ArticlePDF AvailableLiterature Review

Can injured adult CNS axons regenerate by recapitulating development?

Authors:

Abstract

In the adult mammalian central nervous system (CNS), neurons typically fail to regenerate their axons after injury. During development, by contrast, neurons extend axons effectively. A variety of intracellular mechanisms mediate this difference, including changes in gene expression, the ability to form a growth cone, differences in mitochondrial function/axonal transport and the efficacy of synaptic transmission. In turn, these intracellular processes are linked to extracellular differences between the developing and adult CNS. During development, the extracellular environment directs axon growth and circuit formation. In adulthood, by contrast, extracellular factors, such as myelin and the extracellular matrix, restrict axon growth. Here, we discuss whether the reactivation of developmental processes can elicit axon regeneration in the injured CNS.
REVIEW
Can injured adult CNS axons regenerate by recapitulating
development?
Brett J. Hilton*and Frank Bradke*
ABSTRACT
In the adult mammalian central nervous system (CNS), neurons
typically fail to regenerate their axons after injury. During
development, by contrast, neurons extend axons effectively. A
variety of intracellular mechanisms mediate this difference,
including changes in gene expression, the ability to form a growth
cone, differences in mitochondrial function/axonal transport and the
efficacy of synaptic transmission. In turn, these intracellular
processes are linked to extracellular differences between the
developing and adult CNS. During development, the extracellular
environment directs axon growth and circuit formation. In adulthood,
by contrast, extracellular factors, such as myelin and the extracellular
matrix, restrict axon growth. Here, we discuss whether the
reactivation of developmental processes can elicit axon
regeneration in the injured CNS.
KEY WORDS: Axon regeneration, Central nervous system,
Development, Spinal cord, Spinal cord injury
Introduction
Neurons transmit signals to their targets through specialized cellular
processes called axons. These processes extend to other neurons,
muscles, sensory organs and glands, conducting electrical impulses
across long distances. The axons of most neurons grow towards their
targets during embryonic development. The neuron differentiates
and elaborates processes. It specifies one axon and several
dendrites, which are shorter signal-receiving processes (Witte and
Bradke, 2008). The axon elongates towards its target by sensing
guidance molecules (ODonnell et al., 2009). After reaching its
target, the axon forms a specialized connection there called a
synapse (Shen and Scheiffele, 2010). The coordinated growth and
connectivity of axons during development is an impressive feat. In
mammals, billions or trillions of synapses are formed between
neurons and targets that are often long distances away from each
other (Kasthuri et al., 2015; Sporns et al., 2005). As such,
developmental axon growth is an intricate process. It relies on an
axonal environment that enables directed elongation (Seiradake
et al., 2016) and on a variety of intracellular mechanisms that allow
the developing neuron to grow its axon long and rapidly (ODonnell
et al., 2009).
In contrast to the situation in development, and except in a few
unusual cases (Fenrich and Rose, 2009; Jin et al., 2016; Omura
et al., 2015), the axon of an adult mammalian neuron cannot grow
back to its target in the central nervous system (CNS) following
injury; the membrane of the axon reseals to re-establish ionic
homeostasis (Bradke et al., 2012) but axon growth is typically
limited. Most adult neurons fail to express the genes that orchestrate
developmental axon elongation and instead express genes that
restrict growth (He and Jin, 2016). Most adult neurons may also lack
the energy necessary for axon regeneration (Zhou et al., 2016). The
environmental cues that attract the axon to its target during
development are not appropriately expressed in the adult system
(Giger et al., 2010). Instead, the adult environment responds to
injury by expressing factors that impede growth (Cregg et al., 2014;
Schwab and Strittmatter, 2014). In some cases, such as in the
descending corticospinal and rubrospinal systems, and in the
primary sensory neurons that ascend to the brain in the direct dorsal
column pathway, adult neurons survive axonal injury but remain
disconnected from their targets (Kwon et al., 2002; Nielson et al.,
2010; Ylera et al., 2009). In other cases, such as in most subtypes of
retinal ganglion cells (RGCs), injured neurons respond to axon
severance (axotomy) by eventually dying (Duan et al., 2015). Axon
regeneration failure can permanently disrupt CNS connectivity and
can lead to substantial dysfunction in cases of trauma, stroke or
neurodegenerative disease.
A tantalizing strategy for enhancing axon regeneration is to
recapitulate the processes that underlie developmental axon growth
within adult neurons. By examining the processes that mediate axon
outgrowth during development, it may be possible to reactivate the
developmental growth program within the adult neuron and to
unleash robust axon regeneration. Indeed, many experimental
therapeutic strategies to enhance CNS regeneration that
manipulate the extracellular environment or intrinsic growth
capacity of a neuron converge on processes underlying
developmental axon growth (He and Jin, 2016; Schwab and
Strittmatter, 2014). Still, not every process that mediates axon
regeneration is active during embryonic development, and there are
clear differences between these forms of axonal elongation.
In this Review, we highlight the cellular and molecular
mechanisms that orchestrate axon growth during development in
various mammalian species. We then examine how these are
impeded or fail to be activated following adult injury. We describe
divergences between developmental and regenerative axon growth,
and discuss recent work showing that it is possible to promote
regeneration by reactivating developmental growth processes in
adult neurons following CNS injury. Ultimately, a further
understanding of how axons extend and form functional circuits
during development may allow us to recapitulate these processes in
adult neurons and to restore connectivity after disease or injury.
Axon regenerative capacity in the mammalian CNS: a
developmental decline
Although axon regeneration is limited in the adult mammalian CNS,
this is not the case when the CNS is immature. For example,
opossums (such as Monodelphis domestica) are born with an
Laboratory for Axon Growth and Regeneration, German Centre for
Neurodegenerative Diseases (DZNE), Sigmund-Freud-Strasse 27, 53127, Bonn,
Germany.
*Author for correspondence (brett.hilton@dzne.de; frank.bradke@dzne.de)
B.J.H., 0000-0003-2813-2294
3417
© 2017. Published by The Company of Biologists Ltd
|
Development (2017) 144, 3417-3429 doi:10.1242/dev.148312
DEVELOPMENT
immature CNS, including a two-layer cortex and a rudimentary
cerebellum (Kraus and Fadem, 1987). Axons injured early after
birth regenerate and form functional circuits (Saunders et al., 1995).
However, the CNS restricts axon growth after a developmental
transition. For example, retinal ganglion cells (RGCs) in the
opossum regenerate if their axons are injured before postnatal day 12
but not afterwards (MacLaren and Taylor, 1997). Although the
timeframe varies among organisms and pathways, this transition is
virtually ubiquitous in mammals (Nicholls and Saunders, 1996). It
has been referred to as a critical period of CNS axon regeneration,
and understanding its mechanistic basis has been a focus of
regeneration research for many decades.
Classically, it has been thought that extracellular molecules
expressed in adulthood after injury but not during development
underlie this transition. Oligodendrocytes start myelinating axons at
around the same time as the transition and their myelin inhibits axon
growth (Schwab and Strittmatter, 2014). Similarly, astrocytes
respond to developmental and adult lesions differently (Cregg
et al., 2014). However, when neural progenitor cells (NPCs) are
transplanted into the injured adult mammalian spinal cord and
provided with trophic factors and a fibrin matrix, they differentiate
into neurons with a virtually limitless capacity for axon growth (Lu
et al., 2012). These results demonstrate that, when properly
stimulated and in the right environment, newly differentiated
neurons can grow axons that are very long after adult injury. There
are many mechanisms that contribute to the growth capacity of
immature neurons. In the first part of this Review, we focus on
intracellular mechanisms underlying this higher growth capacity
(Fig. 1). In the second part, we discuss how the extracellular
environment changes as the CNS matures in a way that contributes
to regeneration failure. Importantly, intracellular and extracellular
factors are not separate: there is a complex interplay between them,
as the intrinsic growth capacity of a neuron is at least partially
instructed by extracellular interactions (Burnside and Bradbury,
2014; Geoffroy and Zheng, 2014; Tedeschi and Bradke, 2017).
The developing and adult CNS: intracellular factors involved
in regeneration
Gene expression
The difference in gene expression between immature and mature
mammalian neurons is a primary factor that underlies their
differential capacity for axon growth. The capacity of a neuron to
regenerate is defined in part by its expression of regeneration-
associated genes (Tetzlaff et al., 1991). The expression of these
genes is activated in developing neurons and is restricted in adult
neurons (Puttagunta et al., 2014; Tedeschi and Bradke, 2017). For
example, developing mammalian neurons express Krüppel-like
factor 7 (KLF7) and Sox11, but downregulate these transcription
factors as they mature (Blackmore et al., 2012; Moore et al., 2009;
Wang et al., 2015). KLF7 or Sox11 activation in the adult neuron
promotes axon regeneration (Blackmore et al., 2012; Wang et al.,
2015). Similarly, the α9 integrin receptor is downregulated as
primary sensory neurons mature, and the overexpression of α9
promotes axon regeneration (Andrews et al., 2009). In most cases,
transcription factors have been tested singularly for roles in
regeneration. However, bioinformatic approaches have recently
determined functional interactions between pathways involved in
axonal outgrowth and may reveal transcription factor combinations
that can optimize regeneration (Belin et al., 2015; Chandran et al.,
2016; Venkatesh and Blackmore, 2016).
Instead of expressing regeneration-associated genes, mammalian
adult neurons express genes that restrict axon growth. In a screen of
genes differentially expressed by RGCs during the developmental
transition to a restricted capacity to regenerate, Moore and
colleagues identified KLF4 as a transcriptional repressor of
regeneration (Moore et al., 2009). RGCs do not express KLF4
during development but upregulate it upon maturation, when it
restricts growth. Why adult neurons express genes that prohibit
regeneration is unclear. One possibility is that the expression of
these genes is an evolutionarily conserved mechanism to prevent
ectopic axon growth and aberrant synapse formation (Tedeschi and
Bradke, 2017). Indeed, synaptogenesis instructs the expression of
genes that prevent axon outgrowth (Tedeschi et al., 2016), as we
discuss later.
At the translation level, developing neurons may be able to more
effectively generate proteins involved in axonal elongation than
adult neurons (Shigeoka et al., 2016), and signalling via the
mechanistic target of rapamycin (mTOR) pathway may be involved
in this difference (Park et al., 2008). Developing neurons have high
levels of mTOR signalling but this decreases as they mature (Liu
et al., 2010; Park et al., 2008). Enhancing mTOR signalling by
knocking out phosphatase and tensin homolog (Pten) boosts axon
regeneration (Geoffroy et al., 2016; Liu et al., 2010; Park et al.,
2008). The mTOR pathway is pivotal for regeneration following
Pten knockout (Park et al., 2008). Although mTOR regulates many
intracellular events (Laplante and Sabatini, 2012), its activation of
protein translation is particularly important for regeneration (Yang
et al., 2014). To stimulate translation, mTOR activates S6 kinase 1
(S6K1) to generate new ribosomes (Chauvin et al., 2014), and S6K1
activation enhances axon regeneration in the mouse optic nerve
(Yang et al., 2014). In addition to S6K1, mTOR activates eukaryotic
initiation factor 4E (eIF4E) to initiate cap-dependent protein
translation (Brunn et al., 1997). Accordingly, it has been shown
that EIF4E is necessary for regeneration following Pten knockout in
adult mice (Yang et al., 2014). Together, these results support a
model in which developing neurons signal through mTOR to
translate proteins more effectively than adult neurons do, and
suggest that enhancing mTOR-mediated protein translation in
adulthood can boost axon regeneration. However, it has been
reported that S6K1 restricts mouse corticospinal axon regeneration
(Al-Ali et al., 2017). These results highlight the complexity of
mTOR signalling in regeneration and suggest that more research is
needed to understand how different effectors of the PI3K/Akt/
mTOR pathway influence axon growth.
Growth cone formation
Developmental axon growth relies on the formation of a subcellular
structure at the distal tip of the axon called the growth cone
(Fig. 2A). The growth cone comprises an actin-rich peripheral
domain and a microtubule-rich central domain, separated by a
transition zone in which dynamic interactions between these
cytoskeletal elements take place (Coles and Bradke, 2015). It
serves vital functions both in the guidance of the axon to its target
and in the process of axonal outgrowth itself (Dent et al., 2011). At
its distal tips, finger-like filopodia and sheet-like lamellipodia
rapidly extend and retract as they sense the axonal
microenvironment. During outgrowth, filamentous actin (F-actin)
polymerizes, driving the extension of new membrane at the distal
edges of the growth cone (Goldberg and Burmeister, 1986). The
proximal growth cone then stabilizes into the axonal shaft,
elongating the axon (Dent et al., 2011; Goldberg and Burmeister,
1986). After synaptogenesis, axon growth continues to occur,
corresponding to the size of the growing organism, but it does so via
stretch-based mechanisms (Smith, 2009).
3418
REVIEW Development (2017) 144, 3417-3429 doi:10.1242/dev.148312
DEVELOPMENT
After injury, the formation of a growth cone or growth cone-like
structure is vital for axon regeneration (Bradke et al., 2012). Injury
exposes the interior of the axon to calcium, triggering membrane
sealing and growth cone formation. Peripheral nervous system
(PNS) axons form new growth cones after injury in order to
regenerate (Ertürk et al., 2007). However, CNS axons fail to form
growth cones (Bradke et al., 2012; Ertürk et al., 2007). Instead, they
are tipped with dystrophic retraction bulbs; these structures
accumulate anterogradely transported vesicles and mitochondria,
and show microtubule disassembly (Fig. 2A-B). In the human
spinal cord, retraction bulbs have been observed more than four
decades after injury (Ruschel et al., 2015), highlighting their role in
regeneration failure.
Growth cones contain organized microtubules that form tight
bundles parallel to the axonal axis, whereas retraction bulbs have
highly dispersed and disorganized microtubules (Ertürk et al., 2007).
Application of the microtubule-destabilizing agent nocodazole
transforms growth cones into retraction bulb-like structures in vivo,
resulting in halted axon growth (Ertürk et al., 2007). Conversely,
microtubule stabilization through the administration of taxol
interferes with retraction bulb formation and facilitates growth
cone formation in vivo (Fig. 2B,C). When applied in vivo after
CNS injury, taxol administration boosts optic nerve axon regeneration
(Sengottuvel et al., 2011) and serotonergic axon regeneration (Hellal
et al., 2011). Microtubule stabilization also promotes locomotor
recovery in rodent models of traumatic brain injury (Cross et al.,
2015) and spinal cord injury (Hellal et al., 2011). Since microtubule
stabilization drives axon outgrowth during development (Gomis-
Rüth et al., 2008; Witte et al., 2008), these studies highlight how
regeneration can be achieved by activating developmental processes.
Importantly, microtubule stabilization can be stimulated
noninvasively through systemic administration of epothilone B
(Ruschel et al., 2015). This drug binds to the αβ-tubulin heterodimer
subunit of microtubules, decreasing its rate of dissociation and
thus stabilizing the microtubule (Goodin et al., 2004). In vitro,
epothilone B encourages growth cone formation (Ruschel et al.,
2015), while in the adult rat, epothilone B administration
improves skilled locomotion and enhances serotonergic axon
growth following spinal cord injury (Ruschel et al., 2015). These
functional improvements are abrogated by pharmacological ablation
of serotonergic innervation, suggesting that the serotonergic axon
growth is necessary for recovery. In addition to influencing
growth cone formation, microtubule stabilization reduces fibrotic
scarring by mitigating fibroblast polarization and migration
(Hellal et al., 2011; Ruschel et al., 2015) (Fig. 2C,D). Thus,
microtubule stabilization may be a viable strategy for promoting
axon regeneration and functional improvements following CNS
injury.
KLF7
Sox11
KLF4
DNA
Regeneration-associated
gene
Growth-restricting gene
Motor
protein Mitochondrion
Depolarized
mitochondrion
Syntaphilin
Miro/milton complex
ATP
ADP
B Injured adult neuron failing to regenerate axon
A Developing neuron with growing axon
Microtubule
Retraction
bulb
Growth
cone
Nucleus
Nucleus
Key
Fig. 1. Intracellular processes in a developing neuron and its growing axon, and in an adult neuron failing to regenerate. (A) A developing neuron with a
growingaxon. Developing neurons express regeneration-associated genes, such as Klf7 and Sox11, that facilitate axon growth.Mitochondria are highly motile in the
axon, are associatedwith Miro/milton complexes and aretransported by motor proteins (kinesinand dynein) along microtubules. ATPis at a level sufficient to supply
the energy requiredfor axon growth. The axon has a growth cone at its tip (purple) that senses guidance molecules(not shown) and mediates axonal outgrowth. (B)
A schematic of an injured adult neuron failing to regenerate. Adult injured neurons express genes that restrict axon outgrowth, such as Klf4. Mitochondria
are depolarized after injury and are less motile in the adult axon due to the expression of the anchor protein syntaphilin. ATP levels are lower and ADP levels are
higher, such that the axon does not have enough energy to grow. The axon has a retraction bulb with disorganized and disassembled microtubules at its tip.
3419
REVIEW Development (2017) 144, 3417-3429 doi:10.1242/dev.148312
DEVELOPMENT
Actin dynamics are also key to growth cone formation. Following
axotomy, F-actin destabilizes at the injured stump, contributing to
retraction bulb formation (Nawabi et al., 2015). Overexpression of
doublecortin like kinase 2 (DCLK2) can prevent this destabilization
and promotes both growth cone formation and axon regeneration
(Nawabi et al., 2015). DCLK2 is enriched in the transition zone of the
growth cone, at the interface between the actin-rich peripheral
domain and microtubule-rich central domain, where it is thought to
help coordinate microtubule/actin dynamics (Bielas et al., 2007).
Interestingly, although overexpression of either the microtubule-
binding domain or actin-regulatory domain of DLK2 promotes axon
regeneration, simultaneous overexpression of both does not (Nawabi
et al., 2015). These results highlight the complexity of molecular
interactions within the distal tip of the axon following injury.
Although DCLK2 encourages growth cone formation by stabilizing
F-actin, some actin destabilization may be beneficial as it permits
microtubules to protrude into distal areas of the growth cone as the
axon shaft is consolidated (Bradke and Dotti, 1999; Flynn et al.,
2012). Understanding the role of actin in different compartments of
the growth cone is thus essential for understanding how it influences
axon regeneration. Moreover, how the neuron coordinates actin/
microtubule dynamics to effectively elongate its axon remains an
important avenue for further research (Coles and Bradke, 2015).
Mitochondrial function and axonal transport
Axon growth is an energetically intensive process that requires
mitochondrial biogenesis and adenosine triphosphate (ATP)
production (Vaarmann et al., 2016). Although ATP diffuses freely
in the cytosol, neurons, like many other cells, traffic mitochondria to
regions of high ATP consumption (Schwarz, 2013). During axon
specification and outgrowth in development, the growth cone
contains a particularly high concentration of mitochondria (Bradke
and Dotti, 1997; Morris and Hollenbeck, 1993). Axons transport
mitochondria both in the anterograde and retrograde direction along
Fibroblasts
Fibrotic
scar
A
F-actin
bundles
Filopodium
Lamellipodium
Actin
meshwork
Developmental growth cone Retraction bulb
Disorganized and
disassembled
microtubules
Vesicle
Mitochondrion
Accumulation
of vesicles
C Spinal cord injury with microtubule stabilization (epothilone B or taxol treatment)
B Spinal cord injury
Organelle
components
Cytoskeletal
structure
Microtubules
Fig. 2. The growth cone orchestrates
developmental axon growth and is a
target for CNS regeneration. (A)
Cytoskeletal structure and organelle
components of a developmental growth
cone and a retraction bulb. In the growth
cone, microtubules are oriented with the
axonal axis. The distal aspect of the growth
cone is actin rich (purple) and contains
F-actin bundles, filopodia and lamellipodia.
In the retraction bulb, microtubules are
disorganized. Its actin structure is poorly
understood. Mitochondria and vesicles are
found at a higher concentration in the
growth cone than in the axon shaft. In the
retraction bulb, vesicles and mitochondria
accumulate distally. (B) After spinal cord
injury, sensory axons have retraction bulbs
at their tips and axon growth is restricted by
fibroblasts that form scar tissue at the injury
site. (C) Microtubule stabilization (using
epothilone B or taxol treatment) mitigates
fibrotic scarring and enhances the formation
of growth-cone-like structures at the distal
tips of injured sensory axons, enabling
regeneration.
3420
REVIEW Development (2017) 144, 3417-3429 doi:10.1242/dev.148312
DEVELOPMENT
microtubules by using a motor/adaptor complex that contains
kinesin and dynein in addition to the proteins Miro (RhoT1/2) and
milton (Trak1/2) (Schwarz, 2013). The motor adaptor Trak1 is
required for axonal mitochondria transport (van Spronsen et al.,
2013) and its depletion inhibits axon outgrowth (van Spronsen et al.,
2013). Thus, mitochondria and their axonal transport are essential
for sustained axon growth during development.
Axotomy depolarizes mitochondria and depletes ATP in injured
axons (Cavallucci et al., 2014; ODonnell et al., 2013; Zhou et al.,
2016). When axons do not have mitochondria, they rapidly
degenerate (Rawson et al., 2014). Conversely, boosting
mitochondrial ATP production facilitates axon regeneration.
Cytokines stimulate axon regeneration in part by inducing signal
transducer and activator of transcription 3 (STAT3) to translocate
to the inner membrane of the mitochondria (Luo et al., 2016).
There, STAT3 enhances ATP production by optimizing the
function of the electron transport chain (Luo et al., 2016;
Wegrzyn et al., 2009). In cytochrome c oxidase-deficient mice,
which have defective mitochondrial ATP production, cytokine-
stimulated RGC regeneration is reduced (Luo et al., 2016). Thus,
although the cell can produce ATP via glycolysis in the cytoplasm,
ATP generated by the electron transport chain in mitochondria
is important for axon regeneration, and the efficiency of this
process is a target for repair.
Enhancing axonal transport of mitochondria also promotes
regeneration. Adult mammalian neurons transport mitochondria in
their axons less than do developing neurons because they express
syntaphilin (SNPH), a protein that anchors axonal mitochondria to
the axon (Kang et al., 2008). In adult mouse RGCs that have a high
regenerative ability due to dual genetic knockout of Pten and
suppressor of cytokine signalling 3 (Socs3), armadillo repeat
containing X-linked 1 (Armcx1) is highly expressed (Sun et al.,
2011). Armcx1 encodes a mitochondrial protein and belongs to a
family of genes unique to placental mammals (López-Doménech
et al., 2012). The protein contains a putative outer mitochondrial
membrane-targeting sequence, flanking a transmembrane domain
necessary for its mitochondrial localization (Cartoni et al., 2016).
Overexpression of Armcx1 in adult mouse increases mitochondrial
motility and enhances RGC survival and axon regeneration (Cartoni
et al., 2016). In addition, upon simultaneous Pten and Socs3
knockout, which greatly enhances neuronal intrinsic growth
capacity (Sun et al., 2011), Armcx1 mediates neuron survival and
axon regeneration (Cartoni et al., 2016). Similarly, a high
mitochondrial density is crucial for axon regeneration in the
nematode, C. elegans (Han et al., 2016). However, it is unclear how
the axonal transport of mitochondria influences regeneration. While
enhancing ATP production in the axon is likely to be one factor,
mitochondrial density might also influence regeneration by altering
signalling through changes in metabolite production or intracellular
calcium levels (Chandel, 2014; Williams et al., 2013). It is also
known that mitochondria can regulate axon branching by
determining sites of axonal protein synthesis (Spillane et al.,
2013) and that mitochondrial deficiency can activate signalling
cascades that result in dendritic branching (Gioran et al., 2014). The
relationship between the mitochondria in the signal transduction
cascades that mediate axonal outgrowth is an important issue for
future research.
Synaptic transmission
During embryonic mammalian development, axons are connected
to far more targets than those in the adult (Luo and OLeary, 2005).
This exuberance is mediated in part by neurons extending linearly
and rapidly past their target cells (Stanfield et al., 1982). Following
this, side branches form on the elongated axon and grow slowly into
the target area. The overshooting axon degenerates, and
terminations are refined during a protracted period of
synaptogenesis and synapse elimination (Hua and Smith, 2004;
Low and Cheng, 2006; Purves and Lichtman, 1980). Together,
exuberant outgrowth and pruning during development generate
precise connectivity in the adult (Luo and OLeary, 2005).
After adult connectivity is established, neurons employ multiple
strategies to avoid ectopic axon growth and to maintain appropriate
synaptic connectivity (Shen and Scheiffele, 2010). These same
strategies might restrict axon regeneration and rely on synaptic
transmission. In an elegant series of experiments, Lorenzana and
colleagues used in vivo two-photon imaging of adult mouse spinal
dorsal column sensory axons to assess whether a surviving intact
branch influences the ability of the neuron to regenerate (Lorenzana
et al., 2015). These axons present a particularly good model for this
because they bifurcate into an ascending branch, which extends
rostrally towards the brain, and a descending branch, which courses
down the spinal cord. Eliminating either of these branches is
followed by a poor regenerative response but eliminating both
increases regeneration. Thus, an intact axonal process suppresses
regeneration. Although the neurons projecting these central axonal
processes do not effectively regenerate them, they can regenerate
their axonal processes in the PNS. One signal underlying this
difference is electrical activity (Enes et al., 2010). After peripheral
nerve injury, electrical activity is lost, which may signal regeneration
to initiate. In contrast, electrical activity is maintained after central
axon injury and may suppress growth by triggering an increase in
intracellular calcium concentration (Enes et al., 2010). Thus, an
intact axonal process can discourage regeneration and may do so by
maintaining electrical activity. This exciting possibility needs to be
directly tested.
The synaptic-based suppression of axon growth is something that
can be targeted therapeutically. Cacna2d2, the gene encoding the
α2δ2 subunit of voltage-gated calcium channels, limits sensory
axon regeneration (Tedeschi et al., 2016). α2δ2 subunits promote
synapse formation and enhance the probability of synaptic
transmission (Eroglu et al., 2009). Thus, they may act as a
molecular switch to synaptically suppress axonal outgrowth.
Gabapentinoids, including pregabalin and gabapentin, are
clinically approved drugs used to treat epilepsy, neuropathic pain
and fibromyalgia, which bind with high affinity and selectivity to
α2δ1/2 subunits (Gee et al., 1996). Systemic gabapentinoid
administration induces adult mouse sensory axon regeneration
following spinal cord injury (Tedeschi et al., 2016). A meta-analysis
found that spinal cord-injured individuals that received these drugs
have enhanced motor recovery if they are administered in the first
month after injury (Warner et al., 2017). Together, these promising
results pave the way for a detailed exploration of gabapentinoid use
to treat spinal cord injury and other neurological disorders
characterized by paralysis.
In summary, intracellular processes define the capacity of
developing neurons, and the inability of adult neurons, to
regenerate. In turn, these processes occur together with changes in
the extracellular environment of the neuron that are also instructive
to axon growth and restraint.
Developing and adult CNS: extracellular factors involved in
regeneration
In addition to intracellular differences between developing and adult
neurons, the extracellular environment of the neuron changes as the
3421
REVIEW Development (2017) 144, 3417-3429 doi:10.1242/dev.148312
DEVELOPMENT
CNS matures in a way that restricts regeneration. However, it is
becoming clear that extracellular factors that guide developmental
axon growth can facilitate regeneration in the adult.
Axon guidance cues
As an axon extends during development, its growth cone senses an
array of guidance molecules expressed by the environment that
allow it to steer towards the correct target. Several phylogenetically
conserved families of guidance cues have been discovered,
including netrins, slits, semaphorins and ephrins (ODonnell
et al., 2009). The growth cone integrates these signals and
transmits them into cytoskeletal changes that underlie steering
decisions, such as attraction and repulsion (ODonnell et al., 2009).
Some guidance molecules that function primarily in an inhibitory
or repulsive role developmentally are expressed following adult
injury and inhibit axon growth (Giger et al., 2010). For example,
during development, Wnt proteins are expressed in spinal cord grey
matter along the rostral/caudal axis in a concentration gradient that
decreases caudally (Liu et al., 2005). Corticospinal axons express
the repulsive Wnt receptor Ryk (related to receptor tyrosine kinase),
which repels these axons to steer them away from the brain and
towards their spinal cord targets (Li et al., 2009; Liu et al., 2005)
(Fig. 3A,B). Wnt proteins are virtually undetectable in the adult
spinal cord (Fig. 3C), but their expression is induced around the
injury site following spinal cord injury, where they inhibit axon
growth (Liu et al., 2008) (Fig. 3D). Reducing the expression or
axonal detection of repulsive guidance cues can thus enhance axon
growth. Blocking Wnt/Ryk signalling enhances corticospinal axon
sprouting in adult mice after cervical spinal cord injury (Hollis et al.,
2016; Liu et al., 2008) (Fig. 3E). The corticospinal tract mediates
skilled forelimb motor control in rodents and primates (Lawrence
and Kuypers, 1968; Starkey et al., 2005). Interestingly, although
blocking Wnt-Ryk signalling can promote corticospinal sprouting,
this is not sufficient to promote forelimb motor recovery after spinal
cord injury (Hollis et al., 2016). Instead, recovery requires the
combination of this blockade with a form of rehabilitative, task-
specific training. If mice are coaxed to perform repeated reaching
movements, Ryk inhibition promotes extensive forelimb motor
recovery. Importantly, this recovery is associated with motor
cortical reorganization. Motor cortical areas that drive hindlimb
motor responses prior to injury acquire the capacity to activate
forelimb muscle groups controlled by motor neurons located rostral
to injury in Ryk-inhibited, rehabilitative trained mice. This provides
strong evidence that releasing corticospinal axons from Wnt
inhibition promotes circuit plasticity when combined with
activity-based rehabilitative approaches (Dietz and Fouad, 2014).
More generally, these results highlight the importance of
developing approaches that direct growing axons into forming
functional circuits and of strengthening connections.
In addition to blocking the action of repulsive or inhibitory
guidance cues, an alternative strategy to promote regeneration is to
provide injured axons with attractive signals that they receive
developmentally (Fig. 4). Grafting NPCs into sites of spinal cord
injury in the adult rat promotes robust corticospinal axon
regeneration via contact-based mechanisms (Kadoya et al., 2016).
However, NPCs derived from the rat spinal cord facilitate greater
regeneration than do NPCs derived from the rat brain. Moreover,
NPCs driven toward a spinal cord-like fate via treatment with
retinoic acid, which caudalizes the cells in a concentration-
dependent manner (Okada et al., 2004), also facilitate more
corticospinal axon regeneration than do brain-derived NPCs.
Interestingly, the maturity of rat NPCs following their engraftment
is not a primary factor in the regenerative response: when NPCs are
matured for 70 days prior to axonal injury, they support regeneration
as well as do immature NPCs grafted the same day as the injury
(Kadoya et al., 2016). These data support a model in which the
regional similarity of transplanted cells to the original substrate of an
axon is an essential factor in determining their capacity to support
regeneration. Understanding the molecular factors underlying this
effect may allow us to provide these signals to axons without the
need for cell transplantation (Assinck et al., 2017).
Astrocytes and the lesion site scar
Cells of the astrocyte lineage play a vital role in axon pathfinding
and circuit formation during development (Clarke and Barres, 2013;
Fitch and Silver, 1997). They secrete extracellular matrix (ECM)
Wnt
Rostral
Caudal
Wnt
Development
(P 1) Adult Adult
injury
Adult
injury
Ryk KO
Wnt
GM WM GM WM GM WM GM WM
ABCD
Ryk
Frizzled
Wnt
Ca2+
Ca2+
Growth
cone
repulsion
TRP
channel
E
(Injury site) (Injury
site)
Fig. 3. Wnt signalling directs corticospinal axon guidance during development but restricts corticospinal sprouting after injury. (A) A Wnt/Ryk
signalling pathway underlies corticospinal axon repulsion (Li et al., 2009). Wnt (blue) binds to the Ryk (orange) and Frizzled (black) receptors, leading to
activation of TRP channels (green). Calcium influx through TRP channels leads to growth cone repulsion. (B) During development, pioneer corticospinal
axons (red) steer away from the brain towards their spinal targets in white matter (WM) by sensing Wnt, a repulsive guidance cue expressed in grey matter (GM)
that instructs the axons to grow caudally. (C) In adulthood, the corticospinal tract is fully mature, with axonal sprouts extending into the GM, and Wnt is not
expressed. (D) After spinal cord injury, corticospinal axons are inhibited from sprouting by Wnt, which is expressed atthe injury site. (E) Knocking out the repulsive
Wnt receptor Ryk enables corticospinal axon sprouting around a site of spinal cord injury.
3422
REVIEW Development (2017) 144, 3417-3429 doi:10.1242/dev.148312
DEVELOPMENT
molecules, such as chondroitin sulphate proteoglycans (CSPGs),
that form repulsive barriers to developmental axon growth (Brittis
et al., 1992). After adult injury, astrocytes become reactive and
direct the formation of a scar at the injury site by walling off a core of
cells, including fibroblasts, macrophages, NG2 glia, pericytes and
ependymal cells (Cregg et al., 2014). This response to injury serves
an important purpose: in directing scar formation, astrocytes
mitigate inflammatory damage and are thus neuroprotective
(Faulkner et al., 2004; Hilton et al., 2016b). However, many scar
cells inhibit axon growth, typically by physically interacting with
the distal tips of axons (Filous et al., 2014) or by secreting ECM
molecules such as CSPGs (Burnside and Bradbury, 2014; Cregg
et al., 2014; Tan et al., 2011). When single mouse sensory axons are
axotomized by a laser with minimal scarring, the axons regenerate
robustly a few days after receiving a peripheral conditioning lesion
that boosts their intrinsic growth potential (Ylera et al., 2009). In
contrast, sensory neurons that are growth competent due to a
peripheral conditioning lesion cannot regenerate axons through
mature scar territory (Ylera et al., 2009).
Therapeutic strategies that target these extracellular impediments
to axon regeneration can reactivate developmental growth
processes. For example, during development, visual circuits are
plastic during a well-defined critical period that ends as the ECM
matures (Hubel and Wiesel, 1970; Pizzorusso et al., 2002).
Digesting CSPGs via treatment with chondroitinase-ABC
reactivates this developmental plasticity (Pizzorusso et al., 2002),
while chondroitinase-ABC treatment enhances axon growth in part
by facilitating expression of the same regeneration-associated genes
that mediate developmental outgrowth (Bradbury et al., 2002).
Interestingly, chondroitinase-ABC decreases the amplitude and
charge of excitatory postsynaptic currents in vitro, suggesting that
its influence on axon growth may relate to its modulation of synaptic
activity (Pyka et al., 2011). Of note, treatment with the microtubule-
stabilizing drugs taxol and epothilone B reduce CSPG deposition
and fibrotic scarring after spinal cord injury (Hellal et al., 2011;
Ruschel et al., 2015).
Immature astrocytes can support the regeneration of adult central
axons (Filous et al., 2010; Reier et al., 1986; Smith et al., 1986) and
synthesize less CSPGs than adult astrocytes do (Dow et al., 1994).
This is similar to glial cells in lower vertebrate species, which form
multicellular structures (bridges) that provide a substrate onto
which axons regenerate and form new connections after injury
(Butler and Ward, 1967; Zukor et al., 2011). In zebrafish, this
bridging is directed by connective tissue growth factor a (CTGFa)
signalling (Mokalled et al., 2016). Interestingly, some adult
astrocytes retain a capacity to form bridges that facilitate
regeneration following Pten knockout in adult mouse (Liu et al.,
2010). It is tempting to speculate that axon growth could be
promoted by manipulating astrocytes into forming regenerative
bridges. However, it would be important to promote bridge
formation in a way that does not stop astrocytes from preventing
inflammatory cell loss.
Recent work has suggested that astrocytic scar formation aids
CNS axon regeneration (Anderson et al., 2016). Sensory axons
regenerate after spinal cord injury when they are administered a
peripheral conditioning lesion, trophic factors and hydrogels.
However, axons do not regenerate after this treatment if lesion site
astrocytes are depleted by selectively killing proliferating astrocytes
or preventing astrocytic STAT3 signalling (Anderson et al., 2016).
As reactive astrocytes sequester inflammatory cells to protect
nervous tissue following injury (Faulkner et al., 2004), their ablation
likely enhances inflammatory cell damage (Sofroniew, 2015),
Rostral
Caudal
RA
Telencephalon
NPCs
Hindbrain
NPCs
Spinal cord
NPCs
Caudalized
NPCs
Human
iPSCs
Rostralized
NPCs
Neural
induction
medium
Neural
induction
medium
+ RA
CST
SCI and
NPC graft
CST
SCI (no graft)
Rostral Caudal
A E14 rat CNS
B Adult rat CNS
C NPC culture
Telencephalon
Hindbrain
Spinal cord
Fig. 4. The capability of neural progenitor cells to support regeneration is influenced by their site of origin. (A) A schematic of an embryonic day
(E) 14 rat embryo showing itsdeveloping CNS. During embryonic development, the rostrocaudal axis of the CNS is specified by morphogens such as retinoic acid
(RA), which caudalizes tissue in a concentration-dependent manner (green). This process distinguishes the spinal cord (green) from the more rostral
hindbrain (khaki) and telencephalon (orange). (B) In the adult, axons in the corticospinal tract (CST) project from the cortex to the spinal cord and fail to regenerate
after spinal cord injury (SCI). (C) Neural progenitor cells (NPCs) can be dissociated and cultured from an E14 rat CNS, embedded in a fibrin matrix with a cocktail
of trophic factors, and grafted into sites of spinal cord injury. NPCs derived from the telencephalon do not support CST regeneration into the graft (orange),
whereas spinal cord graft NPCs do (green). Hindbrain-derived NPCs (khaki) support some CST regeneration (but less than that produced byspinal-cord NPCs).
In addition, human induced-pluripotent stem cells (iPSCs) driven towards a rostralized NPC fate (orange) do not support CST regeneration but iPSCs caudalized
by treatment with RA (green) do.
3423
REVIEW Development (2017) 144, 3417-3429 doi:10.1242/dev.148312
DEVELOPMENT
precluding the possibility of regeneration (Silver, 2016). Indeed,
astrocytic scar formation is driven by type 1 collagen, which
instructs astrocytes to adhere to one another by activating the
integrin/N-Cadherin signalling pathway (Hara et al., 2017).
Blockade of collagen 1 signalling in reactive astrocytes, which
prevents scar formation without depleting reactive astrocytes from
the lesion site, enhances axon regeneration (Hara et al., 2017). Thus,
although astrocyte depletion does not promote regeneration, some
astrocyte phenotypes produced in response to spinal cord injury
inhibit axon growth.
Intriguingly, recent work in mouse and rat has highlighted the
heterogeneity of reactive astrocytes following CNS disease and
injury by providing evidence that classically activated microglia
induce astrocytes to become neurotoxic (Liddelow et al., 2017).
Microglia secrete interleukin 1α(IL1α), tumour necrosis factor
(TNF) and complement component 1, subcomponent q (C1q),
inducing the formation of neurotoxic A1astrocytes that kill RGCs
after axotomy (Liddelow et al., 2017). In contrast, trophic-factor
producing A2astrocytes form in response to ischemia and may be
neuroprotective. The relationship of this type of astrocyte
heterogeneity to the bridge-forming or scar-forming
characteristics of some astrocytes is unknown. Nonetheless,
blocking the neurotoxic activity of A1 reactive astrocytes or
directing A2 astrocyte formation may promote neuron survival and
regeneration following disease or injury.
Oligodendrocytes and myelin
Myelin insulates axons and accelerates the conduction of electrical
impulses (Fancy et al., 2011). Most myelin-forming
oligodendrocytes are born in the early postnatal period at
approximately the same time as the developmental transition
occurs that restricts axon regeneration (Fancy et al., 2011;
Keirstead et al., 1992). As such, they have long been considered
potential mediators of this transition (Schwab et al., 1993).
Suppressing the onset of myelination can extend the permissive
period of functional regeneration in the chick (Keirstead et al.,
1992). Various myelin-associated molecules inhibit axon
outgrowth, including Nogo, myelin-associated glycoprotein
(MAG) and oligodendrocyte-myelin glycoprotein (OMgp)
(Schwab and Strittmatter, 2014). Nogo blockade boosts central
axon growth and circuit plasticity following a stroke in an adult rat if
followed by rehabilitation training to consolidate newly grown
axons into mature circuits (Wahl et al., 2014).
Although oligodendrocytes restrict axon growth, they may play a
vital role after injury by myelinating newly regenerated axons
(Fig. 5). If the axon has regenerated, it may require new myelin for
efficient conduction and to form a functional circuit, particularly if
that axon was well myelinated prior to injury. In adult mice,
knockout of Pten and Socs3 or overexpression of osteopontin
(OPN), insulin-like growth factor 1 (IGF1) and ciliary neurotrophic
factor (CNTF) induces RGC regeneration (Bei et al., 2016). After
regenerating, some axons form functional synapses with their
presumptive targets in the superior colliculus. However, these axons
lack myelin, fail to conduct action potentials and do not restore
visual function. Administration of the voltage-gated potassium
channel blocker 4-aminopyridine (4-AP) improves visual function
by promoting conduction in regenerated axons (Bei et al., 2016).
Neurotrophin 3 expression guides sensory axon regeneration and
synapse formation but the axons remain unmyelinated (Alto et al.,
2009). In contrast, after Pten knockout, cAMP administration and
inflammatory stimulation, regenerating RGC axons are myelinated
and assemble nodes of Ranvier (Marin et al., 2016). After NPCs are
transplanted into the injured spinal cord and differentiate into
neurons that extend axons long distances, about a quarter of their
axons are myelinated by host oligodendrocytes (Hunt et al., 2017).
As such, myelination of newly regenerated axons is not an automatic
process and varies based on the treatment that induces regeneration to
begin with. Similar to analyses of axon regeneration, assessment of
myelination in regeneration studies must be carefully interpreted
because these results can be confounded by the presence of myelin on
uninjured axons (Tuszynski and Steward, 2012).
The signals that direct myelination of regenerating axons are
unclear. Following focal chemical demyelination, oligodendrocytes
effectively regenerate myelin on denuded axons in a process that
recapitulates aspects of developmental myelination (Fancy et al.,
2011). Both oligodendrocytes and Schwann cells are capable of
regenerating myelin (Plemel et al., 2017), with oligodendrocyte
remyelination relying on myelin regulatory factor signalling
(Duncan et al., 2017) and Schwann cell remyelination requiring
neuregulin 1 signalling (Bartus et al., 2016). One possibility is that
denuded axons contain signals necessary for myelination that
regenerated axons may lack. In this regard, one potential signal is
electrical activity (Fields, 2015). Another possibility is that
myelination of regenerated axons requires inflammatory
stimulation, given that macrophages drive oligodendrocyte
differentiation during remyelination of denuded axons (Miron
et al., 2013). Understanding the signals that regulate the myelination
of regenerated axons is a new area of research that will yield
valuable insight into developing effective repair strategies.
Do adult axons continue to diminish in their regenerative
capacity with age?
Although the CNS diminishes in regenerative capacity early after
birth, one issue of substantial therapeutic relevance is whether it
continues to decline as it ages. In other words, do mid-adult-aged
neurons have even less capacity to regenerate relative to their young
adult counterparts? Age is an important risk factor in cases of stroke
and glaucoma (Leske et al., 1995; Sacco et al., 1997), both of which
would benefit from axon regeneration therapies. Additionally, the
average age of incidence of spinal cord injury has increased since
the 1970s and now occurs in the late 30s or early 40s (DeVivo and
Chen, 2011). However, assessing the role of aging in axon
regeneration has been a challenge because young adult mammals
already have an extremely low capacity to regenerate to begin with.
A number of key insights into neuronal aging have come from
studies of C. elegans. After C. elegans reach adulthood, aging
diminishes their ability to generate growth cones and to extend axons
after injury (Byrne et al., 2014). This restricted axon regeneration is
not merely a consequence of organismal aging. Instead, it is an active
intracellular process: in aged C. elegans, insulin signalling via the
insulin growth factor 1 (IGF1) receptor inhibits forkhead
transcription factor daf-16/FOXO activity, suppressing
regeneration (Byrne et al., 2014). In the mammalian peripheral
nervous system, axon regenerative capacity declines with age
because Schwann cells are impaired in their ability to dedifferentiate
and clear myelin debris (Painter et al., 2014). Mammalian central
neurons also decline in their regenerative capacity as they age from
young adulthood to mid-adulthood (Geoffroy et al., 2016). After
Pten knockout in 4-week-old mice, corticospinal and rubrospinal
neurons regenerate axons past sites of spinal cord injury (Geoffroy
et al., 2016; Liu et al., 2010). However, when Pten is knocked out at
older ages (10 weeks or older), axons fail to navigate past the injury
site by 8 weeks post-injury (Geoffroy et al., 2016). Importantly, Pten
knockout boosts axon sprouting above the lesion in older animals,
3424
REVIEW Development (2017) 144, 3417-3429 doi:10.1242/dev.148312
DEVELOPMENT
demonstrating that it remains effective in enhancing axon growth.
Pten restricts axon regeneration in C. elegans independently of age
(Byrne et al., 2014). Hence, Pten is likely an evolutionarily
conserved inhibitor of regeneration. After knockout of the myelin
associated inhibitor Nogo, the regeneration of corticospinal and
serotonergic axons is much less after spinal cord injury in 14-week-
old mice relative to 8-week-old mice (Cafferty et al., 2007).
How mammalian neurons continue to decline in their
regenerative capacity as they age is unknown. As in C. elegans,
specific signalling pathways might be activated in older mammalian
neurons that further restrict their regeneration, but this is unclear.
Another possibility is that older neurons have the capacity to
regenerate following treatment but their growth through scar
territory is much slower. When Pten is knocked out in
corticospinal neurons a full year after spinal cord injury, axon
regeneration is observed 7, but not 4, months later (Du et al., 2015).
This is a much longer timeframe for regeneration than in studies in
young adult mice, where knockout is induced prior to injury
(Geoffroy et al., 2016; Liu et al., 2010). In those cases, axon
regeneration is observed 6-12 weeks post-injury. One major
difference between these scenarios is that the scar at the injury
site is significantly denser and more mature at 1 year post-injury
(Cregg et al., 2014). When the mid-adult mouse spinal cord is
injured, it has enhanced macrophage density and astrogliosis
relative to an injured spinal cord from a young adult (Geoffroy
et al., 2016). Moreover, in the adult rat, aging increases the density
of CSPGs sulphated at position 6, which cause more inhibition of
axon growth, and decreases CSPGs sulphated at position 4, which
are more growth permissive (Foscarin et al., 2017). As such, aging
might accelerate the formation or maturity of inhibitory scar
territory, making it harder for axons to regenerate. Understanding
the mechanistic basis of this decline in central axon regeneration
from young adulthood to mid-adulthood is an exciting avenue of
research with important implications for CNS repair.
Divergences between developmental and regenerative axon
growth
Although recapitulating aspects of developmental growth enhances
axon regeneration, it is instructive to understand how these two
processes differ. In zebrafish, axon regeneration is mediated by
activating regeneration-associated gene expression using promoter
elements that are not active during embryonic development
(Udvadia et al., 2001). Similarly, mouse DRG neurons that
regenerate because of a conditioning lesion rely on signalling
pathways that are partially distinct from those mediating their
outgrowth during development (Liu and Snider, 2001). Indeed, the
global patterns of gene expression that are active during peripheral
axon regeneration only partially overlap with those that are active
during developmental axon growth (Chandran et al., 2016; Tedeschi
et al., 2016). Thus, the gene expression programs underlying
regeneration in non-mammalian vertebrates and in the mammalian
PNS are at least partially distinct from those underlying
developmental axon growth.
Some intracellular processes key to axon outgrowth are either
unique to or particularly important for regeneration. For example, in
adulthood, the distance between the axon tip and cell body of most
neurons is far longer than at any stage of embryonic development. In
Axotomy
*Cell death
A No treatment
Axotomy
*
B PTEN and SOCS3 knockout or
overexpression of OPN/IGF1/CNTF
Regeneration
and
synapse
formation
C PTEN knockout, cAMP
and inflammatory stimulation
Axotomy
*
Regeneration,
synapse
formation
and
myelination
RGC
Dying
RGC
Target
cell *Injury site
Degenerating
distal axon
Regenerated
axon and
synapse
Original
myelin
New
myelin
Key
Fig. 5. Regenerated axon myelination is not an automatic process and
may involve inflammatory signalling. (A) A mouse retinal ganglion cell
(RGC) projects its axon via the optic nerve to a target cell in the superior
colliculus. The axon is myelinated by oligodendrocytes. After axotomy, the
RGC axon distal to the injury site degenerates and the RGC eventually dies.
(B) The combined knockout of PTEN and SOCS3 or the overexpression of
OPN, IGF1 and CNTF stimulates an axotomized RGC to survive and to
regenerate its axon back to, and form a synapse with, its putative target.
However, the regenerated axon lacks new myelin, leading to poor action
potential conduction, which prevents functional recovery. (C) PTEN knockout,
cAMP and inflammatory (zymosan) stimulation also prompts an axotomized
RGC to regenerate its axon back to its target, where it forms a synapse. With
this treatment, the regenerated axon is myelinated by oligodendrocytes and
forms new nodes of Ranvier.
3425
REVIEW Development (2017) 144, 3417-3429 doi:10.1242/dev.148312
DEVELOPMENT
line with this, axonal transport is crucial for regeneration (He and
Jin, 2016). In the retrograde direction, initiation of the regenerative
program requires signalling from the injured axon to the cell body
(Plunet et al., 2002). The upregulation of regeneration-associated
genes in response to injury is less pronounced in the neuron when it
is axotomized at a longer distance from its cell body compared with
when it is axotomized closer to its cell body (Fernandes et al., 1999).
STAT3 and dual leucine zipper kinase (DLK) are retrograde
messengers key to injury signalling (Lee et al., 2004; Shin et al.,
2012). In the anterograde direction, the building blocks of the axon,
including membrane, proteins and organelles are transported
towards the tip from the cell body to generate new axon. We
surmise that the inability of long descending and ascending axons in
the adult mammalian spinal cord to regenerate is due in part to the
failure of these systems to adequately transport new materials.
Conclusions
Recent studies have demonstrated that recapitulating aspects of
developmental axon growth can promote regeneration in the adult.
However, many challenges remain. For example, it is unclear how
regenerating axons form functional circuits. For many
manipulations, whether regenerated axons find their targets, form
synapses, are myelinated or become functional circuits following
injury or disease is unknown. The development of circuit-specific
genetic technologies to activate or silence neurons will allow
researchers to assess functional connectivity following CNS injury
(Hilton et al., 2016a; Kadoya et al., 2016). These technologies allow
transient activation or silencing of neuronal activity in response to
light (Deisseroth, 2011) or to designer drugs (Roth, 2016). For
example, a dual-virus approach was recently used to express an
engineered G-protein-coupled receptor, the designer receptor
exclusively activated by designer drug hM4Di (DREADD),
within dorsolaterally projecting mouse corticospinal neurons
(Hilton et al., 2016a). hM4Di activation transiently hyperpolarizes
the neuron and suppresses presynaptic glutamate release (Roth,
2016), thus allowing the role of the dorsolateral corticospinal
pathway in mediating spontaneous recovery following cervical
spinal cord injury to be determined (Hilton et al., 2016a). In
principle, such strategies will allow researchers to assess whether a
manipulation promotes synaptic transmission or to probe the
necessity of specific neuron populations in function after injury
(Hilton et al., 2016a; Jayaprakash et al., 2016; Kadoya et al., 2016).
The employment of novel rabies and adeno-associated viral
strategies to dissect the anatomy and functionality of genetically
defined circuits will also help the field of neuroregeneration (Azim
et al., 2014; Esposito et al., 2014; Zampieri et al., 2014). Bioengineered
rabies viruses permit the illumination and interrogation of synaptic
partners of genetically and/or anatomically defined subpopulations
of neurons, and were recently used to analyse synaptic connectivity
between transplanted cells and host axons after spinal cord injury
(Adler et al., 2017). With the recent development of self-
inactivating rabies virus (Ciabatti et al., 2017), which expands the
temporal window of the study of neural circuitry, it is possible to
precisely determine the synaptic connectivity of regenerating axons.
Ultimately, our understanding of the relationship between axon
regeneration and functional recovery remains rudimentary. In some
cases, axon regeneration is associated with worse functional
outcome (Takeoka et al., 2011; Wang et al., 2015), signifying the
importance of understanding the relationship between these
variables. In this regard, investigating synaptic connectivity
during embryonic development may provide clues as to how
regenerating axons can be directed into forming functional circuits.
For example, classical studies in the visual system have
demonstrated the influence of neuronal activity in synapse
formation, elimination and re-arrangement (Katz and Shatz,
1996). Whether neuronal activity can influence synaptic
connectivity of regenerating axons in a similar manner is unclear.
At the same time, axotomized neurons can form new circuits,
underlying recovery following injury in the absence of regeneration-
promoting treatment (Bareyre et al., 2004). Thus, exploring how
therapies that promote axon regeneration can build on, and not
antagonize, endogenous repair mechanisms will be essential for
their translation to the clinic.
Acknowledgements
We thank Dr Daniele Bano and Dr Andreas Husch for discussions on the review.
Competing interests
The authors declare no competing or financial interests.
Funding
B.J.H. is supported by a Wings for Life (WfL) Aguayo-Tator Mentoring Fellowship
and a non-stipendiary European Molecular Biology Organization (EMBO) long-term
fellowship (ALTF 28-2017). F.B. is supported by the Deutsches Zentrum fu
̈r
Neurodegenerative Erkrankungen, the International Foundation for Research in
Paraplegia, WfL and the Deutsche Forschungsgemeinschaft.
References
Adler, A. F., Lee-Kubli, C., Kumamaru, H., Kadoya, K. and Tuszynski, M. H.
(2017). Comprehensive monosynaptic rabies virus mapping of host
connectivity with neural progenitor grafts after spinal cord injury. Stem Cell Rep.
8, 1525-1533.
Al-Ali, H., Ding, Y., Slepak, T., Wu, W., Sun, Y., Martinez, Y., Xu, X.-M., Lemmon,
V. and Bixby, J. (2017). The mTOR substrate S6 Kinase 1 (S6K1) is a negative
regulator of axon regeneration and a potential drug target for Central Nervous
System injury. J. Neurosci. 37, 7079-7095.
Alto, L. T., Havton, L. A., Conner, J. M., Hollis, E. R., II, Blesch, A. and
Tuszynski, M. H. (2009). Chemotropic guidance facilitates axonal regeneration
and synapse formation after spinal cord injury. Nat. Neurosci. 12, 1106-1113.
Anderson, M. A., Burda, J. E., Ren, Y., Ao, Y., OShea, T. M., Kawaguchi, R.,
Coppola, G., Khakh, B. S., Deming, T. J. and Sofroniew, M. V. (2016). Astrocyte
scar formation aids central nervous system axon regeneration. Nature 532,
195-200.
Andrews, M. R., Czvitkovich, S., Dassie, E., Vogelaar, C. F., Faissner, A., Blits,
B., Gage, F. H. and Fawcett, J. W.(2009). α9 integ rin promotes neurite outgrowth
on tenascin-C and enhances sensory axon regeneration. J. Neurosci. 29,
5546-5557.
Assinck, P., Duncan, G. J., Hilton, B. J., Plemel, J. R. and Tetzlaff, W.(2017). Cell
transplantation therapy for spinal cord injury. Nat. Neurosci. 20, 637-647.
Azim, E., Jiang, J., Alstermark, B. and Jessell, T. M. (2014). Skilled reaching
relies on a V2a propriospinal internal copy circuit. Nature 508, 357.
Bareyre, F. M., Kerschensteiner, M., Raineteau, O., Mettenleiter, T. C.,
Weinmann, O. and Schwab, M. E. (2004). The injured spinal cord
spontaneously forms a new intraspinal circuit in adult rats. Nat. Neurosci. 7,
269-277.
Bartus, K., Galino, J., James, N. D., Hernandez-Miranda, L. R., Dawes, J. M.,
Fricker, F. R., Garratt, A. N., McMahon, S. B., Ramer, M. S. and Birchmeier, C.
(2016). Neuregulin-1 controls an endogenous repair mechanism after spinal cord
injury. Brain 139, 1394-1416.
Bei, F., Lee, H. H. C., Liu, X., Gunner, G., Jin, H., Ma, L., Wang, C., Hensch, T. K.,
Frank, E. and Sanes, J. R. (2016). Restoration of visual function by enhancing
conduction in regenerated axons. Cell 164, 219-232.
Belin, S., Nawabi, H., Wang, C., Tang, S., Latremoliere, A., Warren, P., Schorle,
H., Uncu, C., Woolf, C. J. and He, Z. (2015). Injury-induced decline of intrinsic
regenerative ability revealed by quantitative proteomics. Neuron 86, 1000-1014.
Bielas, S. L., Serneo, F. F., Chechlacz, M., Deerinck, T. J., Perkins, G. A., Allen,
P. B., Ellisman, M. H. and Gleeson, J. G. (2007). Spinophilin facilitates
dephosphorylation of doublecortin by PP1 to mediate microtubule bundling at the
axonal wrist. Cell 129, 579-591.
Blackmore, M. G., Wang, Z., Lerch, J. K., Motti, D., Zhang, Y. P., Shields, C. B.,
Lee, J. K., Goldberg, J. L., Lemmon, V. P. and Bixby, J. L. (2012). Kru
̈ppel-
like Factor 7 engineered for transcriptional activation promotes axon
regeneration in the adult corticospinal tract. Proc. Natl Acad. Sci. USA 109,
7517-7522.
Bradbury, E. J., Moon, L. D. F., Popat, R. J., King, V. R., Bennett, G. S., Patel,
P. N., Fawcett, J. W. and McMahon, S. B. (2002). Chondroitinase ABC promotes
functional recovery after spinal cord injury. Nature 416, 636-640.
3426
REVIEW Development (2017) 144, 3417-3429 doi:10.1242/dev.148312
DEVELOPMENT
Bradke, F. and Dotti, C. G. (1997). Neuronal polarity: vectorial cytoplasmic flow
precedes axon formation. Neuron 19, 1175-1186.
Bradke, F. and Dotti, C. G. (1999). The role of local actin instability in axon
formation. Science 283, 1931-1934.
Bradke, F., Fawcett, J. W. and Spira, M. E. (2012). Assembly of a new growth cone
after axotomy: the precursor to axon regeneration. Nat. Rev. Neurosci. 13,
183-193.
Brittis, P. A., Canning, D. R. and Silver, J. (1992). Chondroitin sulfate as a
regulator of neuronal patterning in the retina. Science 255, 733.
Brunn, G. J., Hudson, C. C., Sekulić, A., Williams, J. M., Hosoi, H., Houghton,
P. J., Lawrence, J. C. and Abraham, R. T. (1997). Phosphorylation of the
translational repressor PHAS-I by the mammalian target of rapamycin. Science
277, 99-101.
Burnside, E. E. and Bradbury, E. J. (2014). Manipulating the extracellular matrix
and its role in brain and spinal cord plasticity and repair. Neuropathol. Appl.
Neurobiol. 40, 26-59.
Butler, E. G. and Ward, M. B. (1967). Reconstitution of the spinal cord after ablation
in adult Triturus. Dev. Biol. 15, 464-486.
Byrne, A. B., Walradt, T., Gardner, K. E., Hubbert, A., Reinke, V. and
Hammarlund, M. (2014). Insulin/IGF1 signaling inhibits age-dependent axon
regeneration. Neuron 81, 561-573.
Cafferty, W. B., Kim, J.-E., Lee, J.-K. and Strittmatter, S. M. (2007). Response to
correspondence: Kim et al., axon regeneration in young adult mice lacking Nogo-
A/B.Neuron 38, 187199. Neuron 54, 195-199.
Cartoni, R., Norsworthy, M. W., Bei, F., Wang, C., Li, S., Zhang, Y., Gabel, C. V.,
Schwarz, T. L. and He, Z. (2016). The mammalian-specific protein armcx1
regulates mitochondrial transport during axon regeneration. Neuron 92,
1294-1307.
Cavallucci, V., Bisicchia, E., Cencioni, M. T., Ferri, A., Latini, L., Nobili, A.,
Biamonte, F., Nazio, F., Fanelli, F. and Moreno, S. (2014). Acute focal brain
damage alters mitochondrial dynamics and autophagy in axotomized neurons.
Cell Death Dis. 5, e1545.
Chandel, N. S. (2014). Mitochondria as signaling organelles. BMC Biol. 12, 34.
Chandran, V., Coppola, G., Nawabi, H., Omura, T., Versano, R., Huebner, E. A.,
Zhang, A., Costigan, M., Yekkirala, A., Barrett, L. et al. (2016). A systems-level
analysis of the peripheral nerve intrinsic axonal growth program. Neuron 89,
956-970.
Chauvin, C., Koka, V., Nouschi, A., Mieulet, V., Hoareau-Aveilla, C., Dreazen, A.,
Cagnard, N., Carpentier, W., Kiss, T. and Meyuhas, O. (2014). Ribosomal
protein S6 kinase activity controls the ribosome biogenesis transcriptional
program. Oncogene 33, 474-483.
Ciabatti, E., González-Rueda, A., Mariotti, L., Morgese, F. and Tripodi, M.
(2017). Life-long genetic and functional access to neural circuits using self-
inactivating rabies virus. Cell 170, 382-392.e14.
Clarke, L. E. and Barres, B. A. (2013). Emerging roles of astrocytes in neural circuit
development. Nat. Rev. Neurosci. 14, 311-321.
Coles, C. H. and Bradke, F. (2015). Coordinating neuronal actinmicrotubule
dynamics. Curr. Biol. 25, R677-R691.
Cregg, J. M., DePaul, M. A., Filous, A. R., Lang, B. T., Tran, A. and Silver, J.
(2014). Functional regeneration beyond the glial scar. Exp. Neurol. 253, 197-207.
Cross, D. J., Garwin, G. G., Cline, M. M., Richards, T. L., Yarnykh, V., Mourad,
P. D., Ho, R. J. and Minoshima, S. (2015). Paclitaxel improves outcome from
traumatic brain injury. Brain Res. 1618, 299-308.
Deisseroth, K. (2011). Optogenetics. Nat. Methods 8, 26-29.
Dent, E. W., Gupton, S. L. and Gertler, F. B. (2011). The growth cone cytoskeleton
in axon outgrowth and guidance. Cold Spring Harbor Perspect. Biol. 3, a001800.
DeVivo, M. J. and Chen, Y. (2011). Trends in new injuries, prevalent cases, and
aging with spinal cord injury. Arch. Phys. Med. Rehabil. 92, 332-338.
Dietz, V. and Fouad, K. (2014). Restoration of sensorimotor functions after spinal
cord injury. Brain 137, 654-667.
Dow, K. E., Ethell, D. W., Steeves, J. D. and Riopelle, R. J. (1994). Molecular
correlates of spinal cord repair in the embryonic chick: heparan sulfate and
chondroitin sulfate proteoglycans. Exp. Neurol. 128, 233-238.
Du, K., Zheng, S., Zhang, Q., Li, S., Gao, X., Wang, J., Jiang, L. and Liu, K.
(2015). Pten deletion promotes regrowth of corticospinal tract axons 1 year after
spinal cord injury. J. Neurosci. 35, 9754-9763.
Duan, X., Qiao, M., Bei, F., Kim, I.-J., He, Z. and Sanes, J. R. (2015). Subtype-
specific regeneration of retinal ganglion cells following axotomy: effects of
osteopontin and mTOR signaling. Neuron 85, 1244-1256.
Duncan, G. J., Plemel, J. R., Assinck, P., Manesh, S. B., Muir, F. G., Hirata, R.,
Berson, M., Liu, J., Wegner, M. and Emery, B. (2017). Myelin regulatory factor
drives remyelination in multiple sclerosis. Acta Neuropathol. 1-20 (in press)
Enes, J., Langwieser, N., Ruschel, J., Carballosa-Gonzalez, M. M., Klug, A.,
Traut, M. H., Ylera, B., Tahirovic, S., Hofmann, F. and Stein, V. (2010).
Electrical activity suppresses axon growth through Ca v 1.2 channels in adult
primary sensory neurons. Curr. Biol. 20, 1154-1164.
Eroglu, C., Allen, N. J., Susman, M. W., ORourke, N. A., Park, C. Y., O
̈
zkan, E.,
Chakraborty, C., Mulinyawe, S. B., Annis, D. S. and Huberman, A. D. (2009).
Gabapentin receptor α2δ-1 is a neuronal thrombospondin receptor responsible for
excitatory CNS synaptogenesis. Cell 139, 380-392.
Ertu
̈rk, A., Hellal, F., Enes, J. and Bradke, F. (2007). Disorganized microtubules
underlie the formation of retraction bulbs and the failure of axonal regeneration.
J. Neurosci. 27, 9169-9180.
Esposito, M. S., Capelli, P. and Arber, S. (2014). Brainstem nucleus MdV mediates
skilled forelimb motor tasks. Nature 508, 351-356.
Fancy, S. P. J., Chan, J. R., Baranzini, S. E., Franklin, R. J. M. and Rowitch, D. H.
(2011). Myelin regeneration: a recapitulation of development? Annu. Rev.
Neurosci. 34, 21-43.
Faulkner, J. R., Herrmann, J. E., Woo, M. J., Tansey, K. E., Doan, N. B. and
Sofroniew, M. V. (2004). Reactiveastrocytes protect tissue and preserve function
after spinal cord injury. J. Neurosci. 24, 2143-2155.
Fenrich, K. K. and Rose, P. K. (2009). Spinal interneuron axons spontaneously
regenerate after spinal cord injury in the adult feline. J. Neurosci. 29,
12145-12158.
Fernandes, K. J. L., Fan, D.-P., Tsui, B. J., Cassar, S. L. and Tetzlaff, W. (1999).
Influence of the axotomy to cell body distance in rat rubrospinal and spinal
motoneurons: differential regulation of GAP-43, tubulins, and neurofilament-M.
J. Comp. Neurol. 414, 495-510.
Fields, R. D. (2015). A new mechanism of nervous system plasticity: activity-
dependent myelination. Nat. Rev. Neurosci. 16, 756-767.
Filous, A. R., Miller, J. H., Coulson-Thomas, Y. M., Horn, K. P., Alilain, W. J. and
Silver, J. (2010). Immature astrocytes promote CNS axonal regeneration when
combined with chondroitinase ABC. Dev. Neurobiol. 70, 826-841.
Filous, A. R., Tran, A., Howell, C. J., Busch, S. A., Evans, T. A., Stallcup, W. B.,
Kang, S. H., Bergles, D. E., Lee, S.-I. and Levine, J. M. (2014). Entrapment via
synaptic-like connections between NG2 proteoglycan+ cells and dystrophic axons
in the lesion plays a role in regeneration failure after spinal cord injury. J. Neurosci.
34, 16369-16384.
Fitch, M. T. and Silver, J. (1997). Glial cell extracellular matrix: boundaries for axon
growth in development and regeneration. Cell Tissue Res. 290, 379-384.
Flynn, K. C., Hellal, F., Neukirchen, D., Jacob, S., Tahirovic, S., Dupraz, S.,
Stern, S., Garvalov, B. K., Gurniak, C., Shaw, A. E. et al. (2012). ADF/cofilin-
mediated actin retrograde flow directs neurite formation in the developing brain.
Neuron 76, 1091-1107.
Foscarin, S., Raha-Chowdhury, R., Fawcett, J. and Kwok, J. (2017). Brain ageing
changes proteoglycan sulfation, rendering perineuronal nets more inhibitory.
Aging 9, 1607-1622.
Gee, N. S., Brown, J. P., Dissanayake, V. U. K., Offord, J., Thurlow, R. and
Woodruff, G. N. (1996). The novel anticonvulsant drug, gabapentin (Neurontin),
binds to the subunit of a calcium channel. J. Biol. Chem. 271, 5768-5776.
Geoffroy, C. G. and Zheng, B. (2014). Myelin-associated inhibitors in axonal
growth after CNS injury. Curr. Opin. Neurobiol. 27, 31-38.
Geoffroy, C. G., Hilton, B. J., Tetzlaff, W. and Zheng, B. (2016). Evidence for an
age-dependent decline in axon regeneration in the adult mammalian central
nervous system. Cell Rep. 15, 238-246.
Giger, R. J., Hollis, E. R. and Tuszynski, M. H. (2010). Guidance molecules in
axon regeneration. Cold Spring Harbor Perspect. Biol. 2, a001867.
Gioran, A., Nicotera, P. and Bano, D. (2014). Impaired mitochondrial respiration
promotes dendritic branching via the AMPK signaling pathway. Cell Death Dis. 5,
e1175.
Goldberg, D. J. and Burmeister, D. W. (1986). Stages in axon formation:
observations of growth of Aplysia axons in culture using video-enhanced contrast-
differential interference contrast microscopy. J. Cell Biol. 103, 1921-1931.
Gomis-Ru
̈th, S., Wierenga, C. J. and Bradke, F. (2008). Plasticity of polarization:
changing dendrites into axons in neurons integrated in neuronal circuits. Curr.
Biol. 18, 992-1000.
Goodin, S., Kane, M. P. and Rubin, E. H. (2004). Epothilones: mechanism of action
and biologic activity. J. Clin. Oncol. 22, 2015-2025.
Han, S. M., Baig, H. S. and Hammarlund, M. (2016). Mitochondria localize to
injured axons to support regeneration. Neuron 92, 1308-1323.
Hara, M., Kobayakawa, K., Ohkawa, Y., Kuma maru, H., Yokota, K ., Saito, T.,
Kijima, K ., Yoshizaki, S., Har imaya, K. and Nakashima, Y. (2017). Interaction
of reactive astrocytes with type I co llagen induces astrocytic scar formation
through the integrin-N-cadherin pathway after spinal cord injury. Nat. Med. 23,
818-828.
He, Z. and Jin, Y. (2016). Intrinsic control of axon regeneration. Neuron 90, 437-451.
Hellal, F., Hurtado, A., Ruschel, J., Flynn, K. C., Laskowski, C. J., Umlauf, M.,
Kapitein, L. C., Strikis, D., Lemmon, V. and Bixby, J. (2011). Microtubule
stabilization reduces scarring and causes axon regeneration after spinal cord
injury. Science 331, 928-931.
Hilton, B. J., Anenberg, E., Harrison, T. C., Boyd, J. D., Murphy, T. H. and
Tetzlaff, W. (2016a). Re-establishment of cortical motor output maps and
spontaneous functional recovery via spared dorsolaterally projecting corticospinal
neurons after dorsal column spinal cord injury in adult mice. J. Neurosci. 36,
4080-4092.
Hilton, B. J., Moulson, A. J. and Tetzlaff, W. (2016b). Neuroprotection and
secondary damage following spinal cord injury: concepts and methods. Neurosci.
Lett. 652, 3-10.
Hollis, E. R., II, Ishiko, N., Yu, T., Lu, C.-C., Haimovich, A., Tolentino, K.,
Richman, A., Tury, A., Wang, S.-H. and Pessian, M. (2016). Ryk controls
3427
REVIEW Development (2017) 144, 3417-3429 doi:10.1242/dev.148312
DEVELOPMENT
remapping of motor cortex during functional recovery after spinal cord injury. Nat.
Neurosci. 19, 697-705.
Hua, J. Y. and Smith, S. J. (2004). Neural activity and the dynamics of central
nervous system development. Nat. Neurosci. 7, 327.
Hubel, D. H. and Wiesel, T. N. (1970). The period of susceptibility to the
physiological effects of unilateral eye closure in kittens. J. Physiol. 206, 419-436.
Hunt, M., Lu, P. and Tuszynski, M. H. (2017). Myelination of axons emerging from
neural progenitor grafts after spinal cord injury. Exp. Neurol. 296, 69-73.
Jayaprakash, N., Wang, Z., Hoeynck, B., Krueger, N., Kramer, A., Balle, E.,
Wheeler, D. S., Wheeler, R. A. and Blackmore, M. G. (2016). Optogenetic
interrogation of functional synapse formation by corticospinal tract axons in the
injured spinal cord. J. Neurosci. 36, 5877-5890.
Jin, Y., Dougherty, S. E., Wood, K., Sun, L., Cudmore, R. H., Abdalla, A.,
Kannan, G., Pletnikov, M., Hashemi, P. and Linden, D. J. (2016). Regrowth
of serotonin axons in the adult mouse brain following injury. Neuron 91,
748-762.
Kadoya, K., Lu, P., Nguyen, K., Lee-Kubli, C., Kumamaru, H., Yao, L., Knackert,
J., Poplawski, G., Dulin, J. N., Strobl, H. et al. (2016). Spinal cord reconstitution
with homologous neural grafts enables robust corticospinal regeneration. Nat.
Med. 22, 479-487.
Kang, J.-S., Tian, J.-H., Pan, P.-Y., Zald, P., Li, C., Deng, C. and Sheng, Z.-H.
(2008). Docking of axonal mitochondria by syntaphilin controls their mobility and
affects short-term facilitation. Cell 132, 137-148.
Kasthuri, N., Hayworth, K. J., Berger, D. R., Schalek, R. L., Conchello, J. A.,
Knowles-Barley, S., Lee, D., Vázquez-Reina, A., Kaynig, V. and Jones, T. R.
(2015). Saturated reconstruction of a volume of neocortex. Cell 162, 648-661.
Katz, L. C. and Shatz, C. J. (1996). Synaptic activity and the construction of cortical
circuits. Science 274, 1133.
Keirstead, H. S., Hasan, S. J., Muir, G. D. and Steeves, J. D. (1992). Suppression
of the onset of myelination extends the permissive period for the functional repair
of embryonic spinal cord. Proc. Natl Acad. Sci. USA 89, 11664-11668.
Kraus, D. B. and Fadem, B. H. (1987). Reproduction, development and physiology
of the gray short-tailed opossum (Monodelphis domestica). Lab. Anim. Sci. 37,
478-482.
Kwon, B. K., Liu, J., Messerer, C., Kobayashi, N. R., McGraw, J., Oschipok, L.
and Tetzlaff, W. (2002). Survival and regeneration of rubrospinal neurons 1 year
after spinal cord injury. Proc. Natl Acad. Sci. USA 99, 3246-3251.
Laplante, M. and Sabatini, D. M. (2012). mTOR signaling in growth control and
disease. Cell 149, 274-293.
Lawrence, D. G. and Kuypers, H. G. (1968). The functional organization of the
motor system in the monkey. Brain 91, 15-36.
Lee, N., Neitzel, K. L., Devlin, B. K. and MacLennan, A. J. (2004). STAT3
phosphorylation in injured axons before sensory and motor neuron nuclei:
potential role for STAT3 as a retrograde signaling transcription factor. J. Comp.
Neurol. 474, 535-545.
Leske, M. C., Connell, A., Wu, S.-Y., Hyman, L. G. and Schachat, A. P. (1995).
Risk factors for open-angle glaucoma: the barbado s eye study. Arch. Ophthalmol.
113, 918-924.
Li, L., Hutchins, B. I. and Kalil, K. (2009). Wnt5a induces simultaneous cortical
axon outgrowth and repulsive axon guidance through distinct signaling
mechanisms. J. Neurosci. 29, 5873-5883.
Liddelow, S. A., Guttenplan, K. A., Clarke, L. E., Bennett, F. C., Bohlen, C. J.,
Schirmer, L., Bennett, M. L., Mu
̈nch, A. E., Chung, W.-S. and Peterson, T. C.
(2017). Neurotoxic reactive astrocytes are induced by activated microglia. Nature
541, 481-487.
Liu, R.-Y. and Snider, W. D. (2001). Different signaling pathways mediate
regenerative versus developmental sensory axon growth. J. Neurosci. 21,
RC164-RC164.
Liu, Y., Shi, J., Lu, C.-C., Wang, Z.-B., Lyuksyutova, A. I., Song, X.-J. and Zou, Y.
(2005). Ryk-mediated Wnt repulsion regulates posterior-directed growth of
corticospinal tract. Nat. Neurosci. 8, 1151-1159.
Liu, Y., Wang, X., Lu, C.-C., Sherman-Kermen, R., Steward, O., Xu, X.-M. and
Zou, Y. (2008). Repulsive Wnt signaling inhibits axon regeneration after CNS
injury. J. Neurosci. 28, 8376-8382.
Liu, K., Lu, Y., Lee, J. K., Samara, R., Willenberg, R., Sears-Kraxberger, I.,
Tedeschi, A., Park, K. K., Jin, D. and Cai, B. (2010). PTEN deletion enhances
the regenerative ability of adult corticospinal neurons. Nat. Neurosci. 13,
1075-1081.
López-Doménech, G., Serrat, R., Mirra, S., Daniello, S., Somorjai, I., Abad, A.,
Vitureira, N., Garcı
́
a-Arumı
́
, E., Alonso, M. T. and Rodriguez-Prados, M.
(2012). The Eutherian Armcx genes regulate mitochondrial trafficking in neurons
and interact with Miro and Trak2. Nat. Commun. 3, 814.
Lorenzana, A. O., Lee, J. K., Mui, M., Chang, A. and Zheng, B. (2015). A surviving
intact branch stabilizes remaining axon architecture after injury as revealed by in
vivo imaging in the mouse spinal cord. Neuron 86, 947-954.
Low, L. K. and Cheng, H.-J. (2006). Axon pruning: an essential step underlying the
developmental plasticity of neuronal connections. Phil. Trans. R. Soc. B Biol. Sci.
361, 1531-1544.
Lu, P., Wang, Y., Graham, L., McHale, K., Gao, M., Wu, D., Brock, J., Blesch, A.,
Rosenzweig, E. S. and Havton, L. A. (2012). Long-distance growth and
connectivity of neural stem cells after severe spinal cord injury. Cell 150,
1264-1273.
Luo, L. and OLeary, D. D. M. (2005). Axon retraction and degeneration in
development and disease. Annu. Rev. Neurosci. 28,127-156.
Luo, X., Ribeiro, M., Bray, E. R., Lee, D.-H., Yungher, B. J., Mehta, S. T., Thakor,
K. A., Diaz, F., Lee, J. K. and Moraes, C. T. (2016). Enhanced transcriptional
activity and mitochondrial localization of STAT3 co-induce axon regrowth in the
adult central nervous system. Cell Rep. 15, 398-410.
MacLaren, R. E. and Taylor, J. S. H. (1997). Regeneration in the developing optic
nerve: correlating observations in the opossum to other mammalian systems.
Prog. Neurobiol. 53, 381-398.
Marin, M. A., de Lima, S., Gilbert, H.-Y., Giger, R. J., Benowitz, L. and Rasband,
M. N. (2016). Reassembly of excitable domains after CNS axon regeneration.
J. Neurosci. 36, 9148-9160.
Miron, V. E., Boyd, A., Zhao, J.-W., Yuen, T. J., Ruckh, J. M., Shadrach,J. L., van
Wijngaarden, P., Wagers, A. J., Williams, A. and Franklin, R. J. (2013). M2
microglia and macrophages drive oligodendrocyte differentiation during CNS
remyelination. Nat. Neurosci. 16, 1211-1218.
Mokalled, M. H., Patra, C., Dickson, A. L., Endo, T., Stainier, D. Y. R. and Poss,
K. D. (2016). Injury-induced ctgfa directs glial bridging and spinal cord
regeneration in zebrafish. Science 354, 630-634.
Moore, D. L., Blackmore, M. G., Hu, Y., Kaestner, K. H., Bixby, J. L., Lemmon,
V. P. and Goldberg, J. L. (2009). KLF family members regulate intrinsic axon
regeneration ability. Science 326, 298-301.
Morris, R. L. and Hollenbeck, P. J. (1993). The regulation of bidirectional
mitochondrial transport is coordinated with axonal outgrowth. J. Cell Sci. 104,
917-927.
Nawabi, H., Belin, S., Cartoni, R., Williams, P. R., Wang, C., Latremolière, A.,
Wang, X., Zhu, J., Taub, D. G. and Fu, X. (2015). Doublecortin-like kinases
promote neuronal survival and induce growth cone reformation via distinct
mechanisms. Neuron 88, 704-719.
Nicholls, J. and Saunders, N. (1996). Regeneration of immature mammalian spinal
cord after injury. Trends Neurosci. 19, 229-234.
Nielson, J. L., Sears-Kraxberger, I., Strong, M. K., Wong, J. K., Willenberg, R.
and Steward, O. (2010). Unexpected survival of neurons of origin of the pyramidal
tract after spinal cord injury. J. Neurosci. 30, 11516-11528.
ODonnell, M., Chance, R. K. and Bashaw, G. J. (2009). Axon growth and
guidance: receptor regulation and signal transduction. Annu. Rev. Neurosci. 32,
383-412.
ODonnell, K. C., Vargas, M. E. and Sagasti, A. (2013). WldS and PGC-1αregulate
mitochondrial transport and oxidation state after axonal injury. J. Neurosci. 33,
14778-14790.
Okada, Y., Shimazaki, T., Sobue, G. and Okano, H. (2004). Retinoic-acid-
concentration-dependent acquisition of neural cell identity during in vitro
differentiation of mouse embryonic stem cells. Dev. Biol. 275, 124-142.
Omura, T., Omura, K., Tedeschi, A., Riva, P., Painter, M. W., Rojas, L., Martin, J.,
Lisi, V., Huebner, E. A. and Latremoliere, A. (2015). Robustaxonal regeneration
occurs in the injured CAST/Ei mouse CNS. Neuron 86, 1215-1227.
Painter, M. W., Lutz, A. B., Cheng, Y.-C., Latremoliere, A., Duong, K., Miller,
C. M., Posada, S., Cobos, E. J., Zhang, A. X. and Wagers, A. J. (2014).
Diminished Schwann cell repair responses underlie age-associated impaired
axonal regeneration. Neuron 83, 331-343.
Park, K. K., Liu, K., Hu, Y., Smith, P. D., Wang, C., Cai, B., Xu, B., Connolly, L.,
Kramvis, I. and Sahin, M. (2008). Promoting axon regeneration in the adult CNS
by modulation of the PTEN/mTOR pathway. Science 322, 963-966.
Pizzorusso, T., Medini, P., Berardi, N., Chierzi, S., Fawcett, J. W. and Maffei, L.
(2002). Reactivation of ocular dominance plasticity in the adult visual cortex.
Science 298, 1248-1251.
Plemel, J. R., Liu, W.-Q. and Yong, V. W. (2017). Remyelination therapies: a new
direction and challenge in multiple sclerosis. Nat. Rev. Drug Discov. (in press).
Plunet, W., Kwon, B. K. and Tetzlaff, W. (2002). Promoting axonal regeneration in
the central nervous system by enhancing the cell body response to axotomy.
J. Neurosci. Res. 68, 1-6.
Purves, D. and Lichtman, J. W. (1980). Elimination of synapses in the developing
nervous system. Science 210, 153-157.
Puttagunta, R., Tedeschi, A., Sória, M. G., Hervera, A., Lindner, R., Rathore,
K. I., Gaub, P., Joshi, Y., Nguyen, T. and Schmandke, A. (2014). PCAF-
dependent epigenetic changes promote axonal regeneration in the central
nervous system. Nat. Commun. 5, 3527.
Pyka, M., Wetzel, C., Aguado, A., Geissler, M., Hatt, H. and Faissner, A. (2011).
Chondroitin sulfate proteoglycans regulate astrocyte-dependent synaptogenesis
and modulate synaptic activity in primary embryonic hippocampal neurons.
Eur. J. Neurosci. 33, 2187-2202.
Rawson, R. L., Yam, L., Weimer, R. M., Bend, E. G., Hartwieg, E., Horvitz, H. R.,
Clark, S. G. and Jorgensen, E. M. (2014). Axons degenerate in the absence of
mitochondria in C. elegans. Curr. Biol. 24, 760-765.
Reier, P. J., Bregman, B. S. and Wujek, J. R. (1986). Intraspinal transplantation of
embyronic spinal cord tissue in neonatal and adult rats. J. Comp. Neurol. 247,
275-296.
Roth, B. L. (2016). DREADDs for neuroscientists. Neuron 89, 683-694.
3428
REVIEW Development (2017) 144, 3417-3429 doi:10.1242/dev.148312
DEVELOPMENT
Ruschel, J., Hellal, F., Flynn, K. C., Dupraz, S., Elliott, D. A., Tedeschi, A., Bates,
M., Sliwinski, C., Brook, G., Dobrindt, K. et al. (2015). Systemic administration
of epothilone B promotes axon regeneration after spinal cord injury. Science 348,
347-352.
Sacco, R. L., Benjamin, E. J., Broderick, J. P., Dyken, M., Easton, J. D.,
Feinberg, W. M., Goldstein, L. B., Gorelick, P. B., Howard, G. and Kittner, S. J.
(1997). Risk factors. Stroke 28, 1507-1517.
Saunders, N. R., Deal, A., Knott, G. W., Varga, Z. M. and Nicholls, J. G. (1995).
Repair and recovery following spinal cord injury in a neonatal marsupial
(Monodelphis domestica). Clin. Exp. Pharmacol. Physiol. 22, 518-526.
Schwab, M. E. and Strittmatter, S. M. (2014). Nogo limits neural plasticity and
recovery from injury. Curr. Opin. Neurobiol. 27, 53-60.
Schwab, M. E., Kapfhammer, J. P. and Bandtlow, C. E. (1993). Inhibitors of
neurite growth. Annu. Rev. Neurosci. 16, 565-595.
Schwarz, T. L. (2013). Mitochondrial trafficking in neurons. Cold Spring Harbor
Perspect. Biol. 5, a011304.
Seiradake, E., Jones, E. Y. and Klein, R. (2016). Structural perspectives on axon
guidance. Annu. Rev. Cell Dev. Biol. 32, 577-608.
Sengottuvel, V., Leibinger, M., Pfreimer, M., Andreadaki, A. and Fischer, D.
(2011). Taxol facilitates axon regeneration in the mature CNS. J. Neurosci. 31,
2688-2699.
Shen, K. and Scheiffele, P. (2010). Genetics and cell biology of building specific
synaptic connectivity. Annu. Rev. Neurosci. 33, 473-507.
Shigeoka, T., Jung, H., Jung, J., Turner-Bridger, B., Ohk, J., Lin, J. Q., Amieux,
P. S. and Holt, C. E. (2016). Dynamic axonal translation in developing and mature
visual circuits. Cell 166, 181-192.
Shin, J. E., Cho, Y., Beirowski, B., Milbrandt, J., Cavalli, V. and DiAntonio, A.
(2012). Dual leucine zipper kinase is required for retrograde injury signaling and
axonal regeneration. Neuron 74, 1015-1022.
Silver, J. (2016). The glial scar is more than just astrocytes. Exp. Neurol. 286,
147-149.
Smith, D. H. (2009). Stretch growth of integrated axon tracts: extremes and
exploitations. Prog. Neurobiol. 89, 231-239.
Smith, G. M., Miller, R. H. and Silver, J. (1986). Changing role of forebrain
astrocytes during development, regenerative failure, and induced regeneration
upon transplantation. J. Comp. Neurol. 251, 23-43.
Sofroniew, M. V. (2015). Astrocyte barriers to neurotoxic inflammation. Nat. Rev.
Neurosci. 16, 249-263.
Spillane, M., Ketschek, A., Merianda, T. T., Twiss, J. L. and Gallo, G. (2013).
Mitochondria coordinate sites of axon branching through localized intra-axonal
protein synthesis. Cell reports 5, 1564-1575.
Sporns, O., Tononi, G. and Ko
̈tter, R. (2005). The human connectome: a structural
description of the human brain. PLoS Comput. Biol. 1, e42.
Stanfield, B. B., OLeary, D. D. M. and Fricks, C. (1982). Selective collateral
elimination in early postnatal development restricts cortical distribution of rat
pyramidal tract neurones. Nature 298, 371-373.
Starkey, M. L., Barritt, A. W., Yip, P. K., Davies, M., Hamers, F. P. T., McMahon,
S. B. and Bradbury, E. J. (2005). Assessing behavioural function following a
pyramidotomy lesion of the corticospinal tract in adult mice. Exp. Neurol. 195,
524-539.
Sun, F., Park, K. K., Belin, S., Wang, D., Lu, T., Chen, G., Zhang, K., Yeung, C.,
Feng, G. and Yankner, B. A. (2011). Sustained axon regeneration induced by co-
deletion of PTEN and SOCS3. Nature 480, 372-375.
Takeoka, A., Jindrich, D. L., Muñoz-Quiles, C., Zhong, H., van den Brand, R.,
Pham, D. L., Ziegler, M. D., Ramón-Cueto, A., Roy, R. R. and Edgerton, V. R.
(2011). Axon regeneration can facilitate or suppress hindlimb function after
olfactory ensheathing glia transplantation. J. Neurosci. 31, 4298-4310.
Tan, C. L., Kwok, J. C. F., Patani, R., Chandran, S. and Fawcett, J. W. (2011).
Integrin activation promotes axon growth on inhibitory chondroitin sulfate
proteoglycans by enhancing integrin signaling. J. Neurosci. 31, 6289-6295.
Tedeschi, A. and Bradke, F. (2017). Spatial and temporal arrangement of neuronal
intrinsic and extrinsic mechanisms controlling axon regeneration. Curr. Opin.
Neurobiol. 42, 118-127.
Tedeschi, A., Dupraz, S., Laskowski, C. J., Xue, J., Ulas, T., Beyer,M., Schultze,
J. L. and Bradke, F. (2016). The calcium channel subunit Alpha2delta2
suppresses axon regeneration in the adult CNS. Neuron 92, 419-434.
Tetzlaff, W., Alexander, S. W., Miller, F. D. and Bisby, M. A. (1991). Response of
facial and rubrospinal neurons to axotomy: changes in mRNA expression for
cytoskeletal proteins and GAP-43. J. Neurosci. 11, 2528-2544.
Tuszynski, M. H. and Steward, O. (2012). Concepts and methods for the study of
axonal regeneration in the CNS. Neuron 74, 777-791.
Udvadia, A. J., Koster, R. and Skene, J. (2001). GAP-43 promoter elements in
transgenic zebrafish reveal a difference in signals for axon growth during CNS
development and regeneration. Development 128, 1175-1182.
Vaarmann, A., Mandel, M., Zeb, A., Wareski, P., Liiv, J., Kuum, M., Antsov, E.,
Liiv, M., Cagalinec, M. and Choubey, V. (2016). Mitochondrial biogenesis is
required for axonal growth. Development 143, 1981-1992.
van Spronsen, M., Mikhaylova, M., Lipka, J., Schlager, M. A., van den Heuvel,
D. J., Kuijpers, M., Wulf, P. S., Keijzer, N., Demmers, J. and Kapitein, L. C.
(2013). TRAK/Milton motor-adaptor proteins steer mitochondrial trafficking to
axons and dendrites. Neuron 77, 485-502.
Venkatesh, I. and Blackmore, M. G. (2016). Selecting optimal combinations of
transcription factors to promote axon regeneration: why mechanisms matter.
Neurosci. Lett. 652, 64-73.
Wahl, A. S., Omlor, W., Rubio, J. C., Chen, J. L., Zheng, H., Schro
̈ter, A., Gullo,
M., Weinmann, O., Kobayashi, K. and Helmchen, F. (2014). Asynchronous
therapy restores motor control by rewiring of the rat corticospinal tract after stroke.
Science 344,1250-1255.
Wang, Z., Reynolds, A., Kirry, A., Nienhaus, C. and Blackmore, M. G. (2015).
Overexpression of sox11 promotes corticospinal tract regeneration after
spinal injury while interfering with functional recovery. J. Neurosci. 35,
3139-3145.
Warner, F. M., Cragg, J. J., Jutzeler, C. R., Ro
̈hrich, F., Weidner, N., Saur, M.,
Maier, D. D., Schuld, C., Sites, E. and Curt, A. (2017). Early administration of
Gabapentinoids improves motor recovery after human spinal cord injury. Cell Rep.
18, 1614-1618.
Wegrzyn, J., Potla, R., Chwae, Y.-J., Sepuri, N. B. V., Zhang, Q., Koeck, T.,
Derecka, M., Szczepanek, K., Szelag, M. and Gornicka, A. (2009). Function of
mitochondrial Stat3 in cellular respiration. Science 323, 793-797.
Williams, G. S. B., Boyman, L., Chikando, A. C., Khairallah, R. J. and Lederer,
W. (2013). Mitochondrial calcium uptake. Proc. Natl Acad. Sci. USA 110,
10479-10486.
Witte, H. and Bradke, F. (2008). The role of the cytoskeleton during neuronal
polarization. Curr. Opin. Neurobiol. 18, 479-487.
Witte, H., Neukirchen, D. and Bradke, F. (2008). Microtubule stabilization specifies
initial neuronal polarization. J. Cell Biol. 180, 619-632.
Yang, L., Miao, L., Liang, F., Huang, H., Teng, X., Li, S., Nuriddinov, J., Selzer,
M. E. and Hu, Y. (2014). The mTORC1 effectors S6K1 and 4E-BP play different
roles in CNS axon regeneration. Nat. Commun. 5, 5416.
Ylera, B., Ertu
̈rk, A., Hellal, F., Nadrigny, F., Hurtado, A., Tahirovic, S., Oudega,
M., Kirchhoff, F. and Bradke, F. (2009). Chronically CNS-injured adult sensory
neurons gain regenerative competence upon a lesion of their peripheral axon.
Curr. Biol. 19, 930-936.
Zampieri, N., Jessell, T. M. and Murray, A. J. (2014). Mapping sensory circuits by
anterograde transsynaptic transfer of recombinant rabies virus. Neuron 81,
766-778.
Zhou, B., Yu, P., Lin, M.-Y., Sun, T., Chen, Y. and Sheng, Z.-H. (2016). Facilitation
of axon regeneration by enhancing mitochondrial transport and rescuing energy
deficits. J. Cell Biol. 214, 103-119.
Zukor, K. A., Kent, D. T. and Odelberg, S. J. (2011). Meningeal cells and glia
establish a permissive environment for axon regeneration after spinal cord injury
in newts. Neural Dev. 6,1.
3429
REVIEW Development (2017) 144, 3417-3429 doi:10.1242/dev.148312
DEVELOPMENT
... Previous studies have shown the significance of the CST in driving functional recovery, while mTOR signaling stands out as one of the pivotal intrinsic regenerative pathways responsible for axonal regeneration and sprouting following various central nervous system (CNS) injuries [21,33,34]. Therefore, we investigated the activation level of mTOR signaling and functional recovery after treadmill training. ...
Article
Full-text available
In this study, we examined the combined impact of osteopontin (OPN) and treadmill training on mice with spinal cord injury (SCI). OPN was overexpressed by injecting AAV9-SPP1-GFP into the sensorimotor cortex, followed by a left incomplete C5 crush injury two weeks later. Mice (Ex or Ex + OPN group) were trained at 50% maximum running speed for 8 weeks. To analyze the effects, we used biotinylated dextran amine (BDA) for tracing the corticospinal tract (CST) and performed Western blotting and immunohistochemical methods to assess the activation of the mammalian target of rapamycin (mTOR). We also examined axonal regeneration and conducted behavioral tests to measure functional recovery. The results demonstrated that treadmill training promoted the expression of neurotrophic factors such as brain-derived neurotrophic factor (BNDF) and insulin-like growth factor I (IGF-1) and activated mTOR signaling. OPN amplified the effect of treadmill training on activating mTOR signaling indicated by upregulated phosphorylation of ribosomal protein S6 kinase (S6). The combination of OPN and exercise further promoted functional recovery and facilitated limited CST axonal regeneration which did not occur with treadmill training and OPN treatment alone. These findings indicate that OPN enhances the effects of treadmill training in the treatment of SCI and offer new therapeutic insights for spinal cord injury.
... Complex wiring of the mature CST, together with its central role in voluntary and fine motor control, means that damage leads to significant and lasting functional impairments. Efforts to repair the damaged CST have focused on either nullifying the effects of the axon growth inhibitory environment of the mature CNS or recapitulating cell autonomous developmental mechanisms (Schwab and Strittmatter, 2014;Gutilla and Steward, 2016;He and Jin, 2016;Hilton and Bradke, 2017;Bradbury and Burnside, 2019;Fawcett, 2020). Although strides have been made in stimulating axotomized CSNs to regenerate or intact CSNs to undergo plasticity after injury, recovery of significant motor function remains to be realized (Park et al., 2008;Fink et al., 2017;Kauer et al., 2022). ...
Article
Full-text available
The corticospinal tract (CST) forms a central part of the voluntary motor apparatus in all mammals. Thus, injury, disease, and subsequent degeneration within this pathway result in chronic irreversible functional deficits. Current strategies to repair the damaged CST are sub-optimal in part due to underexplored molecular heterogeneity within the adult tract. Here we combine spinal retrograde CST tracing with single-cell RNA sequencing in adult male and female mice to index corticospinal neuron (CSN) subtypes that differentially innervate the forelimb and hindlimb. We exploit publicly available datasets to confer anatomical specialization among CSNs and show that CSNs segregate not only along the forelimb and hindlimb axis but also by supraspinal axon collateralization. These anatomically defined transcriptional data allow us to use machine learning tools to build classifiers that discriminate between CSNs and cortical layer 2/3 and non-spinally terminating layer 5 neurons in M1, and separately identify limb specific CSNs. Utilizing these tools, CSN subtypes can be differentially identified to study postnatal patterning of the CST in vivo , leveraged to screen for novel limb-specific axon growth survival and growth activators in vitro , and ultimately exploited to repair the damaged CST after injury and disease. Significance Statement Therapeutic interventions designed to repair the damaged corticospinal tract (CST) after spinal cord injury have remained functionally sub-optimal in part due to an incomplete understanding of the molecular heterogeneity among subclasses of corticospinal tract neurons (CSNs). Here, we combine spinal retrograde labeling with scRNAseq and annotate a CSN index by the termination pattern of their primary axon in the cervical or lumbar spinal cord and supraspinal collateral terminal fields. Using machine learning we have confirmed the veracity of our CSN gene lists to train classifiers to identify CSNs among all classes of neurons in M1 to study the development, patterning, homeostasis, and response to injury and disease, and ultimately target streamlined repair strategies to this critical motor pathway.
... Axons are very thin nerve fibers that carry nerve impulses away from one neuron to another. Axonal outgrowth is essential for wiring the central nervous system during development [50]. Numerous studies found that axonal injury is a critical initiating event in a variety of central nervous system diseases, which is implicated in the neuropathology of depression, cognitive dysfunction, metabolic encephalopathies, and multiple sclerosis [51,52]. ...
Article
Full-text available
Background As one of the environmental risk factors for human health, atmospheric fine particulate matter (PM2.5) contributes to cognitive deterioration in addition to respiratory and cardiovascular injuries. Recently, increasing evidence implicates that PM2.5 inhalation can affect neurological functions in offspring, but the sex-specific outcomes and the underlying biological processes are largely unknown. Objectives To observe the influence of prenatal PM2.5 exposure on cognitive performance in offspring, to elucidate the neuronal morphological alterations and possible transcriptional regulation based on mRNA-sequencing (mRNA-Seq) data after birth, and to determine the key components of PM2.5 contributing to the adverse effects. Methods Pregnant C57BL/6J mice were exposed to sterile saline or PM2.5 suspension. Morris water maze test was used to assess the cognitive function in weanling offspring. Microscopic observation was applied to detect neuronal morphogenesis in vivo and in vitro. The cortex tissues from male offspring were collected on postnatal days (PNDs) 1, 7, and 21 for mRNA-Seq analysis. The organic and inorganic components of PM2.5 were separated to assess their contributions using primary cultured neurons. Results Prenatal PM2.5 exposure impaired spatial learning and memory in weanling male mice, but not female mice. The sex-specific outcomes were associated with mRNA expression profiles of the cortex during postnatal critical windows, and the annotations in Gene Ontology (GO) of differentially expressed genes (DEGs) revealed that the exposure persistently disrupted the expression of genes involved in neuronal features in male offspring. Consistently, axonal growth impairment and dendritic complexity reduction were observed. Importantly, Homeobox A5 (Hoxa5), a critical transcription factor regulating all of the neuronal morphogenesis-associated hub genes on PNDs 1, 7, and 21, significantly decreased in the cortex of male offspring following PM2.5 exposure. In addition, both inorganic and organic components were harmful to axonal and dendritic growth, with organic components exhibiting stronger inhibition than inorganic ones. Conclusion Prenatal PM2.5 exposure affected spatial learning and memory in male mice by disrupting Hoxa5-mediated neuronal morphogenesis, and the organic components, including polycyclic aromatic hydrocarbons (PAHs), posed more adverse effects than the inorganic components.
... Next, we identified groups of up-and down-regulated genes, which partially reverted their expression towards an embryonic RGC state, as well as those that are not developmentally-regulated ( Fig. 3C and Supplementary Data 1). Axon regeneration may require recapitulation of the molecular mechanisms involved in developmental embryonic axon growth (Filbin, 2006;Hilton and Bradke, 2017;Yaniv et al., 2012;Harel and Strittmatter, 2006;Goldberg et al., 2002;Chen, Jhaveri and Schneider, 1995;Poplawski et al., 2020;Moore et al., 2009). However, the adult CNS environment is different from embryonic and is further altered by lesion. ...
Article
Full-text available
Central nervous system (CNS) projection neurons fail to spontaneously regenerate injured axons. Targeting developmentally regulated genes in order to reactivate embryonic intrinsic axon growth capacity, or targeting pro-growth tumor suppressor genes such as Pten, promotes long-distance axon regeneration in only a small subset of injured retinal ganglion cells (RGCs), despite many RGCs regenerating short-distance axons. A recent study identified αRGCs as the primary type that regenerates short-distance axons in response to Pten inhibition, but the rare types which regenerate long-distance axons, and cellular features that enabled such response, remained unknown. Here, we used a new method for capturing specifically the rare long-distance axon-regenerating RGCs, and also compared their transcriptomes to embryonic RGCs, in order to answer these questions. We found the existence of adult non-α intrinsically-photosensitive (ip) M1 RGC subtypes that retained features of embryonic cell state, and showed that these subtypes partially dedifferentiated towards an embryonic state and regenerated long-distance axons in response to Pten inhibition. We also identified Pten inhibition-upregulated mitochondria-associated genes, Dynlt1a and Lars2, which promote axon regeneration on their own, and thus present novel therapeutic targets.
Article
Full-text available
Spinal cord injury (SCI) is a devastating neurological disorder, leading to loss of motor or somatosensory function, which is the most challenging worldwide medical problem. Re-establishment of intact neural circuits is the basis of spinal cord regeneration. Considering the crucial role of electrical signals in the nervous system, electroactive bioscaffolds have been widely developed for SCI repair. They can produce conductive pathways and a pro-regenerative microenvironment at the lesion site similar to that of the natural spinal cord, leading to neuronal regeneration and axonal growth, and functionally reactivating the damaged neural circuits. In this review, we first demonstrate the pathophysiological characteristics induced by SCI. Then, the crucial role of electrical signals in SCI repair is introduced. Based on a comprehensive analysis of these characteristics, recent advances in the electroactive bioscaffolds for SCI repair are summarized, focusing on both the conductive bioscaffolds and piezoelectric bioscaffolds, used independently or in combination with external electronic stimulation. Finally, thoughts on challenges and opportunities that may shape the future of bioscaffolds in SCI repair are concluded.
Chapter
The second chapter focuses on spinal cord injuries (SCIs). In addition to the biological and clinical basics, the differences to peripheral nerve regeneration are discussed, including the causes of axonal growth inhibition in the adult CNS. The current treatment approaches in the acute and chronic phase following SCIs are presented, and novel therapies using biopolymers, stem cell transplants, electrical stimulation, bioprostheses, and rehabilitation measures are introduced.
Chapter
Axonal regrowth is not possible in the adult mammalian brain due to unsupportive gene expression and microenvironment. Remyelination is different. Recovery of neuronal functions is aided by the secretion of neurotrophic factors from reactive astrocytes. Uninjured and intact axons can sprout and establish new synapses. Most prominent are reparative axonal sprouting and long-distance sprouting. It occurs over long distances and connects functionally related areas. It is therefore most important for functional recovery and compensation. Reactive astrocytes are involved as they repulse and promote the sprouting. In other words, they serve as guides that direct the sprouting axons to their targets. Blocking this astrocytic activity does not lead to functional recovery. This important role of astrocytes pertains only to moderate reactive subtypes but not to scar-forming astrocytes. In addition, reactive astrocytes are known to regulate spine stabilization, plasticity, and maturation of the newly sprouted axons with target dendrites by releasing synaptogenic factors in the appropriate areas. Macrophages can induce a proinflammatory shift in microglia after an injury. This can lead to chronic inflammation which is completely detrimental to the healing process. If this macrophage influence is absent microglia shift after a short pro-inflammatory phase into an anti-inflammatory phase and assist the astrocytes in restoring function by degrading fibrotic tissue and releasing of factors that promote sprouting and synaptogenesis.
Article
Spinal cord injury (SCI) is devastating, with limited treatment options and variable outcomes. Most in vivo SCI research has focused on the acute and early post-injury periods, and the promotion of axonal growth, so little is understood about the clinically stable chronic state, axonal growth over time, and what plasticity endures. Here, we followed animals into the chronic phase following SCI, to address this gap. Male macaques received targeted deafferentation, affecting three digits of one hand, and were divided into short (4–6 months) or long-term (11–12 months) groups, based on post-injury survival times. Monkeys were assessed behaviorally, where possible, and all exhibited an initial post-injury deficit in manual dexterity, with gradual functional recovery over 2 months. We previously reported extensive sprouting of somatosensory corticospinal (S1 CST) fibers in the dorsal horn in the first five post-injury months. Here, we show that by 1 year, the S1 CST sprouting is pruned, with the terminal territory resembling control animals. This was reflected in the number of putatively “functional” synapses observed, which increased over the first 4–5 months, and then returned to baseline by 1 year. Microglia density also increased in the affected dorsal horn at 4–6 months and then decreased, but did not return to baseline by 1 year, suggesting refinement continues beyond this time. Overall, there is a long period of reorganization and consolidation of adaptive circuitry in the dorsal horn, extending well beyond the initial behavioral recovery. This provides a potential window to target therapeutic opportunities during the chronic phase.
Preprint
In adult mammalian central nervous system (CNS) neurons, axon regeneration after injury remains limited due to the preset unfavorable gene regulatory programs. Factors enabling comprehensive epigenetic and transcriptional shifts, for example, the pivotal transcription factors which mediate the neurogenesis and morphogenesis may be essential for driving CNS axon regeneration. Based on the analysis of multiple public whole-genome mRNA and chromatin accessibility sequencing datasets of CNS neuronal development, as well as previous functional studies on the regeneration-capable dorsal root ganglion peripheral neurons, we hypothesize that overexpression of the pioneer transcription factor Achaete-Scute homolog 1 (Ascl1) will promote axon regeneration in the adult mammalian CNS neurons. We employed the optic nerve crush (ONC) in mice, a common model for studying CNS axon regeneration, neuron survival and glaucoma, to investigate the effect of Ascl1 overexpression. We found that Ascl1 could sufficiently promote axon regeneration past the crush site and significantly preserve the survival of retinal ganglion cells. Mechanistically, we revealed that the effect of Ascl1 was mediated by known pro-regeneration factor Sox11 but not others. Together, our study established an effective pipeline for discovering functionally important novel target for promoting CNS neuron axon regeneration and survival.
Article
Full-text available
Neural networks are emerging as the fundamental computational unit of the brain and it is becoming progressively clearer that network dysfunction is at the core of a number of psychiatric and neurodegenerative disorders. Yet, our ability to target specific networks for functional or genetic manipulations remains limited. Monosynaptically restricted rabies virus facilitates the anatomical investigation of neural circuits. However, the inherent cytotoxicity of the rabies largely prevents its implementation in long-term functional studies and the genetic manipulation of neural networks. To overcome this limitation, we developed a self-inactivating ΔG-rabies virus (SiR) that transcriptionally disappears from the infected neurons while leaving permanent genetic access to the traced network. SiR provides a virtually unlimited temporal window for the study of network dynamics and for the genetic and functional manipulation of neural circuits in vivo without adverse effects on neuronal physiology and circuit function.
Article
Full-text available
Chondroitin sulfate (CS) proteoglycans in perineuronal nets (PNNs) from the central nervous system (CNS) are involved in the control of plasticity and memory. Removing PNNs reactivates plasticity and restores memory in models of Alzheimer's disease and ageing. Their actions depend on the glycosaminoglycan (GAG) chains of CS proteoglycans, which are mainly sulfated in the 4 (C4S) or 6 (C6S) positions. While C4S is inhibitory, C6S is more permissive to axon growth, regeneration and plasticity. C6S decreases during critical period closure. We asked whether there is a late change in CS-GAG sulfation associated with memory loss in aged rats. Immunohistochemistry revealed a progressive increase in C4S and decrease in C6S from 3 to 18 months. GAGs extracted from brain PNNs showed a large reduction in C6S at 12 and 18 months, increasing the C4S/C6S ratio. There was no significant change in mRNA levels of the chondroitin sulfotransferases. PNN GAGs were more inhibitory to axon growth than those from the diffuse extracellular matrix. The 18-month PNN GAGs were more inhibitory than 3-month PNN GAGs. We suggest that the change in PNN GAG sulfation in aged brains renders the PNNs more inhibitory, which lead to a decrease in plasticity and adversely affect memory.
Article
Full-text available
Remyelination is limited in the majority of multiple sclerosis (MS) lesions despite the presence of oligodendrocyte precursor cells (OPCs) in most lesions. This observation has led to the view that a failure of OPCs to fully differentiate underlies remyelination failure. OPC differentiation requires intricate transcriptional regulation, which may be disrupted in chronic MS lesions. The expression of few transcription factors has been differentially compared between remyelinating lesions and lesions refractory to remyelination. In particular, the oligodendrocyte transcription factor myelin regulatory factor (MYRF) is essential for myelination during development, but its role during remyelination and expression in MS lesions is unknown. To understand the role of MYRF during remyelination, we genetically fate mapped OPCs following lysolecithin-induced demyelination of the corpus callosum in mice and determined that MYRF is expressed in new oligodendrocytes. OPC-specific Myrf deletion did not alter recruitment or proliferation of these cells after demyelination, but decreased the density of new glutathione S-transferase π positive oligodendrocytes. Subsequent remyelination in both the spinal cord and corpus callosum is highly impaired following Myrf deletion from OPCs. Individual OPC-derived oligodendrocytes, produced in response to demyelination, showed little capacity to express myelin proteins following Myrf deletion. Collectively, these data demonstrate a crucial role of MYRF in the transition of oligodendrocytes from a premyelinating to a myelinating phenotype during remyelination. In the human brain, we find that MYRF is expressed in NogoA and CNP-positive oligodendrocytes. In MS, there was both a lower density and proportion of oligodendrocyte lineage cells and NogoA+ oligodendrocytes expressing MYRF in chronically demyelinated lesions compared to remyelinated shadow plaques. The relative scarcity of oligodendrocyte lineage cells expressing MYRF in demyelinated MS lesions demonstrates, for the first time, that chronic lesions lack oligodendrocytes that express this necessary transcription factor for remyelination and supports the notion that a failure to fully differentiate underlies remyelination failure.
Article
Full-text available
Themammaliantarget of rapamycin(mTOR)positively regulatesaxongrowth in themammaliancentral nervous system (CNS). Althoughaxon regeneration and functional recovery from CNS injuries are typically limited, knockdown or deletion of PTEN, a negative regulator of mTOR, increases mTOR activity and induces robust axon growth and regeneration. It has been suggested that inhibition of S6 kinase 1 (S6K1, gene symbol: RPS6KB1), a prominent mTOR target, would blunt mTOR’s positive effect on axon growth. In contrast to this expectation, we demonstrate that inhibition of S6K1 inCNSneurons promotes neurite outgrowth in vitro by twofold to threefold. Biochemical analysis revealed that an mTOR-dependent induction of PI3K signaling is involved in mediating this effect of S6K1 inhibition. Importantly, treating female mice in vivo with PF-4708671, a selective S6K1 inhibitor, stimulated corticospinal tract regeneration across a dorsal spinal hemisection between the cervical 5 and 6 cord segments (C5/C6), increasing axon counts for at least 3mmbeyond the injury site at 8 weeks after injury. Concomitantly, treatment withPF-4708671producedsignificantlocomotorrecovery. Pharmacological targeting ofS6K1maytherefore constituteanattractive strategy for promoting axon regeneration following CNS injury, especially given that S6K1 inhibitors are being assessed in clinical trials for nononcological indications.
Article
Full-text available
Neural progenitor cells grafted to sites of spinal cord injury have supported electrophysiological and functional recovery in several studies. Mechanisms associated with graft-related improvements in outcome appear dependent on functional synaptic integration of graft and host systems, although the extent and diversity of synaptic integration of grafts with hosts are unknown. Using transgenic mouse spinal neural progenitor cell grafts expressing the TVA and G-protein components of the modified rabies virus system, we initiated monosynaptic tracing strictly from graft neurons placed in sites of cervical spinal cord injury. We find that graft neurons receive synaptic inputs from virtually every known host system that normally innervates the spinal cord, including numerous cortical, brainstem, spinal cord, and dorsal root ganglia inputs. Thus, implanted neural progenitor cells receive an extensive range of host neural inputs to the injury site, potentially enabling functional restoration across multiple systems.
Article
Full-text available
The anticonvulsant pregabalin promotes neural regeneration in a mouse model of spinal cord injury (SCI). We have also previously observed that anticonvulsants improve motor outcomes following human SCI. The present study examined the optimal timing and type of anticonvulsants administered in a large, prospective, multi-center, cohort study in acute SCI. Mixed-effects regression techniques were used to model total motor scores at 1, 3, 6, and 12 months post injury. We found that early (not late) administration of anticonvulsants significantly improved motor recovery (6.25 points over 1 year). The beneficial effect of anticonvulsants remained significant after adjustment for differences in 1-month motor scores and injury characteristics. A review of a subset of patients revealed that gabapentinoids were the most frequently administrated anticonvulsant. Together with preclinical findings, intervention with anticonvulsants represents a potential pharmacological strategy to improve motor function after SCI.
Article
Neural progenitor cells (NPCs) grafted to sites of spinal cord injury (SCI) extend numerous axons over long distances and form new synaptic connections with host neurons. In the present study we examined the myelination of axons emerging from NPC grafts. Rat E14 multipotent NPCs constitutively expressing GFP were grafted into adult C5 spinal cord hemisection lesions; 3 months later we examined graft-derived axonal diameter and myelination using transmission electron microscopy. 104 graft-derived axons were characterized. Axon diameter ranged from 0.15 to 1.70 μm, and 24% of graft-derived axons were myelinated by host oligodendrocytes caudal to the lesion. The average diameter of myelinated axons (0.72 ± 0.06 μm) was significantly larger than that of non-myelinated axons (0.61 ± 0.03 μm, p < 0.05). Notably, the G-ratio of myelinated graft-derived axons (0.77 ± 0.01) was virtually identical to that of the normal, intact spinal cord described in published reports. These findings indicate that axons emerging from early stage neural grafts into the injured spinal cord recapitulate both the small/medium size range and myelin thickness of intact spinal cord axons.
Article
Multiple sclerosis is characterized by inflammatory activity that results in destruction of the myelin sheaths that enwrap axons. The currently available medications for multiple sclerosis are predominantly immune-modulating and do not directly promote repair. White matter regeneration, or remyelination, is a new and exciting potential approach to treating multiple sclerosis, as remyelination repairs the damaged regions of the central nervous system. A wealth of new strategies in animal models that promote remyelination, including the repopulation of oligodendrocytes that produce myelin, has led to several clinical trials to test new reparative therapies. In this Review, we highlight the biology of, and obstacles to, remyelination. We address new strategies to improve remyelination in preclinical models, highlight the therapies that are currently undergoing clinical trials and discuss the challenges of objectively measuring remyelination in trials of repair in multiple sclerosis.
Article
Central nervous system (CNS) injury transforms naive astrocytes into reactive astrocytes, which eventually become scar-forming astrocytes that can impair axonal regeneration and functional recovery. This sequential phenotypic change, known as reactive astrogliosis, has long been considered unidirectional and irreversible. However, we report here that reactive astrocytes isolated from injured spinal cord reverted in retrograde to naive astrocytes when transplanted into a naive spinal cord, whereas they formed astrocytic scars when transplanted into injured spinal cord, indicating the environment-dependent plasticity of reactive astrogliosis. We also found that type I collagen was highly expressed in the spinal cord during the scar-forming phase and induced astrocytic scar formation via the integrin-N-cadherin pathway. In a mouse model of spinal cord injury, pharmacological blockade of reactive astrocyte-type I collagen interaction prevented astrocytic scar formation, thereby leading to improved axonal regrowth and better functional outcomes. Our findings reveal environmental cues regulating astrocytic fate decisions, thereby providing a potential therapeutic target for CNS injury.
Article
Spinal cord injury can lead to severe motor, sensory and autonomic dysfunction. Currently, there is no effective treatment for the injured spinal cord. The transplantation of Schwann cells, neural stem cells or progenitor cells, olfactory ensheathing cells, oligodendrocyte precursor cells and mesenchymal stem cells has been investigated as potential therapies for spinal cord injury. However, little is known about the mechanisms through which these individual cell types promote repair and functional improvements. The five most commonly proposed mechanisms include neuroprotection, immunomodulation, axon regeneration, neuronal relay formation and myelin regeneration. A better understanding of the mechanisms whereby these cells promote functional improvements, as well as an appreciation of the obstacles in implementing these therapies and effectively modeling spinal cord injury, will be important to make cell transplantation a viable clinical option and may lead to the development of more targeted therapies.