development, 117(2), 525–533.
remyelination: The road toward therapies for spinal cord injury. Glia, Pu, A., Stephenson, E. L., & Yong, V. W. (2018). The extracellular matrix: Focus
63(7), 1101–1125. https://doi.org/10.1002/glia.22809 on oligodendrocyte biology and targeting CSPGs for remyelination thera-
Patrikios, P., Stadelmann, C., Kutzelnigg, A., Rauschka, H., pies. Glia, 66(9), 1809–1825. https://doi.org/10.1002/glia.23333
Schmidbauer, M., Laursen, H., … Lassmann, H. (2006). Remyelination is Pukos, N., Goodus, M. T., Sahinkaya, F. R., & McTigue, D. (2019). Persis-
extensive in a subset of multiple sclerosis patients. Brain, 129, tent oligodendrocyte formation and remyelination occurs in a dynamic
3165–3172. https://doi.org/10.1093/Brain/Awl217 microenvironment after spinal cord injury. Glia, in press.
Pearse, D. D., Pereira, F. C., Marcillo, A. E., Bates, M. L., Berrocal, Y. A., Raff, M. C., Lillien, L. E., Richardson, W. D., Burne, J. F., & Noble, M. D.
Filbin, M. T., & Bunge, M. B. (2004). cAMP and Schwann cells promote (1988). Platelet-derived growth factor from astrocytes drives the clock
axonal growth and functional recovery after spinal cord injury. Nature that times oligodendrocyte development in culture. Nature, 333(6173),
Medicine, 10(6), 610–616. https://doi.org/10.1038/Nm1056 562–565. https://doi.org/10.1038/333562a0
Pedraza, L., Huang, J. K., & Colman, D. (2009). Disposition of axonal caspr Raineteau, O., Fouad, K., Bareyre, F. M., & Schwab, M. E. (2002). Reorgani-
with respect to glial cell membranes: Implications for the process of zation of descending motor tracts in the rat spinal cord. The European
myelination. Journal of Neuroscience Research, 87(15), 3480–3491. Journal of Neuroscience, 16(9), 1761–1771.
https://doi.org/10.1002/jnr.22004 Ramer, M. S., Priestley, J. V., & McMahon, S. B. (2000). Functional regener-
Peles, E., Nativ, M., Lustig, M., Grumet, M., Schilling, J., Martinez, R., … ation of sensory axons into the adult spinal cord. Nature, 403(6767),
Schlessinger, J. (1997). Identification of a novel contactin-associated 312–316. https://doi.org/10.1038/35002084
18 DUNCAN ET AL.

Rasband, M. N., Trimmer, J. S., Schwarz, T. L., Levinson, S. R., Snaidero, N., Mobius, W., Czopka, T., Hekking, L. H., Mathisen, C.,
Ellisman, M. H., Schachner, M., & Shrager, P. (1998). Potassium chan- Verkleij, D., … Simons, M. (2014). Myelin membrane wrapping of CNS
nel distribution, clustering, and function in remyelinating rat axons. axons by PI(3,4,5)P3-dependent polarized growth at the inner tongue.
The Journal of Neuroscience, 18(1), 36–47. Cell, 156(1–2), 277–290. https://doi.org/10.1016/j.cell.2013.11.044
Redford, E. J., Kapoor, R., & Smith, K. J. (1997). Nitric oxide donors revers- Snaidero, N., Velte, C., Myllykoski, M., Raasakka, A., Ignatev, A.,
ibly block axonal conduction: Demyelinated axons are especially sus- Werner, H. B., … Simons, M. (2017). Antagonistic functions of MBP
ceptible. Brain, 120(Pt 12), 2149–2157. and CNP establish cytosolic channels in CNS myelin. Cell Reports, 18
Rempe, R. G., Hartz, A. M. S., & Bauer, B. (2016). Matrix (2), 314–323. https://doi.org/10.1016/j.celrep.2016.12.053
metalloproteinases in the brain and blood-brain barrier: Versatile brea- Sofroniew, M. V. (2015). Astrocyte barriers to neurotoxic inflammation.
kers and makers. Journal of Cerebral Blood Flow and Metabolism, 36(9), Nature Reviews. Neuroscience, 16(5), 249–263. https://doi.org/10.
1481–1507. https://doi.org/10.1177/0271678X16655551 1038/nrn3898
Richardson, W. D., Pringle, N., Mosley, M. J., Westermark, B., & Dubois- Son, Y. J. (2015). Synapsing with NG2 cells (polydendrocytes),
Dalcq, M. (1988). A role for platelet-derived growth factor in normal unappreciated barrier to axon regeneration? Neural Regeneration
gliogenesis in the central nervous system. Cell, 53(2), 309–319. Research, 10(3), 346–348. https://doi.org/10.4103/1673-5374.
Rodriguez, J. P., Coulter, M., Miotke, J., Meyer, R. L., Takemaru, K., & 153672
Levine, J. M. (2014). Abrogation of beta-catenin signaling in oligoden- Spitzer, S. O., Sitnikov, S., Kamen, Y., Evans, K. A., Kronenberg-
drocyte precursor cells reduces glial scarring and promotes axon Versteeg, D., Dietmann, S., … Karadottir, R. T. (2019). Oligodendrocyte
regeneration after CNS injury. The Journal of Neuroscience, 34(31), progenitor cells become regionally diverse and heterogeneous with
10285–10297. https://doi.org/10.1523/JNEUROSCI.4915-13.2014 age. Neuron, 101, 459–471.e5. https://doi.org/10.1016/j.neuron.
Rossignol, S., & Frigon, A. (2011). Recovery of locomotion after spinal cord 2018.12.020
injury: Some facts and mechanisms. Annual Review of Neuroscience, 34, Stirling, D. P., Khodarahmi, K., Liu, J., McPhail, L. T., McBride, C. B.,
413–440. https://doi.org/10.1146/annurev-neuro-061010-113746 Steeves, J. D., … Tetzlaff, W. (2004). Minocycline treatment reduces
Saab, A. S., & Nave, K. A. (2017). Myelin dynamics: Protecting and shaping delayed oligodendrocyte death, attenuates axonal dieback, and
neuronal functions. Current Opinion in Neurobiology, 47, 104–112. improves functional outcome after spinal cord injury. The Journal of
https://doi.org/10.1016/j.conb.2017.09.013 Neuroscience, 24(9), 2182–2190. https://doi.org/10.1523/
Sabelstrom, H., Stenudd, M., Reu, P., Dias, D. O., Elfineh, M., Zdunek, S., … JNEUROSCI.5275-03.2004
Frisen, J. (2013). Resident neural stem cells restrict tissue damage and Stys, P. K., Waxman, S. G., & Ransom, B. R. (1991). Na(+)-Ca2+ exchanger
neuronal loss after spinal cord injury in mice. Science, 342(6158), mediates Ca2+ influx during anoxia in mammalian central nervous sys-
637–640. https://doi.org/10.1126/science.1242576 tem white matter. Annals of Neurology, 30(3), 375–380. https://doi.
Schucht, P., Raineteau, O., Schwab, M. E., & Fouad, K. (2002). Anatomical org/10.1002/ana.410300309
correlates of locomotor recovery following dorsal and ventral lesions Tan, A. M., Colletti, M., Rorai, A. T., Skene, J. H., & Levine, J. M. (2006).
of the rat spinal cord. Experimental Neurology, 176(1), 143–153. Antibodies against the NG2 proteoglycan promote the regeneration of
Seif, G. I., Nomura, H., & Tator, C. H. (2007). Retrograde axonal degenera- sensory axons within the dorsal columns of the spinal cord. The Journal
tion "dieback" in the corticospinal tract after transection injury of the of Neuroscience, 26(18), 4729–4739. https://doi.org/10.1523/
rat spinal cord: A confocal microscopy study. Journal of Neurotrauma, JNEUROSCI.3900-05.2006
24(9), 1513–1528. https://doi.org/10.1089/neu.2007.0323 Tang, D. G., Tokumoto, Y. M., & Raff, M. C. (2000). Long-term culture of
Sellers, D. L., Maris, D. O., & Horner, P. J. (2009). Postinjury niches induce purified postnatal oligodendrocyte precursor cells. Evidence for an
temporal shifts in progenitor fates to direct lesion repair after spinal intrinsic maturation program that plays out over months. The Journal
cord injury. The Journal of Neuroscience, 29(20), 6722–6733. https:// of Cell Biology, 148(5), 971–984. https://doi.org/10.1083/jcb.148.
doi.org/10.1523/JNEUROSCI.4538-08.2009 5.971
Seo, J. H., Miyamoto, N., Hayakawa, K., Pham, L. D., Maki, T., Ayata, C., … Tetzlaff, W., Okon, E. B., Karimi-Abdolrezaee, S., Hill, C. E., Sparling, J. S.,
Arai, K. (2013). Oligodendrocyte precursors induce early blood-brain Plemel, J. R., … Kwon, B. K. (2011). A systematic review of cellular
barrier opening after white matter injury. The Journal of Clinical Investi- transplantation therapies for spinal cord injury. Journal of Neurotrauma,
gation, 123(2), 782–786. https://doi.org/10.1172/JCI65863 28(8), 1611–1682. https://doi.org/10.1089/neu.2009.1177
Shrager, P., Custer, A. W., Kazarinova, K., Rasband, M. N., & Mattson, D. Timsit, S., Martinez, S., Allinquant, B., Peyron, F., Puelles, L., & Zalc, B.
(1998). Nerve conduction block by nitric oxide that is mediated by the (1995). Oligodendrocytes originate in a restricted zone of the embry-
axonal environment. Journal of Neurophysiology, 79(2), 529–536. onic ventral neural tube defined by DM-20 mRNA expression. The
https://doi.org/10.1152/jn.1998.79.2.529 Journal of Neuroscience, 15(2), 1012–1024.
Shultz, R. B., Wang, Z., Nong, J., Zhang, Z., & Zhong, Y. (2017). Local deliv- Torres-Espin, A., Beaudry, E., Fenrich, K., & Fouad, K. (2018). Rehabilitative
ery of thyroid hormone enhances oligodendrogenesis and myelination training in animal models of spinal cord injury. Journal of Neurotrauma,
after spinal cord injury. Journal of Neural Engineering, 14(3), 036014. 35(16), 1970–1985. https://doi.org/10.1089/neu.2018.5906
https://doi.org/10.1088/1741-2552/aa6450 Totoiu, M. O., & Keirstead, H. S. (2005). Spinal cord injury is accompanied
Siebert, J. R., Conta Steencken, A., & Osterhout, D. J. (2014). Chondroitin by chronic progressive demyelination. The Journal of Comparative Neu-
sulfate proteoglycans in the nervous system: Inhibitors to repair. rology, 486(4), 373–383. https://doi.org/10.1002/cne.20517
BioMed Research International, 2014, 845315–845323. https://doi. Trapp, B. D., & Stys, P. K. (2009). Virtual hypoxia and chronic necrosis of
org/10.1155/2014/845323 demyelinated axons in multiple sclerosis. Lancet Neurology, 8(3),
Silver, J., Schwab, M. E., & Popovich, P. G. (2014). Central nervous system 280–291. https://doi.org/10.1016/S1474-4422(09)70043-2
regenerative failure: Role of oligodendrocytes, astrocytes, and Tripathi, R., & McTigue, D. M. (2007). Prominent oligodendrocyte genesis
microglia. Cold Spring Harbor Perspectives in Biology, 7(3), a020602. along the border of spinal contusion lesions. Glia, 55(7), 698–711.
https://doi.org/10.1101/cshperspect.a020602 https://doi.org/10.1002/Glia.20491
Smith, K. J., Blakemore, W. F., & Mcdonald, W. I. (1979). Central Tripathi, R. B., Clarke, L. E., Burzomato, V., Kessaris, N., Anderson, P. N.,
remyelination restores secure conduction. Nature, 280(5721), Attwell, D., & Richardson, W. D. (2011). Dorsally and ventrally derived
395–396. https://doi.org/10.1038/280395a0 oligodendrocytes have similar electrical properties but myelinate pre-
Smith, P. M., & Jeffery, N. D. (2005). Spinal shock––Comparative aspects and ferred tracts. The Journal of Neuroscience, 31(18), 6809–6819. https://
clinical relevance. Journal of Veterinary Internal Medicine, 19(6), 788–793. doi.org/10.1523/JNEUROSCI.6474-10.2011
DUNCAN ET AL. 19

Tsai, H. H., Li, H., Fuentealba, L. C., Molofsky, A. V., Taveira-Marques, R., Witte, M. E., Schumacher, A. M., Mahler, C. F., Bewersdorf, J. P.,
Zhuang, H., … Rowitch, D. H. (2012). Regional astrocyte allocation reg- Lehmitz, J., Scheiter, A., … Kerschensteiner, M. (2019). Calcium influx
ulates CNS synaptogenesis and repair. Science, 337(6092), 358–362. through plasma-membrane nanoruptures drives axon degeneration in
https://doi.org/10.1126/science.1222381 a model of multiple sclerosis. Neuron, 101(4), 615–624 e615. https://
Tsai, H. H., Niu, J., Munji, R., Davalos, D., Chang, J., Zhang, H., … doi.org/10.1016/j.neuron.2018.12.023
Fancy, S. P. (2016). Oligodendrocyte precursors migrate along vascula- Yasuda, A., Tsuji, O., Shibata, S., Nori, S., Takano, M., Kobayashi, Y., …
ture in the developing nervous system. Science, 351(6271), 379–384. Okano, H. (2011). Significance of Remyelination by neural stem/-
https://doi.org/10.1126/science.aad3839 progenitor cells transplanted into the injured spinal cord. Stem Cells,
Ulanska-Poutanen, J., Mieczkowski, J., Zhao, C., Konarzewska, K., Kaza, B., 29(12), 1983–1994. https://doi.org/10.1002/Stem.767
Pohl, H. B., … Zawadzka, M. (2018). Injury-induced perivascular niche Yeung, M. S. Y., Djelloul, M., Steiner, E., Bernard, S., Salehpour, M.,
supports alternative differentiation of adult rodent CNS progenitor Possnert, G., … Frisen, J. (2019). Dynamics of oligodendrocyte genera-
cells. eLife, 7, 7. https://doi.org/10.7554/eLife.30325 tion in multiple sclerosis. Nature, 566, 538–542. https://doi.org/10.
Vallstedt, A., Klos, J. M., & Ericson, J. (2005). Multiple dorsoventral origins 1038/s41586-018-0842-3
of oligodendrocyte generation in the spinal cord and hindbrain. Neu- Young, K. M., Psachoulia, K., Tripathi, R. B., Dunn, S. J., Cossell, L.,
ron, 45(1), 55–67. https://doi.org/10.1016/j.neuron.2004.12.026 Attwell, D., … Richardson, W. D. (2013). Oligodendrocyte dynamics in
van Bruggen, D., Agirre, E., & Castelo-Branco, G. (2017). Single-cell trans- the healthy adult CNS: Evidence for myelin remodeling. Neuron, 77(5),
criptomic analysis of oligodendrocyte lineage cells. Current Opinion in Neu- 873–885. https://doi.org/10.1016/j.neuron.2013.01.006
robiology, 47, 168–175. https://doi.org/10.1016/j.conb.2017.10.005 Yuen, T. J., Silbereis, J. C., Griveau, A., Chang, S. M., Daneman, R.,
Wang, S., Bates, J., Li, X. J., Schanz, S., Chandler-Militello, D., Levine, C., … Fancy, S. P., … Rowitch, D. H. (2014). Oligodendrocyte-encoded HIF
Goldman, S. A. (2013). Human iPSC-derived Oligodendrocyte progeni- function couples postnatal myelination and white matter angiogenesis.
tor cells can Myelinate and rescue a mouse model of congenital Hyp- Cell, 158(2), 383–396. https://doi.org/10.1016/j.cell.2014.04.052
omyelination. Cell Stem Cell, 12(2), 252–264. https://doi.org/10.1016/ Zai, L. J., & Wrathall, J. R. (2005). Cell proliferation and replacement fol-
j.stem.2012.12.002 lowing contusive spinal cord injury. Glia, 50(3), 247–257. https://doi.
Wang, Y., Cheng, X., He, Q., Zheng, Y., Kim, D. H., Whittemore, S. R., & org/10.1002/glia.20176
Cao, Q. L. (2011). Astrocytes from the contused spinal cord inhibit oli- Zawadzka, M., Rivers, L. E., Fancy, S. P., Zhao, C., Tripathi, R., Jamen, F., …
godendrocyte differentiation of adult oligodendrocyte precursor cells Franklin, R. J. (2010). CNS-resident glial progenitor/stem cells produce
by increasing the expression of bone morphogenetic proteins. The Schwann cells as well as oligodendrocytes during repair of CNS demy-
Journal of Neuroscience, 31(16), 6053–6058. https://doi.org/10.1523/ elination. Cell Stem Cell, 6(6), 578–590. https://doi.org/10.1016/j.
JNEUROSCI.5524-09.2011 stem.2010.04.002
Warf, B. C., Fok-Seang, J., & Miller, R. H. (1991). Evidence for the ventral Zhang, Y., Chen, K. N., Sloan, S. A., Bennett, M. L., Scholze, A. R.,
origin of oligodendrocyte precursors in the rat spinal cord. The Journal O'Keeffe, S., … Wu, J. Q. (2014). An RNA-sequencing transcriptome
of Neuroscience, 11(8), 2477–2488. and splicing database of glia, neurons, and vascular cells of the cerebral
Whittaker, M. T., Zai, L. J., Lee, H. J., Pajoohesh-Ganji, A., Wu, J., Sharp, A., cortex. Journal of Neuroscience, 34(36), 11929–11947. https://doi.org/
… Wrathall, J. R. (2012). GGF2 (Nrg1-beta3) treatment enhances NG2 10.1523/Jneurosci.1860-14.2014
+ cell response and improves functional recovery after spinal cord Ziskin, J. L., Nishiyama, A., Rubio, M., Fukaya, M., & Bergles, D. E. (2007).
injury. Glia, 60(2), 281–294. https://doi.org/10.1002/glia.21262 Vesicular release of glutamate from unmyelinated axons in white matter.
Williams, P. R., Marincu, B. N., Sorbara, C. D., Mahler, C. F., Schumacher, A. M., Nature Neuroscience, 10(3), 321–330. https://doi.org/10.1038/nn1854
Griesbeck, O., … Misgeld, T. (2014). A recoverable state of axon injury per-
sists for hours after spinal cord contusion in vivo. Nature Communications,
5, 5683. https://doi.org/10.1038/ncomms6683 How to cite this article: Duncan GJ, Manesh SB, Hilton BJ,
Windrem, M. S., Nunes, M. C., Rashbaum, W. K., Schwartz, T. H.,
Assinck P, Plemel JR, Tetzlaff W. The fate and function of
Goodman, R. A., McKhann, G., 2nd, … Goldman, S. A. (2004). Fetal and
adult human oligodendrocyte progenitor cell isolates myelinate the
oligodendrocyte progenitor cells after traumatic spinal cord
congenitally dysmyelinated brain. Nature Medicine, 10(1), 93–97. injury. Glia. 2019;1–19. https://doi.org/10.1002/glia.23706
https://doi.org/10.1038/nm974

You might also like