Open access peer-reviewed chapter

Schwann Cell Plasticity in Peripheral Nerve Regeneration after Injury

Written By

Emilia Manole, Alexandra Eugenia Bastian, Ana Maria Oproiu, Monica Teodora Neagu, Carolina Constantin and Gheorghita Isvoranu

Submitted: 30 January 2020 Reviewed: 18 February 2020 Published: 03 April 2020

DOI: 10.5772/intechopen.91805

From the Edited Volume

Demyelination Disorders

Edited by Stavros J. Baloyannis, Fabian H. Rossi and Welwin Liu

Chapter metrics overview

776 Chapter Downloads

View Full Metrics

Abstract

In the normal peripheral nervous system, Schwann cells (SCs) are present in two different states of differentiation: myelinating SCs that surround large-caliber axons, forming myelin sheath, and non-myelinating SCs that surround more small-caliber axons forming Remak bundles. Under pathological conditions (injury or inflammation), SCs, with a remarkable plasticity, undergo phenotypic transformations, downregulating the production of myelin proteins mRNAs, upregulating neurotrophic factors and cytokines, thus promoting the axonal regeneration. Dedifferentiated SCs activate the protein degradation, participating in the demyelination process and clearance of myelin debris; attract macrophages helping wound healing; proliferate to replace lost cells; guide axonal growth; and protect against secondary axonal damage. Thus, SC functions have a critical contribution to regeneration processes that occur in peripheral nerve after injury.

Keywords

  • Schwann cell plasticity
  • dedifferentiated Schwann cells
  • peripheral nerve regeneration
  • myelin recovery

1. Introduction

Schwann cells (SCs) are glial cells present in the peripheral nerve system (PNS). The name was given in honor of the German scientist Theodore Schwann, who discovered them in the nineteenth century [1] although they were not the main subject of his research. At that time it was thought that this type of cells is very complex and that the cells merge to supply peripheral nerves. Ramon y Cajal, only about 100 years later, discovered the true structure of the peripheral nerves, composed of axons and SCs that are in a symbiotic connection with it [2]. In the following years, with the evolution of electron microscopy, the study of SC morphology has developed continuously, leading to a better understanding of their complex biology.

It is known that nerves in PNS are much easier to regenerate than those in the central nervous system (CNS). Ramon y Cajal sensed very well that there is a “symbiosis” between the axon and the Schwann cells. Kidd et al. [3] described the Schwann cell as one of the largest and most complex cells in the body, which can develop and evolve rapidly after injury. The origin of the Schwann cell is in the neural crest, and this differentiation is made by the regulation of Sox10 but also in the presence of Notch and endothelin signaling [4, 5].

After a peripheral nerve lesion, a series of cellular changes occur at both axons and Schwann cells, a phenomenon known as Wallerian degeneration: axonal degeneration and myelin destruction, followed by a dedifferentiation (an immature-like phenotype of SCs) and proliferation of Schwann cells [6].

The purpose of this chapter is to highlight the extremely important role of the Schwann cell in the regeneration of the peripheral nerve and its extraordinary plasticity in order to ensure this phenomenon.

Advertisement

2. Peripheral nerve injury

What does peripheral nerve injury mean? This could mean a mechanical trauma, transection or crush, or a pathological condition, when could be affected sensory nerves, motor nerves or autonomic nerves. A peripheral neuropathy may affect one or many nerves, axon, or myelin in the first stage.

In the nerve transection, all nerve fibers are affected, while in a disease manifestation, only a number of nerve fibers are affected, others being normal (Figures 2A and 4).

Very briefly, in peripheral neuropathies, it may be an axonal primary damage or a myelin sheath primary damage. After a period both components of the nerve fiber are affected.

Primary axonal degeneration, whether it is nerve transection or a pathological manifestation, is essentially the same: it starts with a Wallerian degeneration in the distal part of nerve (Figure 1), following the myelin destruction. On semithin transverse sections (Figure 2B and C) and in electron microscopy images (Figure 3), the affected nerve fibers are seen to be in a process of necrobiosis. In electron microscopy images, autophagic vacuoles are seen, near the axon (Figure 3A) or in the exterior layer of SC, under the basal lamina (Figure 3B and C) and macrophages (Figure 3D). After the destruction of the nerve fiber, only irregular structures of myelin residues can be seen (Figure 3E) or myelin debris like ovoids and balls (Figure 4B and C). If it is a chronic process, many nerve fibers disappear, the density of myelinated fibers being very low (Figure 2D and E). When the myelin is affected in the first step, not all Schwann cells are suffering in the same time. One internode with a very thin sheath between two normal internodes may be observed: segmental demyelination (Figure 4A and B). When a myelin protein, PMP22, is genetically affected, in hereditary neuropathy with pressure palsies (HNPP), the nerve biopsy shows demyelination and focal hypermyelination structures, tomacula (sausage-like) (Figures 2F and 4D). In hypertrophic neuropathies, like Charcot-Marie-Tooth disease type 1A (CMT 1A) and chronic inflammatory demyelinating polyneuropathy (CIDP), some structures named “onion bulbs” are present, a result of concentric layers of Schwann cell processes and collagen around the axons (Figure 2E). It is a repetitive segmental demyelination and myelin regeneration.

Figure 1.

Wallerian degeneration. After injury, axon and myelin sheath in the distal stump degenerate. Macrophages migrate to the site of lesion and with proliferating Schwann cells remove myelin debris. After the debris has been removed, dedifferentiated Schwann cells align forming bands of Bungner, guiding axonal sprout regeneration.

Figure 2.

Peripheral nerve pathological modifications (sural nerve biopsy): (A) a very mild affected nerve, with a normal fiber density; some myelinated fibers with small and medium mean diameter with demyelination; (B) a severe axonal destruction, with disappearance of many large diameter axons and with a low-fiber density; a degenerated axon is present; (C) many degenerated axons and demyelination present in the rest of myelinated fibers; (D) a very severe neuropathy with disappearance of most of the myelinated fibers; (E) some small myelinated axons with onion bulbs; (F) a hypermyelinated fiber in an HNPP case (tomacula) in the center of the image; (G) regeneration aspect: cluster of small axons (arrow). Semithin cross sections stained with toluidine blue; (under oil immersion – 60× Objective).

Figure 3.

Electron microscopy aspects of axonal degeneration (sural nerve biopsy). (A) A myelinated axon showing an autophagic vacuole between axon and myelin sheath. (B) A myelinated axon with an autophagic vacuole in the Schwann cell exterior cytoplasma: small myelin debris are seen. (C) The same aspect: an autophagic Schwann cell with many smaller or bigger fragments of myelin inside. (D) A macrophage with lipid droplets is present near myelinated axons. (E) Total myelin degradation; only irregular laminated structure is present, with no axon (cross sections; bar = 2 μm).

Figure 4.

Teased nerve fiber (sural nerve biopsy) panel. (A) A nerve fiber with segmental demyelination near two other normal myelinated fibers. (B) Near normal fibers, a fiber with segmental demyelination (a thin internode) and a fiber with few myelin ovoids and balls (axonal degeneration). (C) More myelin ovoids in an axonal degeneration. (D) A tomacula in myelin sheath of a nerve fiber.

After a segmental demyelination, along the affected internode, several Schwann cells arrive which begin to remyelinate this portion, the sign of remyelination being more short internodes (Figure 5).

Figure 5.

Peripheral nerve remyelination. In demyelinating peripheral neuropathies, the segmental demyelination is often seen. Following a Schwann cell degeneration, the lost myelin internode is replaced by some Schwann cells which generate myelin sheaths, resulting in many shorter internodes.

The sign of axonal regeneration is observed on semithin sections and consists of the presence of some clusters of axons with the same small mean diameter and thinner myelin sheath (Figure 2G).

After these images sowing just few aspects of pathological degradation of peripheral nerve, focusing on myelin sheath damage, let’s take a closer look at what happens in the Schwann cell, at the cellular and molecular level.

Advertisement

3. Myelin protein gene expression in peripheral nerve after injury

Investigating the evolution of the main proteins that enter the composition of myelin sheath during and after nerve injury has been a subject of study for many scientists. These proteins are P zero (P0), myelin basic protein (MPZ), and P2. The first two play an important role in maintaining the integrity and compactness of the myelin sheath. P2 is a lipid-binding protein and participates in fatty acid elongation and transport during the myelination process [7]. Myelin associated glycoprotein (MAG) is a transmembrane protein that is found in the periaxonal region and participates in SC-axon contact organization. It seems to be involved in the myelination process after injury [8]. P0 and MBP mRNA in the distal nerve portion after transection were found to be 20% lower than normal levels but have had normal levels after crushing [9, 10]. In the absence of a contact between SCs and the axon, the levels of mRNAs of P0 and MBP remained low, and mRNA of MAG was undetectable, long time after nerve transection, whereas MAG mRNA was undetectable after lesion; in the case of a crush injury, after a sudden and short decrease, the mRNA levels of these proteins were found to increase rapidly afterwards [10, 11].

Advertisement

4. Biological aspects of Schwann cell

To understand what plasticity of Schwann cells means, we need to understand what the starting point is for their differentiation and evolution.

4.1 Schwann cell differentiation and development

During development, SCs surround bundles of axons and support them to outgrow by releasing growth factors such as nerve growth factor (NGF), glial cell line-derived neurotrophic factor (GDNF), brain-derived neurotrophic factor (BDNF), and neurotrophin (NT3) [12, 13, 14]. It follows a “radial sorting” of axons by extension of cellular process from Schwann cells, which begins to divide axon bundles into smaller ones and finally separate the neighboring axons with cell cytoplasm. Thus two types of fibers are formed: (i) unmyelinated Remak fibers, in which SC surrounds several small-sized axons (sensory and autonomic) and does not produce myelin, and (ii) myelinated fibers in which each large-sized axon is surrounded by a SC cell, 1:1 relationship, and a myelin sheath is formed by SC membrane spirally wrapping the axon [15]. Mesaxon is termed the point where the plasma membrane apposition is formed where the first encircling process meets itself. Remak SCs maintain the proliferative capacity of all the life [16].

During this stage, changes in cell morphology and gene expression occur, mediated by the transcription factor Krox-20 (or Egr-2) [17, 18, 19].

4.2 Interactions between Schwann cells and axons

The differentiation of Schwann cells is controlled by some growth factors among which the most important are in the neuregulin family. Neuregulins (Nrgs) are transmembrane proteins that signal through ErbB tyrosine kinase receptors [20]. Axonal neuregulin-1 (Nrg1), produced in many isoforms by alternative splicing (heregulin, glial growth factor, sensory and motor neuron-derived factor), interacts with ErbB2/ErbB3 receptors tyrosine kinase expressed on Schwann cells [21, 22, 23, 24, 25]. ErbB2 and ErbB3 combine to act as heterodimers and efficiently bind Nrg1. Nrg1/ErbB signaling axis has a critical role in Schwann cell development (for review [26, 27, 28]) like survival, proliferation, migration, differentiation, and myelination [26, 29, 30, 31, 32].

Nrgs need protease involvement for Nrg1-ErbB interactions because Nrgs are synthesized as single-pass transmembrane proteins and shed from the cell surface by the proteolytic cleavage, thus permitting the interaction with ErbB receptors across the periaxonal space [33, 34].

Another enzyme implicated in Nrg1 cleavage is beta-amyloid converting enzyme (BACE1), a beta-secretase present in axon [35, 36]. An in vivo study showed that the BACE1-null mice presented reduced rates of Nrg1 cleavage and decreased PNS myelin, a low capacity of myelination with axons with a thinner myelin sheath [35].

An effect opposite to the BACE activity has tumor necrosis factor-alpha-converting enzyme (TACE), a neuronal alpha-secretase, cleaving Nrg1 into an inactive form [37]. TACE genetic inactivation in motor neurons caused hypermyelination like in Nrg1 overexpression.

Another factor that is essential in SCs-axon interaction, with a protection role for the axon, is Schwann cell basal lamina. The basal lamina together with extracellular collagen fibrils protects axons from extension and compression injuries. They provide good support for axonal outgrowth and guidance (reviewed by [38]). Basal lamina defines also Schwann cell orientation in axonal myelination [39]. More of this, SCs require axonal contact for secreting the components of basal lamina, so the relationship of axon-SCs via basal lamina is interactive and reciprocal [40, 41].

All these interactions described above are very important and may be modulated in the control of nerve regeneration.

Advertisement

5. Schwann cell plasticity

PNS has a very good regenerative capacity, and this is largely due to Schwann cells that develop a high plasticity and can contribute very quickly to the regeneration of peripheral nerves after injury whether it is a trauma or a pathological condition. In these cases, SCs have the ability to transform into an immature-like form, which drives subsequent regeneration of the nerve. These processes of dedifferentiation into non-myelinating cells and redifferentiation after injury are characteristic of these glial cells in PNS, and in the last decade a significant progress has been made in the study of the molecular mechanisms and signaling pathways that regulate this plasticity (reviewed in [42]). More of this, the myelinating and non-myelinating SCs remain bipotential cells all the time, as demonstrated by grafting or nerve cross anastomosis experiments [43, 44, 45]. Many experimental studies on transgenic animals have shown that after nerve cut or crush, both types of SCs reprogram into proliferative progenitor-like repair SCs [46, 47]. This phenomenon involves downregulation of pro-myelinating genes, such as early growth response 2 (Egr-2 or Krox-20), POU domain class 3 transcription factor 1 (Pou3f1 or Oct-6), and myelin protein zero (MPZ)/myelin basic protein (MBP). There is also an upregulation of markers of dedifferentiated (immature) SCs like low affinity neurotrophin receptor (p75NTR), c-Jun, or glial fibrillary acidic protein (GFAP) [6].

After Wallerian degeneration following nerve injury, a downregulation of pro-myelinating genes occurs, and the myelin clearing phenomenon begins after myelin sheath disorganization, through a mechanism of autophagy or myelinophagy [48]. Macrophages also participate in this process, phagocytosing myelin and axonal debris. The recruitment of macrophages is also done by SCs [49, 50, 51].

One of the major problems of human SCs is that as their regenerative capacity decreases in time, they can no longer sustain axonal growth, and their numbers decrease greatly (reviewed in [52]).

Regarding the plasticity of Schwann cells, although not covered by this chapter, we just want to mention here that SC precursors can generate many and different cell types during embryogenesis, besides myelinating and non-myelinating SCs, such as endoneurial fibroblasts, melanocytes, and neurons [52].

5.1 Schwann cell dedifferentiation

After injury, SCs reacquire some capabilities from early development, like proliferation, production of growth factors, sorting, and myelination. A good review regarding the biology of Schwann cells is the one made by Kidd et al. [3].

SC behavior and fate is regulated by two sort of interactions: SCs-axon and SCs-extracellular matrix/basal matrix. After 48 hours following axonal transection, SCs downregulate the production of myelin protein mRNAs [53] and upregulate trophic factors and cytokines [12, 13, 14] like NGF, BDNF, GDNF, and LIF, molecules necessary in axonal regeneration promoting into distal stump (reviewed in [54]). After axonal injury/transection, the axon is rapidly destroyed by a nonapoptotic autonomous mechanism [55]. SCs begin myelin degradation after axon injury, disassembling first the myelin internode starting with Schmidt-Lantermann incisure swelling [56, 57], following the dissolution of myelin in bubbles, ovoids, and balls. Macrophages finish the myelin degradation by phagocytosis [58]. It is not known exactly how much the SCs contribute to myelin degradation compared to macrophage participation, but it seems that it depends on the volume of the internode [59, 60]. During myelin degeneration, changes occur in the SC microtubule network, lysosome, and endosome positioning [61].

After nerve crush or transection, between the two stumps, over the lesion site, fibroblasts form a bridge, interacting with SCs [62]. The newly formed vasculature participates also in guiding the growing axons through this bridge to the distal end [63]. After a period of persistence of distal nerve stumps, distal axons disappear and dedifferentiated SCs proliferate, align, and begin emitting processes, forming the bands of Bungner (Figure 1), offering a physical and trophic support for the regrowth of axon [44, 60].

After the axonal regeneration, SCs differentiate once more in non-myelinating and myelinating cells to finish the functional recovery of the nerve. The regenerated myelin internodes (Figure 5) are shorter and thinner than the rest of the original ones in the proximal part of nerve [64].

5.2 Molecular mechanisms which control SC plasticity

The molecular mechanisms that regulate SC plasticity are very complex and widely described in many studies in recent years (reviewed in [42]). Here we will briefly mention them.

5.2.1 Transcriptional factors

One important transcriptional factor in SC reprogramming is c-Jun. Although it is downregulated or absent in the differentiation of SC, under pathological conditions c-Jun is particularly upregulated as described in various peripheral neuropathies [65, 66, 67, 68, 69], being a cross-antagonist of Krox-20 (a pro-myelinating transcription factor). c-Jun take part at the myelinophagy process [47] and participate also in the macrophage recruitment following nerve injury [70].

Another transcriptional regulator is NICD, an intracellular domain generated from neurogenic locus notch homolog protein (Notch) cleavage. SC proliferation and generation of immature SCs are controlled by Notch. But the same Notch is a negative regulator of myelination [71].

Nuclear factor kB (NF-kB), a transcription factor which regulates many physiological processes especially the inflammatory response, is very important for SC differentiation and myelination as in vitro studies showed [72, 73, 74].

In the recent years, a transcriptional repressor, Zeb2, has been investigated, and the researchers showed that it is implied in SC differentiation and myelination. The lack of Zeb2 in SCs results in a failure of SC maturation and in absence of myelin membranes [75].

Other factors which are overexpressed in SC dedifferentiation are Sox-2, paired box protein 3 (Pax-3), early growth response proteins 1 and 3 (Egr-1 and Egr-3), and DNA-binding protein inhibitor 2 (Id2) [66, 76, 77]. Sox-2 is also necessary for the nerve bridge formation after nerve injury [62].

mTOR complex 1 (mTORC1) (reviewed in [78]) has a significant role on the transcriptome by controlling transcription factors [79, 80, 81, 82]. It promotes anabolism, counting mRNA translation, and purine and pyrimidine synthesis [83, 84]. mTORC1 is necessary in radial sorting of axons by SCs, biosynthesis of lipids, and, on this basis, myelin growth [85, 86]. The mTORC1 activity is higher before myelination onset and decreases when myelination starts [87, 88, 89].

5.2.2 Mitogen-activated protein kinase (MAPK) family proteins

In the distal stump of the peripheral nerve after injury SCs respond by activating MAPK proteins like extracellular signal-regulated kinase (Erk), c-Jun N-terminal kinase (JNK), and p38 MAP kinase [66, 90, 91, 92, 93, 94, 95].

Ras/Raf/Erk signaling in SC dedifferentiation was studied for the first time by Harrisingh et al., and they showed that the Raf activation suppresses the differentiation of primary SCs induced by cyclic adenosine monophosphate (cAMP) [91]. Raf is an activator of Erk. The authors demonstrated that the activation of Ras/Raf/Erk pathway induced demyelination in an in vitro study on cocultured cells—SCs and neurons from dorsal root ganglia.

Erk activation is a pro-myelinating factor, and if Erk is inhibited, the SC differentiation and myelination are blocked, showed many in vivo studies [96, 97, 98].

In conclusion, Erk signaling is required in differentiation (Erk low levels) but also in dedifferentiation (high Erk levels) of SCs after nerve lesion [99, 100].

JNK, another MAPK protein, is implied in SC functions, so when c-Jun is activated by JNK, the migration and proliferation of SCs are produced [19, 101, 102].

Without insisting, we would like just to remember other MAPK proteins and signaling pathways involved in SC plasticity: p38MAPK, PI3K/Akt/mTOR signaling (reviewed by [42]).

5.2.3 TLRs signaling

After nerve injury, inflammation is an important phenomenon that must be considered. Thus, Toll-like receptors (TLRs) are key factors in initiating the immune response. A number of such receptors are expressed by SCs: TLR3, TLR4, and TLR7 [103]. Some experimental studies showed an upregulation of TLRs following nerve injury, the effect being the inflammation trigger with macrophage recruitment and activation and myelin clearance via SCs [50, 104, 105].

5.2.4 Nrg1/ErbB2/3 signaling

SCs express receptors for axonal neuregulins, as it is showed in Ssection 2.2. The neuregulin/Erb2/3 signaling is strongly involved in immature SC development but not in the regulation of adult SC proliferation after injury. An in vivo study on erbB2 wt/lacZ (with highly reduced ErbB2 levels in adult sciatic nerves) mice showed that after sciatic nerve transection, SC proliferation is not affected in adult ErbB2-conditional null nerves. More of this, the maintenance of myelinated peripheral nerves did not require ErbB2 function [106]. Other studies demonstrated that ErbB2 activation after sciatic nerve axotomy induced SC demyelination [107].

Neuregulin Nrg1 is still necessary for adult SC evolution after nerve injury [108, 109]. The absence of Nrg1 in adult axons results in remyelination defects after nerve crush experiments and also in a slower axon regeneration [110].

Advertisement

6. Therapeutical approaches based on Schwann cell plasticity

Although the peripheral nerve has a much greater regenerative capacity than the CNS nerve, the clinical recovery of patients with peripheral neuropathies is difficult, slow and often incomplete. Moreover, this capacity decreases with age.

The rate of nerve regeneration is approximately 1 mm/day, depending on the site of the lesion and on the patient age. SC plasticity diminishes with age, showing an altered expression of c-Jun [111] and a weak regenerative capacity [112, 113].

Understanding the signaling pathways that govern SC reprogramming and plasticity is essential for nerve repair and therapy.

For example, modulating Nrg1/ErbB signaling may improve myelin clearance, axonal regeneration, and finally functional nerve recovery after injury. An inappropriate overactivation of this pathway may lead to demyelinating neuropathies or tumors like neuroepithelioma and neuroplastic SC line [114, 115]. Experiments on transgenic mice with overexpression of Nrg1 showed hypertrophic neuropathies and malignant peripheral nerve sheath tumors [116]. The excessive activation of ErbB2 by Mycobacterium leprae determines one of the symptoms of leprosy, an important peripheral nerve demyelination [117]. In Charcot-Marie-Tooth 1A, abnormal demyelination and axon loss were prevented by Nrg1 therapy during early postnatal period in a rat model [118].

Another approach to stimulate SC regeneration and peripheral nerve functional recovery is the exogenous modulation by electric stimulation with low frequencies, photomodulation with low-level laser, and pharmacotherapy (with pharmacological agents, growth factors, bioproducts, or hormones) (reviewed by [119]).

Advertisement

7. Conclusions

Understanding Schwann cell biology and its extraordinary plasticity can lead to the development of new therapeutic approaches in peripheral nerve pathology and in the improvement of treatment methods in the case of traumatic nerve lesions. Peripheral neuropathies cause a significant morbidity and a decreased life quality. A better understanding of the many SC signaling pathways represents a very important approach for nerve regeneration as long as we have seen that SC is the main engine in nerve damage and repair after injury.

The recovery of the peripheral nerve, although better than that of the CNS nerve, is still quite complicated, difficult many times, and it is never perfect until the end. But in the last years, a huge amount of scientific data drew attention to the role of growth factors, transcriptional factors, inflammatory factors, hormones, and even exogenous modulation factors in the regulation of Schwann cell and of Schwann cell-axon interrelations, a complex integrated system.

It is expected that studies regarding SCs plasticity in peripheral nerve regeneration will continue and expand, improving not only the scientific knowledge but also a targeted more effective therapies, based on the pathology, personalized treatment and specific response of patients.

Advertisement

Acknowledgments

This work was supported by the Ministry of Education and Research in Romania, under the following grants: Program 1, The improvement of the National System of Research and Development and Subprogram 1.2—Institutional Excellence—Projects of Excellence Funding in RDI, Contract nr. 7PFE/16.10.2018; PN 19.29.02.01; by UEFISCDI, grant, Project PN-III-P1-1.2-PCCDI-2017-0782 (REGMED); and by COST 18125 AeroGels.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Schwann TH. Microscopical Researches into the Accordance in the Structure and Growth of Animals and Plants. Vol. 1. London: Sydenham Society; 1847. DOI: 10.1002/j.1550-8528.1993.tb00021.x
  2. 2. Ramón Y, Cajal S, DeFelipe J, Jones EG, May RM. Cajal’s Degeneration and Regeneration of the Nervous System. Oxford, England: Clarendon Press; 2012. DOI: 10.1093/acprof:oso/9780195065169.001.0001
  3. 3. Kidd GJ, Ohno N, Trapp BD. Biology of Schwann cells. In: Said G and Krarup C, editors. Handbook of Clinical Neurology. 1st ed. Vol. 115. Amsterdam: Elsevier B.V. 2013. pp. 55-79. DOI: 10.1016/B978-0-444-52902-2.00005-9
  4. 4. Brennan A, Dean CH, Zhang AL, Cass DT, Mirsky R, Jessen KR. Endothelins control the timing of Schwann cell generation in vitro and in vivo. Developmental Biology. 2000;227(2):545-557. DOI: 10.1006/dbio.2000.9887
  5. 5. Wakamatsu Y, Maynard TM, Weston JA. Fate determination of neural crest cells by NOTCH-mediated lateral inhibition and asymmetrical cell division during gangliogenesis. Development. 2000;127(13):2811-2821
  6. 6. Jessen KR, Mirsky R. Negative regulation of myelination: Relevance for development, injury, and demyelinating disease. Glia. 2008;56(14):1552-1565. DOI: 10.1002/glia.20761
  7. 7. Narayanan V, Barbosa E, Reed R, Tennekoon G. Characterization of a cloned cDNA encoding rabbit myelin P2 protein. The Journal of Biological Chemistry. 1988;263(17):8332-8337
  8. 8. Quarles RH. Myelin-associated glycoprotein in development and disease. Developmental Neuroscience. 1983;6(6):285-303. DOI: 10.1159/000112356
  9. 9. Leblanc AC, Poduslo JF, Mezei C. Gene expression in the presence or absence of myelin assembly. Molecular Brain Research. 1987;2(1):57-67. DOI: 10.1016/0169-328X(87)90021-0
  10. 10. Gupta SK, Poduslo JF, Mezei C. Temporal changes in PO and MBP gene expression after crush-injury of the adult peripheral nerve. Molecular Brain Research. 1988;4(2):133-141. DOI: 10.1016/0169-328X(88)90005-8
  11. 11. LeBlanc AC, Poduslo JF. Axonal modulation of myelin gene expression in the peripheral nerve. Journal of Neuroscience Research. 1990;26(3):317-326. DOI: 10.1002/jnr.490260308
  12. 12. Meyer M, Matsuoka I, Wetmore C, Olson L, Thoenen H. Enhanced synthesis of brain-derived neurotrophic factor in the lesioned peripheral nerve: Different mechanisms are responsible for the regulation of BDNF and NGF mRNA. The Journal of Cell Biology. 1992;119(1):45-54. DOI: 10.1083/jcb.119.1.45
  13. 13. Curtis R, Scherer SS, Somogyi R, et al. Retrograde axonal transport of LIF is increased by peripheral nerve injury: Correlation with increased LIF expression in distal nerve. Neuron. 1994;12(1):191-204. DOI: 10.1016/0896-6273(94)90163-5
  14. 14. Höke A, Redett R, Hameed H, et al. Schwann cells express motor and sensory phenotypes that regulate axon regeneration. The Journal of Neuroscience. 2006;26(38):9646-9655. DOI: 10.1523/JNEUROSCI.1620-06.2006
  15. 15. Webster HDF. The geometry of peripheral myelin sheaths during their formation and growth in rat sciatic nerves. The Journal of Cell Biology. 1971;48(2):348-367. DOI: 10.1083/jcb.48.2.348
  16. 16. Murinson BB, Archer DR, Li Y, Griffin JW. Degeneration of myelinated efferent fibers prompts mitosis in Remak Schwann cells of uninjured C-fiber afferents. The Journal of Neuroscience. 2005;25(5):1179-1187. DOI: 10.1523/JNEUROSCI.1372-04.2005
  17. 17. Topilko P, Schneider-Maunoury S, Levi G, et al. Krox-20 controls myelination in the peripheral nervous system. Nature. 1994;371(6500):796-799. DOI: 10.1038/371796a0
  18. 18. Nagarajan R, Svaren J, Le N, Araki T, Watson M, Milbrandt J. EGR2 mutations in inherited neuropathies dominant-negatively inhibit myelin gene expression. Neuron. 2001;30(2):355-368. DOI: 10.1016/S0896-6273(01)00282-3
  19. 19. Parkinson DB, Bhaskaran A, Droggiti A, et al. Krox-20 inhibits Jun-NH2-terminal kinase/c-Jun to control Schwann cell proliferation and death. The Journal of Cell Biology. 2004;164(3):385-394. DOI: 10.1083/jcb.200307132
  20. 20. Falls DL. Neuregulins: Functions, forms, and signaling strategies. Experimental Cell Research. 2003;284(1):14-30. DOI: 10.1016/S0014-4827(02)00102-7
  21. 21. Cohen JA, Yachnis AT, Arai M, Davis JG, Scherer SS. Expression of the neu proto-oncogene by schwann cells during peripheral nerve development and wallerian degeneration. Journal of Neuroscience Research. 1992;31(4):622-634. DOI: 10.1002/jnr.490310406
  22. 22. Ho WH, Armanini MP, Nuijens A, Phillips HS, Osheroff PL. Sensory and motor neuron-derived factor. A novel heregulin variant highly expressed in sensory and motor neurons. The Journal of Biological Chemistry. 1995;270(24):14523-14532. DOI: 10.1074/jbc.270.24.14523
  23. 23. Levi ADO, Bunge RP, Lofgren JA, et al. The influence of heregulins on human Schwann cell proliferation. The Journal of Neuroscience. 1995;15(2):1329-1340. DOI: 10.1523/jneurosci.15-02-01329.1995
  24. 24. Carroll SL, Miller ML, Frohnert PW, Kim SS, Corbett JA. Expression of neuregulins and their putative receptors, ErbB2 and ErbB3, is induced during Wallerian degeneration. The Journal of Neuroscience. 1997;17(5):1642-1659. DOI: 10.1523/jneurosci.17-05-01642.1997
  25. 25. Vartanian T, Goodearl A, Viehöver A, Fischbach G. Axonal neuregulin signals cells of the oligodendrocyte lineage through activation of HER4 and Schwann cells through HER2 and HER3. The Journal of Cell Biology. 1997;137(1):211-220. DOI: 10.1083/jcb.137.1.211
  26. 26. Newbern J, Birchmeier C. Nrg1/ErbB signaling networks in Schwann cell development and myelination. Seminars in Cell & Developmental Biology. 2010;21(9):922-928. DOI: 10.1016/j.semcdb.2010.08.008
  27. 27. Grigoryan T, Birchmeier W. Molecular signaling mechanisms of axon-glia communication in the peripheral nervous system. BioEssays. 2015;37(5):502-513. DOI: 10.1002/bies.201400172
  28. 28. Willem M. Proteolytic processing of Neuregulin-1. Brain Research Bulletin. 2016;126:178-182. DOI: 10.1016/j.brainresbull.2016.07.003
  29. 29. Syroid DE, Maycox PR, Burrola PG, et al. Cell death in the Schwann cell lineage and its regulation by neuregulin. Proceedings of the National Academy of Sciences of the United States of America. 1996;93(17):9229-9234. DOI: 10.1073/pnas.93.17.9229
  30. 30. Leimeroth R, Lobsiger C, Lüssi A, Taylor V, Suter U, Sommer L. Membrane-bound neuregulin1 type III actively promotes Schwann cell differentiation of multipotent progenitor cells. Developmental Biology. 2002;246(2):245-258. DOI: 10.1006/dbio.2002.0670
  31. 31. Taveggia C, Zanazzi G, Petrylak A, et al. Neuregulin-1 type III determines the ensheathment fate of axons. Neuron. 2005;47(5):681-694. DOI: 10.1016/j.neuron.2005.08.017
  32. 32. Chen S, Velardez MO, Warot X, et al. Neuregulin 1-erbB signaling is necessary for normal myelination and sensory function. The Journal of Neuroscience. 2006;26(12):3079-3086. DOI: 10.1523/JNEUROSCI.3785-05.2006
  33. 33. Kalderon N. Migration of Schwann cells and wrapping of neurites in vitro: A function of protease activity (plasmin) in the growth medium. Proceedings of the National Academy of Sciences of the United States of America. 1979;76(11):5992-5996. DOI: 10.1073/pnas.76.11.5992
  34. 34. Baron-Van Evercooren A, Leprince P, Rogister B, et al. Plasminogen activators in developing peripheral nervous system, cellular origin and mitogenic effect. Developmental Brain Research. 1987;36(1):101-108. DOI: 10.1016/0165-3806(87)90068-X
  35. 35. Hu X, Hicks CW, He W, et al. Bace1 modulates myelination in the central and peripheral nervous system. Nature Neuroscience. 2006;9(12):1520-1525. DOI: 10.1038/nn1797
  36. 36. Willem M, Garratt AN, Novak B, et al. Control of peripheral nerve myelination by the β-secretase BACE1. Science (80-). 2006;314(5799):664-666. DOI: 10.1126/science.1132341
  37. 37. La Marca R, Cerri F, Horiuchi K, et al. TACE (ADAM17) inhibits Schwann cell myelination. Nature Neuroscience. 2011;14(7):857-865. DOI: 10.1038/nn.2849
  38. 38. Bunge MB, Bunge RP, Kleitman N, Dean AC. Role of peripheral nerve extracellular matrix in schwann cell function and in neurite regeneration. Developmental Neuroscience. 1989;11(4-5):348-360. DOI: 10.1159/000111911
  39. 39. Bunge MB, Williams AK, Wood PM. Neuron-schwann cell interaction in basal lamina formation. Developmental Biology. 1982;92(2):449-460. DOI: 10.1016/0012-1606(82)90190-7
  40. 40. Carey DJ, Eldridge CF, Cornbrooks CJ, Timpl R, Bunge RP. Biosynthesis of type IV collagen by cultured rat Schwann cells. The Journal of Cell Biology. 1983;97(2):473-479. DOI: 10.1083/jcb.97.2.473
  41. 41. Carey DJ, Todd MS. Schwann cell myelination in a chemically defined medium: Demonstration of a requirement for additives that promote Schwann cell extracellular matrix formation. Developmental Brain Research. 1987;32(1):95-102. DOI: 10.1016/0165-3806(87)90142-8
  42. 42. Boerboom A, Dion V, Chariot A, Franzen R. Molecular mechanisms involved in schwann cell plasticity. Frontiers in Molecular Neuroscience. 2017;10(February):1-18. DOI: 10.3389/fnmol.2017.00038
  43. 43. Simpson SA, Young JZ. Regeneration of fibre diameter after cross-unions of visceral and somatic nerves. Journal of Anatomy. 1945;79(Pt 2):48-65
  44. 44. Weinberg HJ, Spencer PS. Studies on the control of myelinogenesis. I. Myelination of regenerating axons after entry into a foreign unmyelinated nerve. Journal of Neurocytology. 1975;4(4):395-418. DOI: 10.1007/BF01261372
  45. 45. Aguayo AJ, Epps J, Charron L, Bray GM. Multipotentiality of Schwann cells in cross-anastomosed and grafted myelinated and unmyelinated nerves: Quantitative microscopy and radioautography. Brain Research. 1976;104(1):1-20. DOI: 10.1016/0006-8993(76)90643-0
  46. 46. Chen Z-L, Yu W-M, Strickland S. Peripheral regeneration. Annual Review of Neuroscience. 2007;30(1):209-233. DOI: 10.1146/annurev.neuro.30.051606.094337
  47. 47. Jessen KR, Mirsky R. The repair Schwann cell and its function in regenerating nerves. The Journal of Physiology. 2016;594(13):3521-3531. DOI: 10.1113/JP270874
  48. 48. Gomez-Sanchez JA, Carty L, Iruarrizaga-Lejarreta M, et al. Schwann cell autophagy, myelinophagy, initiates myelin clearance from injured nerves. The Journal of Cell Biology. 2015;210(1):153-168. DOI: 10.1083/jcb.201503019
  49. 49. Hirata K, Kawabuchi M. Myelin phagocytosis by macrophages and nonmacrophages during Wallerian degeneration. Microscopy Research and Technique. 2002;57(6):541-547. DOI: 10.1002/jemt.10108
  50. 50. Lee H, Jo EK, Choi SY, et al. Necrotic neuronal cells induce inflammatory Schwann cell activation via TLR2 and TLR3: Implication in Wallerian degeneration. Biochemical and Biophysical Research Communications. 2006;350(3):742-747. DOI: 10.1016/j.bbrc.2006.09.108
  51. 51. Barrette B, Hébert MA, Filali M, et al. Requirement of myeloid cells for axon regeneration. The Journal of Neuroscience. 2008;28(38):9363-9376. DOI: 10.1523/JNEUROSCI.1447-08.2008
  52. 52. Jessen KR, Mirsky R. The success and failure of the schwann cell response to nerve injury. Frontiers in Cellular Neuroscience. 2019;13:1-14. DOI: 10.3389/fncel.2019.00033
  53. 53. Lemke G, Chao M. Axons regulate Schwann cell expression of the major myelin and NGF receptor genes. Development. 1988;102(3):499-504
  54. 54. Terenghi G. Peripheral nerve regeneration and neurotrophic factors. Journal of Anatomy. 1999;194(1):1-14. DOI: 10.1017/S0021878298004312
  55. 55. Saxena S, Caroni P. Mechanisms of axon degeneration: From development to disease. Progress in Neurobiology. 2007;83(3):174-191. DOI: 10.1016/j.pneurobio.2007.07.007
  56. 56. Ghabriel MN, Allt G. The role of Schmidt-Lanterman incisures in Wallerian degeneration—II. An electron microscopic study. Acta Neuropathologica. 1979;48(2):95-103. DOI: 10.1007/BF00691150
  57. 57. Ghabriel MN, Allt G. Incisures of Schmidt-Lanterman. Progress in Neurobiology. 1981;17(1-2):25-58. DOI: 10.1016/0301-0082(81)90003-4
  58. 58. Stoll G, Griffin JW, Li CY, Trapp BD. Wallerian degeneration in the peripheral nervous system: Participation of both Schwann cells and macrophages in myelin degradation. Journal of Neurocytology. 1989;18(5):671-683. DOI: 10.1007/BF01187086
  59. 59. Beuche W, Friede RL. The role of non-resident cells in Wallerian degeneration. Journal of Neurocytology. 1984;13(5):767-796. DOI: 10.1007/BF01148493
  60. 60. Stoll G, Müller HW. Nerve injury, axonal degeneration and neural regeneration: Basic insights. Brain Pathology. 2006;9(2):313-325. DOI: 10.1111/j.1750-3639.1999.tb00229.x
  61. 61. Kidd G, Andrews SB, Trapp BD. Axons regulate the distribution of Schwann cell microtubules. The Journal of Neuroscience. 1996;16(3):946-954. DOI: 10.1523/jneurosci.16-03-00946.1996
  62. 62. Parrinello S, Napoli I, Ribeiro S, et al. EphB signaling directs peripheral nerve regeneration through Sox2-dependent Schwann cell sorting. Cell. 2010;143(1):145-155. DOI: 10.1016/j.cell.2010.08.039
  63. 63. Cattin AL, Burden JJ, Van Emmenis L, et al. Macrophage-induced blood vessels guide Schwann cell-mediated regeneration of peripheral nerves. Cell. 2015;162(5):1127-1139. DOI: 10.1016/j.cell.2015.07.021
  64. 64. Schröder JM. Altered ratio between axon diameter and myelin sheath thickness in regenerated nerve fibers. Brain Research. 1972;45(1):49-65. DOI: 10.1016/0006-8993(72)90215-6
  65. 65. Stewart HJS. Expression of c-Jun, Jun B, Jun D and cAMP response element binding protein by schwann cells and their precursors in vivo and in vitro. The European Journal of Neuroscience. 1995;7(6):1366-1375. DOI: 10.1111/j.1460-9568.1995.tb01128.x
  66. 66. Parkinson DB, Bhaskaran A, Arthur-Farraj P, et al. c-Jun is a negative regulator of myelination. The Journal of Cell Biology. 2008;181(4):625-637. DOI: 10.1083/jcb.200803013
  67. 67. Hutton EJ, Carty L, Laurá M, et al. C-Jun expression in human neuropathies: A pilot study. Journal of the Peripheral Nervous System. 2011;16(4):295-303. DOI: 10.1111/j.1529-8027.2011.00360.x
  68. 68. Hantke J, Carty L, Wagstaff LJ, et al. c-Jun activation in Schwann cells protects against loss of sensory axons in inherited neuropathy. Brain. 2014;137(11):2922-2937. DOI: 10.1093/brain/awu257
  69. 69. Klein D, Groh J, Wettmarshausen J, Martini R. Nonuniform molecular features of myelinating Schwann cells in models for CMT1: Distinct disease patterns are associated with NCAM and c-Jun upregulation. Glia. 2014;62(5):736-750. DOI: 10.1002/glia.22638
  70. 70. Arthur-Farraj PJ, Latouche M, Wilton DK, et al. c-Jun reprograms schwann cells of injured nerves to generate a repair cell essential for regeneration. Neuron. 2012;75(4):633-647. DOI: 10.1016/j.neuron.2012.06.021
  71. 71. Woodhoo A, Alonso MBD, Droggiti A, et al. Notch controls embryonic Schwann cell differentiation, postnatal myelination and adult plasticity. Nature Neuroscience. 2009;12(7):839-847. DOI: 10.1038/nn.2323
  72. 72. Nickols JC, Valentine W, Kanwal S, Carter BD. Activation of the transcription factor NF-κB in Schwann cells is required for peripheral myelin formation. Nature Neuroscience. 2003;6(2):161-167. DOI: 10.1038/nn995
  73. 73. Yoon C, Korade Z, Carter BD. Protein kinase A-induced phosphorylation of the p65 subunit of nuclear factor-κB promotes Schwann cell differentiation into a myelinating phenotype. The Journal of Neuroscience. 2008;28(14):3738-3746. DOI: 10.1523/JNEUROSCI.4439-07.2008
  74. 74. Limpert AS, Carter BD. Axonal neuregulin 1 type III activates NF-κB in Schwann cells during myelin formation. The Journal of Biological Chemistry. 2010;285(22):16614-16622. DOI: 10.1074/jbc.M109.098780
  75. 75. Quintes S, Brinkmann BG, Ebert M, et al. Zeb2 is essential for Schwann cell differentiation, myelination and nerve repair. Nature Neuroscience. 2016;19(8):1050-1059. DOI: 10.1038/nn.4321
  76. 76. Doddrell RDS, Dun XP, Moate RM, Jessen KR, Mirsky R, Parkinson DB. Regulation of Schwann cell differentiation and proliferation by the Pax-3 transcription factor. Glia. 2012;60(9):1269-1278. DOI: 10.1002/glia.22346
  77. 77. Gao X, Daugherty RL, Tourtellotte WG. Regulation of low affinity neurotrophin receptor (p75NTR) by early growth response (Egr) transcriptional regulators. Molecular and Cellular Neurosciences. 2007;36(4):501-514. DOI: 10.1016/j.mcn.2007.08.013
  78. 78. Norrmén C, Figlia G, Pfistner P, Pereira JA, Bachofner S, Suter U. mTORC1 is transiently reactivated in injured nerves to promote c-Jun elevation and schwann cell dedifferentiation. The Journal of Neuroscience. 2018;38(20):4811-4828. DOI: 10.1523/JNEUROSCI.3619-17.2018
  79. 79. Laughner E, Taghavi P, Chiles K, Mahon PC, Semenza GL. HER2 (neu) signaling increases the rate of hypoxia-inducible factor 1 (HIF-1) synthesis: Novel mechanism for HIF-1-mediated vascular endothelial growth factor expression. Molecular and Cellular Biology. 2001;21(12):3995-4004. DOI: 10.1128/mcb.21.12.3995-4004.2001
  80. 80. Roczniak-Ferguson A, Petit CS, Froehlich F, et al. The transcription factor TFEB links mTORC1 signaling to transcriptional control of lysosome homeostasis. Science Signaling. 2012;5(228):ra42. DOI: 10.1126/scisignal.2002790
  81. 81. Tiebe M, Lutz M, De La Garza A, Buechling T, Boutros M, Teleman AA. REPTOR and REPTOR-BP regulate organismal metabolism and transcription downstream of TORC1. Developmental Cell. 2015;33(3):272-284. DOI: 10.1016/j.devcel.2015.03.013
  82. 82. Park Y, Reyna-Neyra A, Philippe L, Thoreen CC. mTORC1 balances cellular amino acid supply with demand for protein synthesis through post-transcriptional control of ATF4. Cell Reports. 2017;19(6):1083-1090. DOI: 10.1016/j.celrep.2017.04.042
  83. 83. Ben-Sahra I, Manning BD. mTORC1 signaling and the metabolic control of cell growth. Current Opinion in Cell Biology. 2017;45:72-82. DOI: 10.1016/j.ceb.2017.02.012
  84. 84. Saxton RA, Sabatini DM. mTOR signaling in growth, metabolism, and disease. Cell. 2017;168(6):960-976. DOI: 10.1016/j.cell.2017.02.004
  85. 85. Sherman DL, Krols M, Wu LMN, et al. Arrest of myelination and reduced axon growth when Schwann cells lack mTOR. The Journal of Neuroscience. 2012;32(5):1817-1825. DOI: 10.1523/JNEUROSCI.4814-11.2012
  86. 86. Norrmén C, Figlia G, Lebrun-Julien F, et al. mTORC1 controls PNS myelination along the mTORC1-RXRγ-SREBP-lipid biosynthesis axis in Schwann cells. Cell Reports. 2014;9(2):646-660. DOI: 10.1016/j.celrep.2014.09.001
  87. 87. Beirowski B, Wong KM, Babetto E, Milbrandt J. MTORC1 promotes proliferation of immature Schwann cells and myelin growth of differentiated Schwann cells. Proceedings of the National Academy of Sciences of the United States of America. 2017;114(21):E4261-E4270. DOI: 10.1073/pnas.1620761114
  88. 88. Figlia G, Norrmén C, Pereira JA, Gerber D, Suter U. Dual function of the PI3K-Akt-MTORC1 axis in myelination of the peripheral nervous system. eLife. 2017;6:e29241. DOI: 10.7554/eLife.29241
  89. 89. Figlia G, Gerber D, Suter U. Myelination and mTOR. Glia. 2018;66(4):693-707. DOI: 10.1002/glia.23273
  90. 90. Sheu JY, Kulhanek DJ, Eckenstein FP. Differential patterns of ERK and STAT3 phosphorylation after sciatic nerve transection in the rat. Experimental Neurology. 2000;166(2):392-402. DOI: 10.1006/exnr.2000.7508
  91. 91. Harrisingh MC, Perez-Nadales E, Parkinson DB, Malcolm DS, Mudge AW, Lloyd AC. The Ras/Raf/ERK signalling pathway drives Schwann cell dedifferentiation. The EMBO Journal. 2004;23(15):3061-3071. DOI: 10.1038/sj.emboj.7600309
  92. 92. Zrouri H, Le Goascogne C, Li WW, Pierre M, Courtin F. The role of MAP kinases in rapid gene induction after lesioning of the rat sciatic nerve. The European Journal of Neuroscience. 2004;20(7):1811-1818. DOI: 10.1111/j.1460-9568.2004.03641.x
  93. 93. Agthong S, Kaewsema A, Tanomsridejchai N, Chentanez V. Activation of MAPK ERK in peripheral nerve after injury. BMC Neuroscience. 2006;7:45. DOI: 10.1186/1471-2202-7-45
  94. 94. Lee HJ, Shin YK, Park HT. Mitogen activated protein kinase family proteins and c-jun signaling in injury-induced schwann cell plasticity. Experimental Neurobiology. 2014;23(2):130. DOI: 10.5607/en.2014.23.2.130
  95. 95. Ronchi G, Haastert-Talini K, Fornasari BE, Perroteau I, Geuna S, Gambarotta G. The Neuregulin1/ErbB system is selectively regulated during peripheral nerve degeneration and regeneration. The European Journal of Neuroscience. 2016;43(3):351-364. DOI: 10.1111/ejn.12974
  96. 96. Grossmann KS, Wende H, Paul FE, et al. The tyrosine phosphatase Shp2 (PTPN11) directs neuregulin-1/ErbB signaling throughout Schwann cell development. Proceedings of the National Academy of Sciences of the United States of America. 2009;106(39):16704-16709. DOI: 10.1073/pnas.0904336106
  97. 97. He Y, Kim JY, Dupree J, et al. Yy1 as a molecular link between neuregulin and transcriptional modulation of peripheral myelination. Nature Neuroscience. 2010;13(12):1472-1482. DOI: 10.1038/nn.2686
  98. 98. Newbern JM, Li X, Shoemaker SE, et al. Specific functions for ERK/MAPK signaling during PNS development. Neuron. 2011;69(1):91-105. DOI: 10.1016/j.neuron.2010.12.003
  99. 99. Napoli I, Noon LA, Ribeiro S, et al. A central role for the ERK-signaling pathway in controlling schwann cell plasticity and peripheral nerve regeneration in vivo. Neuron. 2012;73(4):729-742. DOI: 10.1016/j.neuron.2011.11.031
  100. 100. Newbern JM, Snider WD. Bers-ERK Schwann cells coordinate nerve regeneration. Neuron. 2012;73(4):623-626. DOI: 10.1016/j.neuron.2012.02.002
  101. 101. Parkinson DB, Dong Z, Bunting H, et al. Transforming growth factor β (TGFβ) mediates Schwann cell death in vitro and in vivo: Examination of c-Jun activation, interactions with survival signals, and the relationship of TGFβ-mediated death to Schwann cell differentiation. The Journal of Neuroscience. 2001;21(21):8572-8585. DOI: 10.1523/jneurosci.21-21-08572.2001
  102. 102. Yamauchi J, Chan JR, Shooter EM. Neurotrophin 3 activation of TrkC induces Schwann cell migration through the c-Jun N-terminal kinase pathway. Proceedings of the National Academy of Sciences of the United States of America. 2003;100(SUPPL. 2):14421-14426. DOI: 10.1073/pnas.2336152100
  103. 103. Thakur KK, Saini J, Mahajan K, et al. Therapeutic implications of toll-like receptors in peripheral neuropathic pain. Pharmacological Research. 2017;115:224-232. DOI: 10.1016/j.phrs.2016.11.019
  104. 104. Goethals S, Ydens E, Timmerman V, Janssens S. Toll-like receptor expression in the peripheral nerve. Glia. 2010;58(14):1701-1709. DOI: 10.1002/glia.21041
  105. 105. Boivin A, Pineau I, Barrette B, et al. Toll-like receptor signaling is critical for Wallerian degeneration and functional recovery after peripheral nerve injury. The Journal of Neuroscience. 2007;27(46):12565-12576. DOI: 10.1523/JNEUROSCI.3027-07.2007
  106. 106. Atanasoski S, Scherer SS, Sirkowski E, et al. ErbB2 signaling in Schwann cells is mostly dispensable for maintenance of myelinated peripheral nerves and proliferation of adult Schwann cells after injury. The Journal of Neuroscience. 2006;26(7):2124-2131. DOI: 10.1523/JNEUROSCI.4594-05.2006
  107. 107. Guertin AD, Zhang DP, Mak KS, Alberta JA, Kim HA. Microanatomy of axon/glial signaling during Wallerian degeneration. The Journal of Neuroscience. 2005;25(13):3478-3487. DOI: 10.1523/JNEUROSCI.3766-04.2005
  108. 108. Ronchi G, Cillino M, Gambarotta G, et al. Irreversible changes occurring in long-term denervated Schwann cells affect delayed nerve repair. Journal of Neurosurgery. 2017;127(4):843-856. DOI: 10.3171/2016.9.JNS16140
  109. 109. Mahanthappa NK, Anton ES, Matthew WD. Glial growth factor 2, a soluble neuregulin, directly increases Schwann cell motility and indirectly promotes neurite outgrowth. The Journal of Neuroscience. 1996;16(15):4673-4683. DOI: 10.1523/jneurosci.16-15-04673.1996
  110. 110. Fricker FR, Bennett DL. The role of neuregulin-1 in the response to nerve injury. Future Neurology. 2011;6(6):809-822. DOI: 10.2217/fnl.11.45
  111. 111. Joshi AR, Holtmann L, Bobylev I, et al. Loss of Schwann cell plasticity in chronic inflammatory demyelinating polyneuropathy (CIDP). Journal of Neuroinflammation. 2016;13(1):255. DOI: 10.1186/s12974-016-0711-7
  112. 112. Zochodne DW. The challenges and beauty of peripheral nerve regrowth. Journal of the Peripheral Nervous System. 2012;17:1-18. DOI: 10.1111/j.1529-8027.2012.00378.x
  113. 113. Painter MW, Brosius Lutz A, Cheng YC, et al. Diminished Schwann cell repair responses underlie age-associated impaired axonal regeneration. Neuron. 2014;83(2):331-343. DOI: 10.1016/j.neuron.2014.06.016
  114. 114. Frohnert PW, Stonecypher MS, Carroll SL. Constitutive activation of the neuregulin-1/ErbB receptor signaling pathway is essential for the proliferation of a neoplastic Schwann cell line. Glia. 2003;43(2):104-118. DOI: 10.1002/glia.10232
  115. 115. Fallon KB, Havlioglu N, Hamilton LH, Cheng TPH, Carroll SL. Constitutive activation of the neuregulin-1/erbB signaling pathway promotes the proliferation of a human peripheral neuroepithelioma cell line. Journal of Neuro-Oncology. 2004;66(3):273-284. DOI: 10.1023/B:NEON.0000014521.28294.84
  116. 116. Huijbregts RPH, Roth KA, Schmidt RE, Carroll SL. Hypertrophic neuropathies and malignant peripheral nerve sheath tumors in transgenic mice overexpressing glial growth factor β3 in myelinating Schwann cells. The Journal of Neuroscience. 2003;23(19):7269-7280. DOI: 10.1523/jneurosci.23-19-07269.2003
  117. 117. Tapinos N, Ohnishi M, Rambukkana A. ErbB2 receptor tyrosine kinase signaling mediates early demyelination induced by leprosy bacilli. Nature Medicine. 2006;12(8):961-966. DOI: 10.1038/nm1433
  118. 118. Fledrich R, Stassart RM, Klink A, et al. Soluble neuregulin-1 modulates disease pathogenesis in rodent models of Charcot-Marie-Tooth disease 1A. Nature Medicine. 2014;20(9):1055-1061. DOI: 10.1038/nm.3664
  119. 119. Manole E, Bastian A, Ristoiu V, Zurac S, Neagu M. The effects of exogenous modulation on the peripheral nerve regeneration after injury and primary surgical repair. Biomedical Journal of Scientific & Technical Research. 2018;4(3):1-5. DOI: 10.26717/bjstr.2018.04.0001043

Written By

Emilia Manole, Alexandra Eugenia Bastian, Ana Maria Oproiu, Monica Teodora Neagu, Carolina Constantin and Gheorghita Isvoranu

Submitted: 30 January 2020 Reviewed: 18 February 2020 Published: 03 April 2020