Next Article in Journal
Schwannoma Gene Therapy via Adeno-Associated Viral Vector Delivery of Apoptosis-Associated Speck-like Protein Containing CARD (ASC): Preclinical Efficacy and Safety
Next Article in Special Issue
Adipose-Derived Stem Cells from Type 2 Diabetic Rats Retain Positive Effects in a Rat Model of Erectile Dysfunction
Previous Article in Journal
Activity- and Enrichment-Based Metaproteomics Insights into Active Urease from the Rumen Microbiota of Cattle
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Brief Review of In Vitro Models for Injury and Regeneration in the Peripheral Nervous System

Centre for Nanosciences & Molecular Medicine, Amrita Institute of Medical Sciences and Research Center, Amrita Vishwa Vidyapeetham, Kochi 682041, India
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(2), 816; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23020816
Submission received: 6 November 2021 / Revised: 3 December 2021 / Accepted: 5 December 2021 / Published: 13 January 2022

Abstract

:
Nerve axonal injury and associated cellular mechanisms leading to peripheral nerve damage are important topics of research necessary for reducing disability and enhancing quality of life. Model systems that mimic the biological changes that occur during human nerve injury are crucial for the identification of cellular responses, screening of novel therapeutic molecules, and design of neural regeneration strategies. In addition to in vivo and mathematical models, in vitro axonal injury models provide a simple, robust, and reductionist platform to partially understand nerve injury pathogenesis and regeneration. In recent years, there have been several advances related to in vitro techniques that focus on the utilization of custom-fabricated cell culture chambers, microfluidic chamber systems, and injury techniques such as laser ablation and axonal stretching. These developments seem to reflect a gradual and natural progression towards understanding molecular and signaling events at an individual axon and neuronal-soma level. In this review, we attempt to categorize and discuss various in vitro models of injury relevant to the peripheral nervous system and highlight their strengths, weaknesses, and opportunities. Such models will help to recreate the post-injury microenvironment and aid in the development of therapeutic strategies that can accelerate nerve repair.

1. Introduction

Mammalian axons in the peripheral nervous system (PNS), unlike their counterparts in the central nervous system (CNS), possess the ability to repair and regenerate to a large extent [1]. In fact, the reason for this observed dichotomy has been the focus of several studies for over a century. However, it is now generally accepted that the microenvironment within the PNS allows cellular processes that favor axonal growth [2,3,4], while that in the CNS is inhibitory [5,6]. Despite our understanding of the PNS, changes that occur during nerve injuries are not completely clear and require many in-depth studies.
Peripheral nerve injuries (PNI) are a frequent and major cause of human morbidity worldwide. Injury to the peripheral nerves results in pain and loss of function and can severely affect the patient’s mental health [7,8]. Additionally, affected patients also have to bear the financial burden of treatment and rehabilitation, thus resulting in a loss of workdays and overall quality of life [8,9]. Currently, the existing line of treatment for partial to complete PNI focuses on the management of specific symptoms, surgical repair [10,11], or complete nerve transplant [7,12]. However, newer strategies that are more practical and cost-effective in managing nerve injuries are much needed. The development of such innovative treatment options requires a thorough understanding of the cellular and molecular mechanisms involved in peripheral nerve repair and healing. In general, isolated cellular systems and animal models have proven to be extremely useful in understanding PNI-associated molecular mechanisms, electrophysiological alterations [13,14] and changes in gene expression. Importantly, in addition to biological information, inputs from experimental systems have immensely contributed to the development of mathematical and computational models to predict neural changes in PNI [15,16].
Amongst the various experimental strategies to address PNI, well-designed in vitro models, in particular, have been crucial in deciphering, discovering, and identifying many of the molecular determinants and events of nerve repair [17]. In this review, we attempt to compile, list, and expand on the existing in vitro models of nerve injury and describe the strengths and weaknesses of each model system. For in vivo and mathematical models of axonal injury, readers are directed towards other excellent reviews [7,18,19] for a better understanding of their use and application.

1.1. In Vitro Nerve Injury: A Window to Observe Cellular Mechanisms

In vitro nerve injury models provide vast opportunities to study cellular responses due to injury, including mechanisms for regeneration, plasticity, and degeneration. Models that mimic axotomy, drug-, or metabolite-induced nerve damage are quite useful in understanding molecular mechanisms and screening potentially beneficial therapeutic strategies. In addition to mechanistic details, it is now well-recognized that the primary initiator of injury or the type of injury plays a significant role in dictating the direction of downstream injury-associated signaling events [20,21]. Hence, it becomes quite vital that there is a thorough understanding of each model system in terms of what they address and, more importantly, what they do not.

1.2. PNI-Associated Cellular Changes Observed In Vitro

Generally, injury-induced interruption of axonal continuity from the neuronal cell soma is known to induce certain well-characterized cellular changes, both in the distal and proximal axonal segments. These changes have been observed in both in vivo and in vitro models. One such commonly observed cellular event is the progressive degeneration of the injured distal axonal segment, called Wallerian degeneration (WD) [22]. Sciatic nerve explants maintained for more than 2 weeks in vitro have been found to mimic several in vivo observations, such as upregulation of repair program protein c-Jun [22,23], downregulation of several myelin protein genes [22,24], and disorganization of the myelin structure around the axon [22,25,26]. In addition, WD in vitro models have been used to understand the role of important injury-induced proteins belonging to the family of calcium-dependent non-lysosomal cysteine proteases [27,28,29] such as calpains. Specifically, in vitro models such as the neurite transection models and the dorsal root ganglion (DRG) explant models have been used to demonstrate the calpain-mediated proteolysis of the neuronal cytoskeleton [30,31].
Another important hallmark of lesioned axons is the formation of retraction bulbs within the length of the axons [32,33]. Retraction bulbs lack microtubule organization and are generally considered the non-growing counterparts of neuronal growth cones [34]. In vitro models have been consistently utilized to study retraction bulb formation and the subsequent inhibition of axonal growth in the presence of inhibitors such as nocodazole [33,35,36]. Other morphological features that are consistently observed in axonal injury, including the dissipation of chromatin [37], chromatolysis [38], endoplasmic blebbing [39], and loss of dendritic spines [40], have been widely demonstrated in various in vitro models [17]. In addition to mimicking injury-induced morphological changes, in vitro models have also been successfully utilized to study the injury-induced influx of Ca2+ ions in axons [41,42] and associated downstream events [43,44]. Overall, it is quite apparent that in vitro models can significantly contribute to our understanding of various neuronal events, both during and after injury. Additionally, appropriate adoption of the PNI in vitro models would also contribute to the reduction, refinements, and replacement (the ‘3R’ principle) of animals in nerve injury research. Below, we list and describe the features, advantages, and disadvantages of in vitro PNI models that are currently in use (Figure 1).

2. Transection Injury Models

Axonal transection, or the ‘acute axotomy model’, is one of the simplest but most effective in vitro models used to investigate the process of axonal regeneration and response to injury. Robert Campenot presented the first in vitro compartment model in 1977 [45] using a three-chamber culture system in which neurites sprouting from the soma of sympathetic neurons were observed to cross a fluid-impermeable barrier. Typically, in this model, dissociated primary neurons are cultured on coated cell culture plates within compartmentalized areas. Uniformly placed scratches on the coated culture plate ensure that the regenerating neurites grow linearly on the coated areas in-between scratches, with minimal bending (Figure 2). Notably, the rake for applying scratches is made by cementing 20 insect pins placed 200 μm apart, which yields a set of 20 parallel scratches on coated cell culture plates [45,46,47].

2.1. The Campenot Chamber Models

The Campenot chamber allows the spatial segregation of neuronal soma and its neurites and facilitates studies related to neuronal and axonal biology [36,45,46,48]. The classical Campenot chamber is a three-compartment culture system that allows the extension of neurites from the central compartment to the two flanking peripheral compartments. A fluidic seal is obtained by the application of a hydrophobic gel such as sterilized silicone grease [46,48,49] between the walls of the compartment and the underlying culture plate surface. The silicone grease allows neurites to penetrate across while providing options to maintain varying media conditions between compartments [45,48,49]. Within the chamber, transectional axonal injury infliction has been attempted by repeatedly spraying cold sterile water through a 0.22-gauge needle into chambers that contain axons [50,51,52]. This process results in the transection of axons that can be easily visualized using conventional immunostaining methods under epifluorescence or confocal microscopes. Studies performed in this model have demonstrated that transection injury affects the rate of axonal elongation and survival [52], as well as inducing a considerable reduction in the ‘slow axonal transport’ mechanism, which is crucial for supplying cytoskeletal proteins [50] to the neuron.

2.2. Microfluidic Chamber Models

Injury-induced axonal changes similar to those described above have also been observed in other compartmental models, such as the 2D- and 3D microfluidic chip systems [53,54]. These model systems fabricated from an inert polymer material such as poly(dimethylsiloxane) [55,56] comprise compartments and microgrooves of varying design, which spatially separate neuronal axons and soma [57]. Axons within these microfluidic devices extend into dedicated microgrooves, which allows the selective manipulation of axons and provides opportunities for the measurement of neurite growth response [58], somal transcription activity [56], and axonal degeneration [59]. In addition, these model systems have also been utilized to understand neuronal injury associated with amyotrophic lateral sclerosis [60], the overall dynamics of the neural network, and applications involving axonal biology, drug discovery, and regenerative medicine [61].

2.3. Forceful Trituration

In conjunction with the microfluidic model, various types of axonal injury strategies have been applied. For instance, aspiration of a small volume of fluid using a fine-tip glass pipette placed at the entrance of the axonal compartment resulted in injury-associated upregulation of the Nav1.6 sodium channel [62]. Similarly, microfluidic vacuum aspiration followed by introducing air bubbles within the axonal compartment resulted in features linked to acute axonal degeneration, delayed bulb formation, and impairment of axonal mitochondrial transport [63]. A recent study used these techniques to mimic the role of myotubular contractility on axonal regeneration, where axons from motor neurons co-cultured with myotubes were injured using fluidic aspiration along with bubble introduction [64].

2.4. Tear on Cell Monolayer

While spatial separation of the cell body and injured axons closely mimics nerve injury in the PNS, a few studies have taken a more direct approach in inflicting neuronal injury. Mixed cultures of glia and neurons with rich axonal growth were injured by running a pipette tip diametrically along the length of the culture plate, resulting in a ‘tear’ measuring approximately 0.6 to 1 mm wide across the cultured cell layer [65]. A second tear was applied perpendicular to the primary tear, thereby yielding four quadrants within the cell culture plate. This model yields areas that contain undisturbed cells deep within the quadrants that are distant to the injury and also cells that are in close proximity to the injury site. Within 2–3 min of tear application, the neuronal soma adjacent to the tear demonstrated swelling and granularity [65,66]. Swelling was also observed in neurons located 1–2 mm away from the in vitro lesion. Shortly thereafter, neurons located much deeper within the quadrants developed features of swelling, while glial cells remained unchanged [65].

2.5. Laser-Induced Transection

To address the complexities of neurite injuries at the level of single cells, studies have utilized strategies that make use of laser-induced microsurgical lesions. Importantly, axonal transections induced by a focused laser are quite precise, with minimal off-target injury and cytoplasmic damage to cells present in the vicinity. Recently, cultures of neuronal and glial cells exposed to multiple 6-nanosecond laser pulses at 355 nm exhibited immediate partial retraction of axons from the site of laser-induced injury, followed by axonal beading and proximal segment retraction after 1 s [67]. Interestingly, cell bodies of neurons that were subjected to laser-induced axonal injury demonstrated numerous vacuoles [68], especially in the mitochondria [69]. Similarly, neurite injury induced by a UV laser beam resulted in proximal and distal segment retraction with signs of degeneration comparatively higher in thicker segments than thinner segments [68,70]. In another laser-based transection model, low-energy picosecond laser pulses were used to induce single axon transection in vitro, with micro-level resolution [71] (Figure 3). To maximize precision and accuracy, axonal injuries were performed on an inverted microscope coupled with a high-resolution motorized stage and an environmental chamber. Similar to single axon injuries observed in other laser-injury models, the low-energy laser pulses induced axonal swelling, distal segment degeneration, axonal retraction, and vesicle formation [67,71,72]. These observations lend strong credence to using lasers for inducing axonal transection in vitro to mimic membranal and cytoplasmic changes that occur in injured neurons in vivo.
However, it should be noted that an inherent drawback of the in vitro technique, no matter the model system used, lies with the isolation of neural tissue, such as the dorsal root ganglia [57] and the sympathetic ganglia that are commonly used in PNI. Isolation of ganglia for neuronal dissociation involves transection of the nerve roots, which itself causes several injury-induced cellular responses within cultured neurons [73]. Hence, it is possible that many experimental in vitro injury-induced responses in neurons may be masked by the initial injury caused during ganglia harvesting and isolation. Investigators need to be aware of this confounding variable before venturing into studies involving the in vitro model systems and design their experiments accordingly.
In summary, the main advantages and disadvantages of the transection model are listed below:

2.6. Advantages

  • Compartmentalization of peripheral neuron cultures provides robust separation of axon and soma that morphologically reflects a more in vivo arrangement.
  • The spatial separation also allows for the application of consistent injury events, while the cellular response can be easily visualized and quantified using conventional approaches.
  • In addition to being cost-effective, this model also allows for the performance of various biochemical assays that require higher volumes of injured cells.
  • The use of laser-induced transection provides precision and accuracy for single-cell studies.

2.7. Disadvantages

  • Though the compartmental models are quite easy to set up, these devices are prone to fluid leak that may warrant repetition of experiments.
  • There may be difficulty in the assembly of the device for advanced live-cell microscopy, and necessary modifications may have to be incorporated.
  • Single-cell studies using laser-induced injuries will need advanced expertise for set-up and may not be possible in a typical laboratory setting.
  • It is quite challenging to limit glial cells within compartments, and eventually, they tend to migrate into all compartments.

3. Axonal Stretch Injury Model

Traction on the peripheral nerve is a common occurrence due to trauma, surgical procedures, and limb lengthening maneuvers. Interestingly, the stretching force exerted on a nerve has been found to be proportional to the ensuing nerve ischemia, with at least 15% nerve elongation required for precipitation of ischemia [74] and up to 50% elongation for the appearance of histological changes [75,76]. Nerve tissue scarring and tearing of the nerve outer layers (epineurium, perineurium) along with the attached blood vessels are some of the commonly observed lesions in nerve-stretch injury [77].
Unlike in an axotomy, where injury-induced damage is all-or-none, stretch-induced injury demonstrates a range of nerve responses, including an axonal growth phase with mild stretch [78,79] and an injury phase above the resting tension threshold [79,80]. One of the first in vitro models designed to understand stretch-induced injury at the cellular level was in CNS neurons cultured on flexible silastic membranes [81,82], while similar methods have been extended to peripheral neurons as well. In these models, neurons cultured on a flexible silastic membrane are exposed to controlled stretching forces, mainly by the application of calibrated pressure pulses that results in global deformation of the silastic membrane, which in turn stretches cells that are adhered to it [83,84,85]. The actual stretching of the membrane may be as brief as 250 ms, but this is sufficient to produce cell deformation [81,82,83,86] (Figure 4). Further, low levels of stretch were found to induce changes in membrane potential along with delayed depolarization, without any significant alteration in cell viability [86]. Interestingly, membrane-stretched cells have been found to depolarize only after stretch release [83].
Although few studies have reported only mild effects on low levels of axonal stretch, there have been other studies that have observed stretch-induced structural abnormalities in mitochondria, and vacuolization of golgi and rough-surfaced endoplasmic reticulum [84,87,88]. In addition, neurons cultured on flexible polymeric substrates subjected to substrate stretching show a strain-rate-dependent decrease in fast axonal transport [89] and immediate hyperpolarization [83].
Another significant advancement in studying axonal injury is the adoption of microfluidic devices. Neurons cultured within microfluidic devices have been subjected to axonal injury by simply pushing fluid into intersecting microgrooves within the device [54,59,62,90]. Axonal stretching using a cantilever attached to a piezoelectric drive is another innovative approach to produce stretch-induced injury changes at the single axon level (Figure 5). Such in vitro models have contributed quite well to our understanding of the biochemical and gene expression changes in the cell soma at an individual neuron level.
It is also interesting to note that there are conflicting reports in terms of injury severity in mild stretch models [91], with some studies demonstrating beneficial effects of mild nerve stretch [92,93]. In contrast, severe axonal stretching induces critical and permanent deficits, including mitochondrial swelling and cell death [94,95,96]. However, what needs to be noted is that unlike single axons, bundles of axons seem to be quite resilient towards external strain. This could be because of axon–axon adhesions that increase the tensile strength of the whole bundle or due to bundling itself [97]. This makes the conclusions from single axon studies difficult to interpret, especially when trying to mimic responses related to strain rates in vivo.
In summary, the main advantages and disadvantages of the cell stretch model are listed below:

3.1. Advantages

  • Direct and real-time measurement of controlled mechanical strain.
  • The cellular response is reproducible, and the principles behind stretching mechanisms are quite simple.

3.2. Disadvantages

  • Difficulty in mimicking various injury dynamics as seen in real-life conditions, where nerves may be exposed to stretching and twisting force.
  • Techniques such as incorporating a PDMS membrane within microfluidic systems requires experience and advanced skills.
  • Sum total of tensile strength of individual axons, being lower than the tensile strength of an entire nerve, makes in vitro models less relatable to in vivo observations.

4. Compression Induced-Injury Models

Compressive neuropathies are conditions that are associated with the physical compression of a nerve, resulting in pain, numbness, and possible loss of function [98,99]. Nerve compression can be either acute, as in traumatic crush injuries, or chronic compression, as observed in vertebral lesions and entrapment syndromes [100,101]. Few in vitro models have been developed to address the electrophysiological and morphological features of compression injury in nerves. Controlled compression applied on excised canine lumbar roots, with a rod equipped with a pressure gauge, and operated by a pulse motor, demonstrated considerable lowering of the threshold for ectopic firing and prolonged firing in DRG neurons [102]. Significantly, these in vitro findings were later validated in chronically compressed DRG neurons in vivo [103,104]. In vitro compression models have also been used to demonstrate altered electrogenic responses in sensory neurons and a likely causal link to inflammatory mediators [105,106], and gene transcription changes [107,108] have been suggested. Encouragingly, such findings have been validated in vivo and even associated with the development of chronic pain conditions in nerve compression events [109,110].
In vitro compression chambers are generally custom-built and typically face the challenge of maintaining sufficient gas exchange during pressurization procedures [111]. To circumvent this challenge, compression systems have been designed with a gas–liquid interface, where a continuous gas flow can be maintained to regulate hydrostatic pressure within the system [112]. Further, compression experiments performed on myelinated DRG-Schwann cell co-cultures with regulated gas exchange have demonstrated axonal degeneration and reduced axon numbers at pressures around 0.008 MPa [112]. Similar compression devices at higher pressures of 0.5 MPa in cultured sensory neurons have demonstrated the activation of apoptotic pathways, cytoskeletal abnormalities, disruption of the cell cycle, and the precipitation of necrosis [106,113].

4.1. Advantages

  • Reasonable similarity of in-vitro-induced cellular changes compared to in vivo experiments.

4.2. Disadvantages

  • Model offers limited insight into the biomechanics of the nerve injury because of the difficulty in measuring the strain produced at the site of injury and the rate of application of the strain.
  • Since the biomechanical force of injury varies, it is difficult to predict and control the lesions formed and control which axons are affected and which are spared.

5. Hydrostatic Pressure Models

Microenvironmental tissue hydrostatic pressure (HP) is an important cellular cue that plays a critical role in regulating cell growth, differentiation, migration, and apoptosis, both in vivo and in vitro [114]. In particular, HP is involved in maintaining homeostasis during the development of the optic system, central nervous system, cardiovascular system, cartilage, and bladder tissue. In the CNS, tissue HP is generally between 5 and 15 mmHg [114] but differs substantially based on cell and tissue type. However, in the PNS, HP is approximately 3–5 mm Hg in the dorsal root ganglia and approximately 2–3 mm Hg in peripheral nerves [115]. The HP gradient in the peripheral nerve is maintained by the interstitial fluid within the endoneurial space (EHP), and any alterations in pressure are known to cause severe pathological effects, secondary to endoneurial ischemia, compression induced injury, and edema. These changes ultimately cause the deformation and collapse of blood vessels, release of osmotic solutes from nerve axons, and build-up of fluid at the proximal site of injury, causing severe damage at the injured nerve [116,117,118].
To mimic in vivo conditions and closely model the mechanical forces affecting the cellular microenvironment, various in vitro HP models have been developed where cells are subjected to HP changes via pressurized chambers, hydrostatic fluid columns, and use of a syringe pump system.

5.1. Pressurized Chambers

Pressurized chambers are widely used and typically contain 95% air and 5% CO2 fitted with automatic pressure gauges and controllers to monitor pCO2, pO2, and pH levels within culture systems [114]. Pressure changes are regulated by micro-balloons or custom-made mechanical pressure controllers that are designed to mimic physiological alteration in blood flow for an extended period of time [112,119]. To mimic chronic nerve compression, studies have been designed where DRG neuron-Schwann cell–fibroblast co-cultures were subjected to increased pressure of 40 kPa for 24 h, resulting in decreased neuronal viability and the release of lactate dehydrogenase (LDH) [112]. Neurons subjected to a 7 d high-pressure environment at 18–46 mmHg demonstrated higher levels of LDH compared to cells that were exposed for only 1 day.
Similarly, to understand cytoskeletal changes due to HP forces acting on DRG cell bodies, a gas-pressurized system exerting 0.5 MPa has been shown to induce lower cell density, shorter neurites, and irregular-shaped cells. In addition, microtubule disruption, structural degradation, oxidative stress, apoptosis, and cell cycle arrest were also observed in some neurons [106]. Similar observations related to cellular apoptosis and oxidative stress induced by high HP have also been reported in PC-12 cells [120].

5.2. Hydrostatic Fluid Columns

The generation of HP using static columns of cell culture media with variable column height to create pressure is another innovative in vitro approach. With a change in column height, corresponding changes in the distribution of gases (pCO2, pO2), nutrition, and pH will vary and, thus, experiments can be designed to study the role of varying HP on cells [114]. The HP experienced by cultured cells within columns can be regulated by either placing cells on coverslips at the bottom of medium-containing vertical columns [121] or by circulating the medium via a pressure-controlled peristaltic pump [122]. The latter arrangement also supplies fresh medium and gas to the cells. This set-up has been tried on cortical astrocytes but can be easily extended to peripheral neurons as well [123].

5.3. Syringe Pump System

Another technique to achieve fluctuations in hydrostatic pressure with culture medium is via the syringe pump method. The pressure monitored by a gauge can deliver positive (compressive), negative (tensile), or cyclic HP to the cells [114]. Injury is via a syringe pump, which applies increased hydraulic pressure to the media present in the cell culture chamber. Although not widely used as an in vitro injury platform, it has been observed that the use of different flow rates, needle bore size, and viscosity of the suspension can greatly affect the response of cells in the injection site [124]. This methodology could be further developed to mimic cellular response to injections and changes after peripheral nerve blocks.

5.4. Advantages

  • Cells can be subjected to a wide variety of controlled hydrostatic pressure ranges.
  • In vitro platform helps in understanding the underlying mechanisms of CNS injuries, glaucoma, and cardiovascular research.

5.5. Disadvantages

  • There may be design complexity with the necessity of a pump system in a standard cell culture set-up.
  • Fluid flow may induce shear stress in cells.

6. Metabolic Nerve Injury Models

Metabolic disorders such as diabetes, Gaucher’s disease, and Krabbe disease have been well-known to precipitate progressive neuronal injury and disrupt normal nerve function [125,126,127]. In addition, impaired axonal regeneration is also a common feature that is observed in many metabolic disorders, primarily due to disruption of vascular inputs, inflammatory response, and glia–neuron communication [128]. In particular, damage to the peripheral nerve (peripheral neuropathy) is one of the most common complications observed in patients with long-standing diabetes mellitus. Current animal models of hyperglycemia have shed considerable light on the pathogenesis of neural complications in diabetes. Several in vitro studies have contributed strongly to our understanding of the peripheral neuron response to hyperglycemic conditions. Most studies addressing metabolic cellular injuries fundamentally follow a common principle where neuron and/or non-neuronal cells are exposed to in vitro media conditions that closely mimic specific components of the metabolic disorder in question. More commonly, sensory neurons and glial cells derived from murine models, PC-12 cells, and the SHSY5Y cell lines are utilized for in vitro metabolic injury experiments. In this segment, we review techniques utilized to demonstrate metabolic injury in cells exposed to hyperglycemic conditions, since diabetes is a commonly researched topic.
Exposure of differentiated PC-12 cells to high concentrations (>9 mg/mL) of glucose has been shown to induce the altered release of dopamine [129], enhance reactive oxygen species (ROS), and reduce cell viability and neurite extension [130,131,132]. Similarly, dissociated sensory neurons cultured in glucose concentrations of more than 60 mM showed an increased number of cells that were positive for the injury marker, ATF3 [133]. Interestingly, increased ATF3 positive cells were also observed in sensory neurons that were harvested from rats with experimentally induced hyperglycemia [133], suggesting a strong co-relation between in vivo and in vitro models. Along the same lines, in vitro models involving neurons cultured from animals with experimentally induced hyperglycemia have consistently demonstrated elevated levels of ROS, reduced neurite outgrowth [134], and aberrant calcium ion homeostasis [135].
Apart from sensory neurons, sympathetic neurons harvested from superior cervical ganglia (SCG) and celiac/superior mesenteric ganglia (CG/SMG) are some of the other peripheral neural tissues that are commonly utilized for exploring the effect of hyperglycemia in vitro [136]. It must also be noted that there have been several studies that have combined metabolic injury models with other PNI models, but the decision remains with the investigator to choose the appropriate combination of models to expand on their specific research questions.

6.1. Advantages

  • Inexpensive and easy to perform.
  • Can be utilized to study the effect of hyperglycemia on specific cell types of the peripheral nervous system.

6.2. Disadvantages

  • Pathogenesis of diabetes-induced nerve injury goes beyond merely the hyperglycemia condition. Hence, an in vitro model of hyperglycemic growth media may not exactly mimic the multi-component involvement of metabolic disorders as observed in vivo.

7. Chemically Induced Nerve Injury Models

PNI induced by chemical interaction with neural components is commonly seen in patients under anti-cancer treatment [137], anti-microbials [137,138], and local anesthetics [139,140,141,142]. Chemotherapy-induced peripheral neuropathies (CIPN) are a set of clinical conditions observed in patients administered with chemotherapeutics. CIPN commonly occurs as an undesirable, dose-limiting effect of several chemotherapeutics such as cisplatin, paclitaxel [143,144], vinblastine, and vincristine [145], mainly due to chemically induced damage to axons and cell bodies of peripheral neurons. Though the underlying mechanism of CIPN is not completely known, axonal degeneration or ‘dying back’ is a well-recognized histopathological hallmark [27,146], and understanding the molecular mechanism behind such pathogenesis is crucial to develop therapeutic strategies. In vitro models, especially using dissociated DRG cultures, have been utilized to understand the key cellular features of CIPN. The dose-dependent loss of neurites, fragmentation of neurites, and enlargement of the neuronal cell body have been observed in DRG neurons exposed to bortezomib, cisplatin, eribulin, paclitaxel, and vincristine [147,148,149].
Interestingly, a few studies have also utilized compartmental models to specifically examine the chemically induced dying back effect of axons. Exposure of axons to target chemicals including paclitaxel and vincristine shows clear dysregulation in axonal growth followed by loss of electrical activity [150,151,152]. In addition, in vitro studies in Schwann cell-DRG co-cultures have also shown that anticancer drugs produce a disruption of myelin formation and induce the de-differentiation of Schwann cells to a more immature state, an observation that is consistent with in vivo findings [153].
As with chemotherapeutics-related cellular changes in vitro, neuro-toxic mechanisms associated with the activation of common signaling pathways have also been observed with other drugs, including local anesthetics, alcohol, and phenol derivatives. Toxicity-inducing cellular mechanisms include the activation of common signaling pathways linked to intrinsic caspases [154], Akt [155], and mitogen-activated protein kinase [156]. Notably, the exposure of cultured rat DRG neurons to increasing concentrations of lidocaine and bupivacaine has been shown to induce neuronal apoptosis, Schwann cell death [157,158,159], and mitochondrial depletion via caspase and p38MAPK signaling pathways [160,161].

7.1. Advantages

  • Effect of chemotherapeutics on non-neuronal cells and its subsequent effect on neurons can be studied separately.

7.2. Disadvantages

  • Caution to be exercised when analyzing results from CIPN in vitro models, mainly because cancer, being a multi-component disease, will invariably have unknown variables that may be missed.

8. Blast-Induced Injury Models

Explosion-related blast injuries are a major health issue in military medicine, with improvised explosive devices causing many casualties in human conflicts [162]. Explosions cause unusual injury patterns that may involve skin, muscle, internal organs, and nerves. Intriguingly, an actual blast-linked injury has been shown to have four distinct mechanisms. The initial intense high-pressure impulse, referred to as a blast wave, is known to induce direct soft-tissue injuries, while the secondary, tertiary, and quaternary injuries relate more to injuries due to flying debris, physical displacement, and other events, respectively [163,164,165].
There have been few in vitro studies that have attempted to mimic the effect of the blast wave on peripheral neurons. One such study exposed DRG cultures enclosed within a rubber glove to the impact of a high-energy missile within a water-filled tube [166]. The impact caused high-frequency oscillating pressure waves within the rubber glove that somewhat resembled the blast wave seen during explosions. Interestingly, this study found that within 6 h of impact, almost all DRG neurons demonstrated cytoskeletal disruption, plasma membrane dysfunction, and neurofilament tangles [166]. In another blast injury model that was closer to reality, PC-12 cells submerged in water were subjected to the effect of single and multiple primary blasts generated from research department explosives (RDX). Blast-exposed cells showed distinct axonal beading and decreased cell viability with multiple explosion exposure [167].

8.1. Advantages

  • Functional or biochemical measures can be taken after injury at multiple time points.

8.2. Disadvantages

  • Limitations in stimulating shock wave under special circumstances such as trenches, cabin, and underwater explosions.
  • Experiments with compounds such as RDX may require additional regulatory approvals and specialized safety oversight.

9. Challenges and Future Perspectives

While in vitro model systems of nerve injury are promising due to their simplicity in experimental design, there are, however, two broad challenges that need to be addressed going forward: (a) the importance of the co-existence of neuronal and non-neuronal cells in models, and (b) the limitations of 2D culture systems. As firmly established in the past decade, the pathogenesis of nerve injury is not limited to neurons or glial cells alone. The surrounding non-neuronal cells are key components in determining the level of injury and regeneration [168,169,170]. Secondary injury or secondary damage results from crosstalk between neighboring cells and neurons, which can be either neuroprotective or destructive. For instance, immune cells play a prominent role after initial nerve injury. Immune cells such as macrophages are known to mediate Wallerian degeneration and, hence, a co-culture system that can incorporate immune cells, glia, and neurons may provide more resemblance to an in vivo setting. However, the main challenge remains that it is difficult to maintain long-duration co-cultures of neuronal and non-neuronal cells since the proliferative rates of cells are quite different.
Secondly, cultures in a 2D platform have significant shortcomings. For instance, regeneration ability declines with a higher grade level of injury [100]. Grades III and IV, which include damage to endoneurial tubes along with axons, are known to disrupt the formation of aligned bands of Bungner, required for regeneration. Replicating Grade III and Grade IV types of injury is not possible in 2D culture models. The mechanical properties of neurons cultured in the 3D extracellular matrix considerably vary from 2D culture, which will invariably affect its response to experimental injury. Though microfluidic devices have made it possible to spatially separate the cell body and axon, somewhat mimicking an in vivo design, the 3D architecture is still lacking.
Organoids provide a platform to culture multiple cell types and develop 3D networks that are physiologically closer to tissues in situ. The field of organoid modeling in the peripheral nervous system is gradually developing, with the incorporation of neuro-immune and neuro-vascular cellular interactions. Though exciting, the field of organoid development for the PNS is still in its infancy, and exciting outcomes can be expected in the future

10. Conclusions

PNI coupled with poor nerve regeneration capabilities can be debilitating and compromises overall quality of life. A better understanding of the molecular mechanisms behind such pathogenesis can aid in developing innovative intervention strategies for repair and regeneration. A variety of in vitro models of PNI exist with their own advantages and disadvantages. We observe that the in vitro models are gradually evolving from simple 2D culture systems to complex 3D platforms, and this, we believe, will extend the field of in vitro studies and take us closer to breaking down the complex cogwheel of nerve injury.

Author Contributions

Conceptualization, S.A.S.; data curation, P.V., G.R., P.M. and C.J.; writing—review and editing, P.V., G.R., P.M., C.J. and S.A.S.; visualization, P.V. and G.R.; supervision, S.A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by a CSIR fellowship awarded to P.M, G.R., Department of Science and Technology, Government of India, and support to S.S., Department of Biotechnology, Government of India, grant BT/PR24515/MED/30/1926/2017.

Acknowledgments

We thank Neeraj Katiyar and Jitin Rajappan for their assistance in fabricating the aligned rake pin device shown in this article. We thank the Department of Science and Technology, Government of India, and the Department of Biotechnology, Government of India, for the fellowship and grant support, respectively.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Qian, C.; Zhou, F.-Q. Updates and challenges of axon regeneration in the mammalian central nervous system. J. Mol. Cell Biol. 2020, 12, 798–806. [Google Scholar] [CrossRef]
  2. Rosso, G.; Guck, J. Mechanical changes of peripheral nerve tissue microenvironment and their structural basis during development. APL Bioeng. 2019, 3, 036107. [Google Scholar] [CrossRef] [Green Version]
  3. Park, S.E.; Ahn, J.; Jeong, H.-E.; Youn, I.; Huh, D.; Chung, S. A Three-dimensional in vitro model of the peripheral nervous system. NPG Asia Mater. 2021, 13, 2. [Google Scholar] [CrossRef]
  4. Babetto, E.; Wong, K.M.; Beirowski, B. A glycolytic shift in schwann cells supports injured axons. Nat. Neurosci. 2020, 23, 1215–1228. [Google Scholar] [CrossRef] [PubMed]
  5. Fawcett, J.W. The struggle to make CNS axons regenerate: Why has it been so difficult? Neurochem. Res. 2020, 45, 144–158. [Google Scholar] [CrossRef] [Green Version]
  6. Fawcett, J.W.; Verhaagen, J. Intrinsic determinants of axon regeneration. Devel. Neurobio. 2018, 78, 890–897. [Google Scholar] [CrossRef]
  7. Hussain, G.; Wang, J.; Rasul, A.; Anwar, H.; Qasim, M.; Zafar, S.; Aziz, N.; Razzaq, A.; Hussain, R.; de Aguilar, J.L.G.; et al. Current status of therapeutic approaches against peripheral nerve injuries: A detailed story from injury to recovery. Int. J. Biol. Sci. 2020, 16, 116–134. [Google Scholar] [CrossRef] [Green Version]
  8. Heary, K.O.; Wong, A.W.K.; Lau, S.C.L.; Dengler, J.; Thompson, M.R.; Crock, L.W.; Novak, C.B.; Philip, B.A.; Mackinnon, S.E. Quality of life and psychosocial factors as predictors of pain relief following nerve surgery. Hand 2020, 155894472091121. [Google Scholar] [CrossRef] [PubMed]
  9. Houshyar, S.; Bhattacharyya, A.; Shanks, R. Peripheral nerve conduit: Materials and structures. ACS Chem. Neurosci. 2019, 10, 3349–3365. [Google Scholar] [CrossRef]
  10. Gopalakrishnan-Prema, V.; Mohanan, A.; Shivaram, S.B.; Madhusudanan, P.; Raju, G.; Menon, D.; Shankarappa, S.A. Electrical stimulation of co-woven nerve conduit for peripheral neurite differentiation. Biomed. Mater. 2020, 15, 065015. [Google Scholar] [CrossRef]
  11. Raza, C.; Riaz, H.A.; Anjum, R.; Shakeel, N. ul A. Repair strategies for injured peripheral nerve: Review. Life Sci. 2020, 243, 117308. [Google Scholar] [CrossRef]
  12. Isaacs, J.; Cochran, A.R. Nerve transfers for peripheral nerve injury in the upper limb: A case-based review. Bone Jt. J. 2019, 101, 124–131. [Google Scholar] [CrossRef]
  13. Mohanty, C.; Bhat, D.; Devi, B. Use of animal models in peripheral nerve surgery and research. Neurol. India 2019, 67, 100. [Google Scholar] [CrossRef]
  14. Vela, F.; Martínez-Chacón, G.; Ballestín, A.; Campos, J.; Sánchez-Margallo, F.; Abellán, E. Animal models used to study direct peripheral nerve repair: A systematic review. Neural. Regen. Res. 2020, 15, 491. [Google Scholar] [CrossRef] [PubMed]
  15. Montanino, A.; Kleiven, S. Utilizing a structural mechanics approach to assess the primary effects of injury loads onto the axon and its components. Front. Neurol. 2018, 9, 643. [Google Scholar] [CrossRef] [PubMed]
  16. Romeo-Guitart, D.; Casas, C. Network-centric medicine for peripheral nerve injury: Treating the whole to boost endogenous mechanisms of neuroprotection and regeneration. Neural. Regen. Res. 2019, 14, 1122. [Google Scholar] [CrossRef] [PubMed]
  17. Lee, B.; Cho, Y. Experimental model systems for understanding human axonal injury responses. Int. J. Mol. Sci. 2021, 22, 474. [Google Scholar] [CrossRef]
  18. Rigoni, M.; Montecucco, C. Animal models for studying motor axon terminal paralysis and recovery. J. Neurochem. 2017, 142, 122–129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Wright, R.M.; Post, A.; Hoshizaki, B.; Ramesh, K.T. A multiscale computational approach to estimating axonal damage under inertial loading of the head. J. Neurotrauma 2013, 30, 102–118. [Google Scholar] [CrossRef]
  20. Höke, A. Mechanisms of disease: What factors limit the success of peripheral nerve regeneration in humans? Nat. Clin. Pract. Neurol. 2006, 2, 448–454. [Google Scholar] [CrossRef]
  21. Navarro, X.; Vivó, M.; Valero-Cabré, A. Neural plasticity after peripheral nerve injury and regeneration. Prog. Neurobiol. 2007, 82, 163–201. [Google Scholar] [CrossRef]
  22. Elsayed, H.; Faroni, A.; Ashraf, M.R.; Osuji, J.; Wunderley, L.; Zhang, L.; Elsobky, H.; Mansour, M.; Zidan, A.S.; Reid, A.J. Development and characterisation of an in vitro model of wallerian degeneration. Front. Bioeng. Biotechnol. 2020, 8, 784. [Google Scholar] [CrossRef] [PubMed]
  23. Jessen, K.R.; Arthur-Farraj, P. Repair schwann cell update: Adaptive reprogramming, EMT, and stemness in regenerating nerves. Glia 2019, 67, 421–437. [Google Scholar] [CrossRef]
  24. Pellegatta, M.; Taveggia, C. The complex work of proteases and secretases in wallerian degeneration: Beyond neuregulin-1. Front. Cell. Neurosci. 2019, 13, 93. [Google Scholar] [CrossRef] [PubMed]
  25. Modrak, M.; Talukder, M.A.H.; Gurgenashvili, K.; Noble, M.; Elfar, J.C. Peripheral Nerve Injury and Myelination: Potential Therapeutic Strategies. J. Neurosci. Res. 2020, 98, 780–795. [Google Scholar] [CrossRef]
  26. Renner, M.; Stute, G.; Alzureiqi, M.; Reinhard, J.; Wiemann, S.; Schmid, H.; Faissner, A.; Dick, H.B.; Joachim, S.C. Optic nerve degeneration after retinal ischemia/reperfusion in a rodent model. Front. Cell. Neurosci. 2017, 11, 254. [Google Scholar] [CrossRef] [Green Version]
  27. Fukuda, Y.; Li, Y.; Segal, R.A. A Mechanistic understanding of axon degeneration in chemotherapy-induced peripheral neuropathy. Front. Neurosci. 2017, 11, 481. [Google Scholar] [CrossRef]
  28. Girouard, M.P.; Simas, T.; Hua, L.; Morquette, B.; Khazaei, M.R.; Unsain, N.; Johnstone, A.D.; Rambaldi, I.; Sanz, R.L.; Di Raddo, M.E.; et al. Collapsin response mediator protein 4 (Crmp4) facilitates wallerian degeneration and axon regeneration following sciatic nerve injury. Eneuro 2020, 7, 1–17. [Google Scholar] [CrossRef]
  29. Rodemer, W.; Selzer, M.E. Role of axon resealing in retrograde neuronal death and regeneration after spinal cord injury. Neural Regen. Res. 2019, 14, 399–404. [Google Scholar] [CrossRef]
  30. Canty, A.J.; Jackson, J.S.; Huang, L.; Trabalza, A.; Bass, C.; Little, G.; Tortora, M.; Khan, S.; De Paola, V. In vivo imaging of injured cortical axons reveals a rapid onset form of wallerian degeneration. BMC Biol. 2020, 18, 170. [Google Scholar] [CrossRef]
  31. Kerstein, P.C.; Patel, K.M.; Gomez, T.M. Calpain-mediated proteolysis of talin and FAK regulates adhesion dynamics necessary for axon guidance. J. Neurosci. 2017, 37, 1568–1580. [Google Scholar] [CrossRef] [Green Version]
  32. Blanquie, O.; Bradke, F. Cytoskeleton dynamics in axon regeneration. Curr. Opin. Neurobiol. 2018, 51, 60–69. [Google Scholar] [CrossRef] [PubMed]
  33. Gu, C. Rapid and reversible development of axonal varicosities: A new form of neural plasticity. Front. Mol. Neurosci. 2021, 14, 1. [Google Scholar] [CrossRef] [PubMed]
  34. Hill, C.E. A view from the ending: Axonal dieback and regeneration following SCI. Neurosci. Lett. 2017, 652, 11–24. [Google Scholar] [CrossRef] [PubMed]
  35. Gu, Y.; Jukkola, P.; Wang, Q.; Esparza, T.; Zhao, Y.; Brody, D.; Gu, C. Polarity of varicosity initiation in central neuron mechanosensation. J. Cell Biol. 2017, 216, 2179–2199. [Google Scholar] [CrossRef] [Green Version]
  36. Shrirao, A.B.; Kung, F.H.; Omelchenko, A.; Schloss, R.S.; Boustany, N.N.; Zahn, J.D.; Yarmush, M.L.; Firestein, B.L. Microfluidic platforms for the study of neuronal injury in vitro. Biotechnol. Bioeng. 2018, 115, 815–830. [Google Scholar] [CrossRef]
  37. Wahane, S.; Halawani, D.; Zhou, X.; Zou, H. Epigenetic regulation of axon regeneration and glial activation in injury responses. Front. Genet. 2019, 10, 640. [Google Scholar] [CrossRef]
  38. Nagendran, T.; Larsen, R.S.; Bigler, R.L.; Frost, S.B.; Philpot, B.D.; Nudo, R.J.; Taylor, A.M. Distal axotomy enhances retrograde presynaptic excitability onto injured pyramidal neurons via trans-synaptic signaling. Nat. Commun. 2017, 8, 625. [Google Scholar] [CrossRef] [Green Version]
  39. Li, Y.; Li, C.; Gan, C.; Zhao, K.; Chen, J.; Song, J.; Lei, T. A Precise, controllable in vitro model for diffuse axonal injury through uniaxial stretch injury. Front. Neurosci. 2019, 13, 1063. [Google Scholar] [CrossRef] [Green Version]
  40. Nagendran, T.; Taylor, A.M. Unique axon-to-soma signaling pathways mediate dendritic spine loss and hyper-excitability post-axotomy. Front. Cell. Neurosci. 2019, 13, 431. [Google Scholar] [CrossRef] [Green Version]
  41. Davies, A.J.; Kim, H.W.; Gonzalez-Cano, R.; Choi, J.; Back, S.K.; Roh, S.E.; Johnson, E.; Gabriac, M.; Kim, M.S.; Lee, J.; et al. Natural killer cells degenerate intact sensory afferents following nerve injury. Cell 2019, 176, 716–728. [Google Scholar] [CrossRef] [Green Version]
  42. Davies, A.J.; Rinaldi, S.; Costigan, M.; Oh, S.B. Cytotoxic immunity in peripheral nerve injury and pain. Front. Neurosci. 2020, 14, 142. [Google Scholar] [CrossRef]
  43. Boerboom, A.; Dion, V.; Chariot, A.; Franzen, R. Molecular mechanisms involved in schwann cell plasticity. Front. Mol. Neurosci. 2017, 10, 38. [Google Scholar] [CrossRef] [Green Version]
  44. Ma, J.J.; Xu, R.J.; Ju, X.; Wang, W.H.; Luo, Z.P.; Liu, C.M.; Yang, L.; Li, B.; Chen, J.Q.; Meng, B.; et al. The telomerase reverse transcriptase (TERT) and P53 regulate mammalian PNS and CNS axon regeneration downstream of c-Myc. Biorxiv 2019, 39, 9107–9118. [Google Scholar] [CrossRef]
  45. Campenot, R.B. Local control of neurite development by nerve growth factor. Proc. Natl. Acad. Sci. USA 1977, 74, 4516–4519. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Campenot, R.B.; Lund, K.; Mok, S.A. Production of compartmented cultures of rat sympathetic neurons. Nat. Protoc. 2009, 4, 1869–1887. [Google Scholar] [CrossRef]
  47. Tierney, W.; Vicino, I.; Sun, S.; Chiu, W.; Engel, E.; Taylor, M.; Hogue, I. Methods and applications of campenot trichamber neuronal cultures for the study of neuroinvasive viruses. Life Sci. 2021. [Google Scholar] [CrossRef]
  48. Fantuzzo, J.A.; Hart, R.P.; Zahn, J.D.; Pang, Z.P. Compartmentalized devices as tools for investigation of human brain network dynamics: Compartmentalized devices to study human brain network dynamics. Dev. Dyn. 2019, 248, 65–77. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Klusch, A.; Gorzelanny, C.; Reeh, P.W.; Schmelz, M.; Petersen, M.; Sauer, S.K. Local NGF and GDNF levels modulate morphology and function of porcine DRG neurites, in vitro. PLoS ONE 2018, 13, e0203215. [Google Scholar] [CrossRef]
  50. Bertrand, J.; Winton, M.J.; Rodriguez-Hernandez, N.; Campenot, R.B.; McKerracher, L. Application of Rho antagonist to neuronal cell bodies promotes neurite growth in compartmented cultures and regeneration of retinal ganglion cell axons in the optic nerve of adult rats. J. Neurosci. 2005, 25, 1113–1121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Kimpinski, K.; Campenot, R.B.; Mearow, K. Effects of the neurotrophins nerve growth factor, neurotrophin-3, and brain-derived neurotrophic factor (BDNF) on neurite growth from adult sensory neurons in compartmented cultures. J. Neurobiol. 1997, 33, 395–410. [Google Scholar] [CrossRef]
  52. MacInnis, B.L.; Senger, D.L.; Campenot, R.B. Spatial requirements for TrkA kinase activity in the support of neuronal survival and axon growth in rat sympathetic neurons. Neuropharmacology 2003, 45, 995–1010. [Google Scholar] [CrossRef]
  53. Ohtake, Y.; Matsuhisa, K.; Kaneko, M.; Kanemoto, S.; Asada, R.; Imaizumi, K.; Saito, A. Axonal activation of the unfolded protein response promotes axonal regeneration following peripheral nerve injury. Neuroscience 2018, 375, 34–48. [Google Scholar] [CrossRef] [PubMed]
  54. Jocher, G.; Mannschatz, S.H.; Offterdinger, M.; Schweigreiter, R. Microfluidics of small-population neurons allows for a precise quantification of the peripheral axonal growth state. Front. Cell. Neurosci. 2018, 12, 166. [Google Scholar] [CrossRef] [Green Version]
  55. Katiyar, N.; Raju, G.; Madhusudanan, P.; Gopalakrishnan-Prema, V.; Shankarappa, S.A. Neuronal delivery of nanoparticles via nerve fibres in the skin. Sci. Rep. 2021, 11, 2566. [Google Scholar] [CrossRef]
  56. Atmaramani, R.; Veeramachaneni, S.; Mogas, L.V.; Koppikar, P.; Black, B.J.; Hammack, A.; Pancrazio, J.J.; Granja-Vazquez, R. Investigating the function of adult DRG neuron axons using an in vitro microfluidic culture system. Micromachines 2021, 12, 1317. [Google Scholar] [CrossRef]
  57. Madhusudanan, P.; Jerard, C.; Katiyar, N.; Raju, G.; Shankarappa, S.A. Effect of gold nanoparticle treated dorsal root ganglion cells on peripheral neurite differentiation. Toxicol. Vitr. 2021, 74, 105175. [Google Scholar] [CrossRef] [PubMed]
  58. Roselló-Busquets, C.; de la Oliva, N.; Martínez-Mármol, R.; Hernaiz-Llorens, M.; Pascual, M.; Muhaisen, A.; Navarro, X.; del Valle, J.; Soriano, E. Cholesterol depletion regulates axonal growth and enhances central and peripheral nerve regeneration. Front. Cell. Neurosci. 2019, 13, 40. [Google Scholar] [CrossRef] [Green Version]
  59. Yong, Y.; Hughes, C.; Deppmann, C. A microfluidic culture platform to assess axon degeneration. Methods Mol. Biol. 2020, 83–86. [Google Scholar] [CrossRef]
  60. Osaki, T.; Uzel, S.G.M.; Kamm, R.D. Microphysiological 3D model of amyotrophic lateral sclerosis (ALS) from human IPS-derived muscle cells and optogenetic motor neurons. Sci. Adv. 2018, 4, 1–16. [Google Scholar] [CrossRef] [Green Version]
  61. Spijkers, X.M.; Pasteuning-Vuhman, S.; Dorleijn, J.C.; Vulto, P.; Wevers, N.R.; Pasterkamp, R.J. A Directional 3D neurite outgrowth model for studying motor axon biology and disease. Sci. Rep. 2021, 11, 2080. [Google Scholar] [CrossRef]
  62. Vysokov, N.; McMahon, S.B.; Raouf, R. The role of NaV channels in synaptic transmission after axotomy in a microfluidic culture platform. Sci. Rep. 2019, 9, 12915. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Zhang, J.N.; Michel, U.; Lenz, C.; Friedel, C.C.; Köster, S.; D’Hedouville, Z.; Tönges, L.; Urlaub, H.; Bähr, M.; Lingor, P.; et al. Calpain-mediated cleavage of collapsin response mediator protein-2 drives acute axonal degeneration. Sci. Rep. 2016, 6, 37050. [Google Scholar] [CrossRef]
  64. Sala-Jarque, J.; Mesquida-Veny, F.; Badiola-Mateos, M.; Samitier, J.; Hervera, A.; Del Río, J.A. Neuromuscular activity induces paracrine signaling and triggers axonal regrowth after injury in microfluidic lab-on-chip devices. Cells 2020, 9, 302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Tecoma, E.S.; Monyer, H.; Goldberg, M.P.; Choi, D.W. Traumatic neuronal injury in vitro is attenuated by NMDA antagonists. Neuron 1989, 2, 1541–1545. [Google Scholar] [CrossRef]
  66. Payette, D.J.; Xie, J.; Shirwany, N.; Guo, Q. Exacerbation of apoptosis of cortical neurons following traumatic brain injury in Par-4 transgenic mice. Int. J. Clin. Exp. Pathol. 2008, 1, 44–56. [Google Scholar] [PubMed]
  67. Datar, A.; Ameeramja, J.; Bhat, A.; Srivastava, R.; Mishra, A.; Bernal, R.; Prost, J.; Callan-Jones, A.; Pullarkat, P.A. The roles of microtubules and membrane tension in axonal beading, retraction, and atrophy. Biophys. J. 2019, 117, 880–891. [Google Scholar] [CrossRef]
  68. Üstün, R.; Oğuz, E.K.; Şeker, A.; Korkaya, H. Thymoquinone protects DRG neurons from axotomy-induced cell death. Neurol. Res. 2018, 40, 930–937. [Google Scholar] [CrossRef]
  69. Gross, G.W.; Lucas, J.H.; Higgins, M.L. Laser microbeam surgery: Ultrastructural changes associated with neurite transection in culture. J. Neurosci. 1983, 3, 1979–1993. [Google Scholar] [CrossRef] [Green Version]
  70. Cengiz, N.; Öztürk, G.; Erdoǧan, E.; Him, A.; Oǧuz, E.K. Consequences of neurite transection in vitro. J. Neurotrauma 2012, 29, 2465–2474. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Kunik, D.; Dion, C.; Ozaki, T.; Levin, L.A.; Costantino, S. Laser-based single-axon transection for high-content axon injury and regeneration studies. PLoS ONE 2011, 6, e26832. [Google Scholar] [CrossRef] [Green Version]
  72. Shao, X.; You, R.; Hui, T.H.; Fang, C.; Gong, Z.; Yan, Z.; Chang, R.C.C.; Shenoy, V.B.; Lin, Y. Tension- and adhesion-regulated retraction of injured axons. Biophys. J. 2019, 117, 193–202. [Google Scholar] [CrossRef]
  73. Song, Y.; Zhang, M.; Tao, X.; Xu, Z.; Zheng, Y.; Zhu, M.; Zhang, L.; Qiao, J.; Gao, L. Difference of acute dissociation and 1-day culture on the electrophysiological properties of rat dorsal root ganglion neurons. J. Physiol. Biochem. 2018, 74, 207–221. [Google Scholar] [CrossRef]
  74. Wall, E.J.; Massie, J.B.; Kwan, M.K.; Rydevik, B.L.; Myers, R.R.; Garfin, S.R. experimental stretch neuropathy: Changes in nerve conduction under tension. J. Bone Jt. Surg. Br. Vol. 1992, 74, 126–129. [Google Scholar] [CrossRef] [PubMed]
  75. Loverde, J.R.; Pfister, B.J. Developmental axon stretch stimulates neuron growth while maintaining normal electrical activity, intracellular calcium flux, and somatic morphology. Front. Cell. Neurosci. 2015, 9, 308. [Google Scholar] [CrossRef] [Green Version]
  76. Umansky, D.; Midha, R. Commentary: Rapid-stretch injury to peripheral nerves: Histologic results. Neurosurgery 2019, 86, nyz196. [Google Scholar] [CrossRef] [Green Version]
  77. Mahan, M.A.; Yeoh, S.; Monson, K.; Light, A. Rapid stretch injury to peripheral nerves: Biomechanical results. Neurosurgery 2019, 85, E137–E144. [Google Scholar] [CrossRef] [PubMed]
  78. Yap, Y.C.; King, A.E.; Guijt, R.M.; Jiang, T.; Blizzard, A.; Breadmore, M.C.; Dickson, T.C. Mild and repetitive very mild axonal stretch injury triggers cystoskeletal mislocalization and growth cone collapse. PLoS ONE 2017, 12, e0176997. [Google Scholar] [CrossRef]
  79. Zhou, M.; Hu, M.; He, S.; Li, B.; Liu, C.; Min, J.; Hong, L. Effects of RSC96 schwann cell-derived exosomes on proliferation, senescence, and apoptosis of dorsal root ganglion cells in vitro. Med. Sci. Monit. 2018, 24, 7841–7849. [Google Scholar] [CrossRef] [PubMed]
  80. Aomura, S.; Nakadate, H.; Kaneko, Y.; Nishimura, A.; Willinger, R. Stretch-induced functional disorder of axonal transport in the cultured rat cortex neuron. Integr. Mol. Med. 2016, 3, 654–660. [Google Scholar] [CrossRef] [Green Version]
  81. Ahmed, S.M.; Rzigalinski, B.A.; Willoughby, K.A.; Sitterding, H.A.; Ellis, E.F. Stretch-induced injury alters mitochondrial membrane potential and cellular ATP in cultured astrocytes and neurons. J. Neurochem. 2000, 74, 1951–1960. [Google Scholar] [CrossRef] [PubMed]
  82. Ellis, E.F.; Mckinney, J.S.; Willoughby, K.A.; Liang, S.; Povlishock, J.T. A new model for rapid stretch-induced injury of cells in culture: Characterization of the model using astrocytes. J. Neurotrauma 1995, 12, 325–339. [Google Scholar] [CrossRef]
  83. Bianchi, F.; Malboubi, M.; George, J.H.; Jerusalem, A.; Thompson, M.S.; Ye, H. Ion current and action potential alterations in peripheral neurons subject to uniaxial strain. J. Neurosci. Res. 2019, 97, 744–751. [Google Scholar] [CrossRef] [PubMed]
  84. Katanosaka, K.; Takatsu, S.; Mizumura, K.; Naruse, K.; Katanosaka, Y. TRPV2 is required for mechanical nociception and the stretch-evoked response of primary sensory neurons. Sci. Rep. 2018, 8, 16782. [Google Scholar] [CrossRef] [PubMed]
  85. Lin, Y.W.; Cheng, C.M.; LeDuc, P.R.; Chen, C.C. Understanding sensory nerve mechanotransduction through localized elastomeric matrix control. PLoS ONE 2009, 4, e4293. [Google Scholar] [CrossRef]
  86. Tavalin, S.J.; Ellis, E.F.; Satin, L.S. Mechanical perturbation of cultured cortical neurons reveals a stretch-induced delayed depolarization. J. Neurophysiol. 1995, 74, 2767–2773. [Google Scholar] [CrossRef]
  87. Chetta, J.; Kye, C.; Shah, S.B. Cytoskeletal dynamics in response to tensile loading of mammalian axons. Cytoskeleton 2010, 67, 650–665. [Google Scholar] [CrossRef]
  88. Mutalik, S.P.; Joseph, J.; Pullarkat, P.A.; Ghose, A. Cytoskeletal mechanisms of axonal contractility. Biophys. J. 2018, 115, 713–724. [Google Scholar] [CrossRef] [Green Version]
  89. Loverde, J.R.; Ozoka, V.C.; Aquino, R.; Lin, L.; Pfister, B.J. Live imaging of axon stretch growth in embryonic and adult neurons. J. Neurotrauma 2011, 28, 2389–2403. [Google Scholar] [CrossRef]
  90. Tsantoulas, C.; Farmer, C.; Machado, P.; Baba, K.; McMahon, S.B.; Raouf, R. Probing functional properties of nociceptive axons using a microfluidic culture system. PLoS ONE 2013, 8, e80722. [Google Scholar] [CrossRef]
  91. Kampanis, V.; Tolou-Dabbaghian, B.; Zhou, L.; Roth, W.; Puttagunta, R. Cyclic stretch of either PNS or CNS located nerves can stimulate neurite outgrowth. Cells 2020, 10, 32. [Google Scholar] [CrossRef] [PubMed]
  92. Mahan, M.A. Nerve stretching: A history of tension. J. Neurosurg. 2020, 132, 252–259. [Google Scholar] [CrossRef] [PubMed]
  93. Katiyar, K.S.; Struzyna, L.A.; Das, S.; Cullen, D.K. Stretch growth of motor axons in custom mechanobioreactors to generate long-projecting axonal constructs. J. Tissue Eng. Regen. Med. 2019, 13, 2040–2054. [Google Scholar] [CrossRef] [PubMed]
  94. Gladman, S.J.; Ward, R.E.; Michael-Titus, A.T.; Knight, M.M.; Priestley, J.V. The effect of mechanical strain or hypoxia on cell death in subpopulations of rat dorsal root ganglion neurons in vitro. Neuroscience 2010, 171, 577–587. [Google Scholar] [CrossRef]
  95. Ji, J.; Tyurina, Y.Y.; Tang, M.; Feng, W.; Stolz, D.B.; Clark, R.S.B.; Meaney, D.F.; Kochanek, P.M.; Kagan, V.E.; Bayír, H. Mitochondrial injury after mechanical stretch of cortical neurons in vitro: Biomarkers of apoptosis and selective peroxidation of anionic phospholipids. J. Neurotrauma 2012, 29, 776–788. [Google Scholar] [CrossRef] [Green Version]
  96. Xu, Z.; Liu, Y.; Yang, D.; Yuan, F.; Ding, J.; Chen, H.; Tian, H. Sesamin protects SH-SY5Y cells against mechanical stretch injury and promoting cell survival. BMC Neurosci. 2017, 18, 57. [Google Scholar] [CrossRef] [Green Version]
  97. Mutalik, S.P.; Ghose, A. Axonal cytomechanics in neuronal development. J. Biosci. 2020, 45, 64. [Google Scholar] [CrossRef]
  98. Faktorovich, S.; Filatov, A.; Rizvi, Z. Common compression neuropathies. Clin. Geriatr. Med. 2021, 37, 241–252. [Google Scholar] [CrossRef]
  99. Lee, E.Y.; Lim, A.Y.T. Nerve compression in the upper limb. Clin. Plast. Surg. 2019, 46, 285–293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Menorca, R.M.G.; Fussell, T.S.; Elfar, J.C. Peripheral nerve trauma: Mechanisms of injury and recovery. Hand Clin. 2013, 29, 317–330. [Google Scholar] [CrossRef] [Green Version]
  101. Tapadia, M.; Mozaffar, T.; Gupta, R. Compressive neuropathies of the upper extremity: Update on pathophysiology, classification, and electrodiagnostic findings. J. Hand Surg. 2010, 35, 668–677. [Google Scholar] [CrossRef] [Green Version]
  102. Sugawara, O.; Atsuta, Y.; Iwahara, T.; Muramoto, T.; Watakabe, M.; Takemitsu, Y. The effects of mechanical compression and hypoxia on nerve root and dorsal root ganglia: An analysis of ectopic firing using an in vitro model. Spine 1996, 21, 2089–2094. [Google Scholar] [CrossRef]
  103. Amir, R.; Kocsis, J.D.; Devor, M. Multiple interacting sites of ectopic spike electrogenesis in primary sensory neurons. J. Neurosci. 2005, 25, 2576–2585. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Ma, C.; LaMotte, R.H. Multiple sites for generation of ectopic spontaneous activity in neurons of the chronically compressed dorsal root ganglion. J. Neurosci. 2007, 27, 14059–14068. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Ma, C.; Greenquist, K.W.; LaMotte, R.H. Inflammatory mediators enhance the excitability of chronically compressed dorsal root ganglion neurons. J. Neurophysiol. 2006, 95, 2098–2107. [Google Scholar] [CrossRef] [Green Version]
  106. Ye, Z.; Wang, Y.; Quan, X.; Li, J.; Hu, X.; Huang, J.; Luo, Z. Effects of mechanical force on cytoskeleton structure and calpain-induced apoptosis in rat dorsal root ganglion neurons in vitro. PLoS ONE 2012, 7, e52183. [Google Scholar] [CrossRef]
  107. Ewan, E.E.; Avraham, O.; Carlin, D.; Gonçalves, T.M.; Zhao, G.; Cavalli, V. Ascending dorsal column sensory neurons respond to spinal cord injury and downregulate genes related to lipid metabolism. Sci. Rep. 2021, 11, 374. [Google Scholar] [CrossRef] [PubMed]
  108. Martin, S.; Reid, A.; Verkhratsky, A.; Magnaghi, V.; Faroni, A. Gene Expression changes in dorsal root ganglia following peripheral nerve injury: Roles in inflammation, cell death and nociception. Neural Regen. Res. 2019, 14, 939–947. [Google Scholar] [CrossRef]
  109. 1Wang, T.; Hurwitz, O.; Shimada, S.G.; Qu, L.; Fu, K.; Zhang, P.; Ma, C.; LaMotte, R.H. Chronic compression of the dorsal root ganglion enhances mechanically evoked pain behavior and the activity of cutaneous nociceptors in mice. PLoS ONE 2015, 10, e0137512. [Google Scholar] [CrossRef]
  110. Zhao, J.Y.; Liang, L.; Gu, X.; Li, Z.; Wu, S.; Sun, L.; Atianjoh, F.E.; Feng, J.; Mo, K.; Jia, S.; et al. DNA methyltransferase DNMT3a contributes to neuropathic pain by repressing kcna2 in primary afferent neurons. Nat. Commun. 2017, 8, 14712. [Google Scholar] [CrossRef]
  111. Myers, K.A.; Rattner, J.B.; Shrive, N.G.; Hart, D.A.; Myers, K.A.; Rattner, J.B.; Shrive, N.G.; Hart, D.A. Hydrostatic pressure sensation in cells: Integration into the tensegrity model. Biochem. Cell Biol. 2007, 85, 543–551. [Google Scholar] [CrossRef] [PubMed]
  112. Frieboes, L.R.; Gupta, R. An in-vitro traumatic model to evaluate the response of myelinated cultures to sustained hydrostatic compression injury. J. Neurotrauma 2009, 26, 2245–2256. [Google Scholar] [CrossRef] [Green Version]
  113. Quan, X.; Guo, K.; Wang, Y.; Huang, L.; Chen, B.; Ye, Z.; Luo, Z. Mechanical compression insults induce nanoscale changes of membrane-skeleton arrangement which could cause apoptosis and necrosis in dorsal root ganglion neurons. Biosci. Biotechnol. Biochem. 2014, 78, 1631–1639. [Google Scholar] [CrossRef] [Green Version]
  114. Liu, S.; Tao, R.; Wang, M.; Tian, J.; Genin, G.M.; Lu, T.J.; Xu, F. Regulation of cell behavior by hydrostatic pressure. Appl. Mech. Rev. 2019, 71, 40803. [Google Scholar] [CrossRef]
  115. Ask, P.; Levitan, H.; Robinson, P.J.; Rapoport, S.I. Peripheral nerve as an osmometer: Role of endoneurial capillaries in frog sciatic nerve. Am. J. Physiol. Cell Physiol. 1983, 244, C75–C81. [Google Scholar] [CrossRef]
  116. Lim, T.K.Y.; Shi, X.Q.; Johnson, J.M.; Rone, M.B.; Antel, J.P.; David, S.; Zhang, J. Peripheral nerve injury induces persistent vascular dysfunction and endoneurial hypoxia, contributing to the genesis of neuropathic pain. J. Neurosci. 2015, 35, 3346–3359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Mizisin, A.P.; Weerasuriya, A. Homeostatic Regulation of the endoneurial microenvironment during development, aging and in response to trauma, disease and toxic insult. Acta Neuropathol. 2011, 121, 291–312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Myers, R.R.; Rydevik, B.L.; Heckman, H.M.; Powell, H.C. Proximodistal gradient in endoneurial fluid pressure. Exp. Neurol. 1988, 102, 368–370. [Google Scholar] [CrossRef]
  119. Sinar, E.J.; Mendelow, A.D.; Graham, D.I.; Teasdale, G.M. Experimental intracerebral hemorrhage: Effects of a temporary mass lesion. J. Neurosurg. 1987, 66, 568–576. [Google Scholar] [CrossRef]
  120. Tök, L.; Naziroʇlu, M.; Uʇuz, A.C.; Tök, Ö. Elevated hydrostatic pressures induce apoptosis and oxidative stress through mitochondrial membrane depolarization in PC12 neuronal cells: A cell culture model of glaucoma. J. Recept. Signal Transduct. 2014, 34, 410–416. [Google Scholar] [CrossRef]
  121. Lei, Y.; Rajabi, S.; Pedrigi, R.M.; Overby, D.R.; Thomas Read, A.; Ross Ethier, C. In vitro models for glaucoma research: Effects of hydrostatic pressure. Investig. Ophthalmol. Vis. Sci. 2011, 52, 6329–6339. [Google Scholar] [CrossRef]
  122. Mandal, A.; Shahidullah, M.; Delamere, N.A. Hydrostatic pressure-induced release of stored calcium in cultured rat optic nerve head astrocytes. Investig. Ophthalmol. Vis. Sci. 2010, 51, 3129–3138. [Google Scholar] [CrossRef] [Green Version]
  123. Guo, T.; Bian, Z.; Trocki, K.; Chen, L.; Zheng, G.; Feng, B. Optical recording reveals topological distribution of functionally classified colorectal afferent neurons in intact lumbosacral DRG. Physiol. Rep. 2019, 7, e14097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Wahlberg, B.; Ghuman, H.; Liu, J.R.; Modo, M. Ex vivo biomechanical characterization of syringe-needle ejections for intracerebral cell delivery. Sci. Rep. 2018, 8, 9194. [Google Scholar] [CrossRef] [PubMed]
  125. Pastores, G.M. Neuropathic gaucher disease. Wien. Med. Wochenschr. 2010, 160, 605–608. [Google Scholar] [CrossRef] [PubMed]
  126. Feldman, E.L.; Callaghan, B.C.; Pop-Busui, R.; Zochodne, D.W.; Wright, D.E.; Bennett, D.L.; Bril, V.; Russell, J.W.; Viswanathan, V. Diabetic neuropathy. Nat. Rev. Dis. Primers 2019, 5, 41. [Google Scholar] [CrossRef]
  127. Yang, H.; Sloan, G.; Ye, Y.; Wang, S.; Duan, B.; Tesfaye, S.; Gao, L. New perspective in diabetic neuropathy: From the periphery to the brain, a call for early detection, and precision medicine. Front. Endocrinol. 2020, 10, 929. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Yagihashi, S.; Mizukami, H.; Sugimoto, K. Mechanism of diabetic neuropathy: Where are we now and where to go? J. Diabetes Investig. 2011, 2, 18–32. [Google Scholar] [CrossRef] [Green Version]
  129. Koshimura, K.; Tanaka, J.; Murakami, Y.; Kato, Y. Effect of high concentration of glucose on dopamine release from pheochromocytoma-12 cells. Metabolism 2003, 52, 922–926. [Google Scholar] [CrossRef]
  130. Koshimura, K.; Tanaka, J.; Murakami, Y.; Kato, Y. Involvement of nitric oxide in glucose toxicity on differentiated pc12 cells: Prevention of glucose toxicity by tetrahydrobiopterin, a cofactor for nitric oxide synthase. Neurosci. Res. 2002, 43, 31–38. [Google Scholar] [CrossRef]
  131. Shi, Y.; Liang, X.; Zhang, H.; Sun, Q.; Wu, Q.; Qu, L. Combination of quercetin, cinnamaldehyde and hirudin protects rat dorsal root ganglion neurons against high glucose-induced injury through Nrf-2/HO-1 activation and NF-ΚB inhibition. Chin. J. Integr. Med. 2017, 23, 663–671. [Google Scholar] [CrossRef] [PubMed]
  132. Liu, Y.; Zhang, S.; Xue, J.; Wei, Z.; Ao, P.; Shen, B.; Ding, L. CGRP reduces apoptosis of DRG cells induced by high-glucose oxidative stress injury through PI3K/AKT induction of heme oxygenase-1 and Nrf-2 expression. Oxidative Med. Cell. Longev. 2019, 2019, 1–9. [Google Scholar] [CrossRef]
  133. Peeraer, E.; van Lutsenborg, A.; Verheyen, A.; de Jongh, R.; Nuydens, R.; Meert, T.F. Pharmacological evaluation of rat dorsal root ganglion neurons as an in vitro model for diabetic neuropathy. J. Pain Res. 2011, 4, 55–65. [Google Scholar] [CrossRef] [Green Version]
  134. Zherebitskaya, E.; Akude, E.; Smith, D.R.; Fernyhough, P. Development of selective axonopathy in adult sensory neurons isolated from diabetic rats: Role of glucose-induced oxidative stress. Diabetes 2009, 58, 1356–1364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Zherebitskaya, E.; Schapansky, J.; Akude, E.; Smith, D.R.; van der Ploeg, R.; Solovyova, N.; Verkhratsky, A.; Fernyhough, P. Sensory neurons derived from diabetic rats have diminished internal Ca2+ stores linked to impaired re-uptake by the endoplasmic reticulum. ASN Neuro. 2012, 4, e00072. [Google Scholar] [CrossRef] [PubMed]
  136. Sango, K.; Saito, H.; Takano, M.; Tokashiki, A.; Inoue, S.; Horie, H. Cultured adult animal neurons and schwann cells give us new insights into diabetic neuropathy. Curr. Diabetes Rev. 2006, 2, 169–183. [Google Scholar] [CrossRef]
  137. Jones, M.R.; Urits, I.; Wolf, J.; Corrigan, D.; Colburn, L.; Peterson, E.; Williamson, A.; Viswanath, O. Drug-induced peripheral neuropathy: A narrative review. Curr. Clin. Pharmacol. 2020, 15, 38–48. [Google Scholar] [CrossRef]
  138. Popescu, C. Severe acute axonal neuropathy induced by ciprofloxacin: A case report. Case Rep. Neurol. 2018, 10, 124–129. [Google Scholar] [CrossRef] [PubMed]
  139. Shankarappa, S.A.; Sagie, I.; Tsui, J.H.; Chiang, H.H.; Stefanescu, C.; Zurakowski, D.; Kohane, D.S. Duration and local toxicity of sciatic nerve blockade with coinjected site 1 sodium-channel blockers and quaternary lidocaine derivatives. Reg. Anesth. Pain Med. 2012, 37, 483–489. [Google Scholar] [CrossRef] [PubMed]
  140. Verlinde, M.; Hollmann, M.W.; Stevens, M.F.; Hermanns, H.; Werdehausen, R.; Lirk, P. Local anesthetic-induced neurotoxicity. Int. J. Mol. Sci. 2016, 17, 339. [Google Scholar] [CrossRef]
  141. Hewson, D.W.; Bedforth, N.M.; Hardman, J.G. Peripheral nerve injury arising in anaesthesia practice. Anaesthesia 2018, 73, 51–60. [Google Scholar] [CrossRef] [Green Version]
  142. Wen, X.; Li, Y.; Liu, X.; Sun, C.; Lin, J.; Zhang, W.; Wu, Y.; Wang, X. Roles of CaMKIIβ in the neurotoxicity induced by ropivacaine hydrochloride in dorsal root ganglion. Artif. Cells Nanomed. Biotechnol. 2019, 47, 2948–2956. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Staff, N.P.; Fehrenbacher, J.C.; Caillaud, M.; Damaj, M.I.; Segal, R.A.; Rieger, S. Pathogenesis of paclitaxel-induced peripheral neuropathy: A current review of in vitro and in vivo findings using rodent and human model systems. Exp. Neurol. 2020, 324, 113121. [Google Scholar] [CrossRef] [PubMed]
  144. Zhuo, M.; Gorgun, M.F.; Englander, E.W. Translesion SYnthesis DNA polymerase kappa is indispensable for DNA repair synthesis in cisplatin Exposed dorsal root ganglion neurons. Mol. Neurobiol. 2018, 55, 2506–2515. [Google Scholar] [CrossRef] [PubMed]
  145. Triarico, S.; Romano, A.; Attinà, G.; Capozza, M.A.; Maurizi, P.; Mastrangelo, S.; Ruggiero, A. Vincristine-induced peripheral neuropathy (VIPN) in pediatric tumors: Mechanisms, risk factors, strategies of prevention and treatment. Int. J. Mol. Sci. 2021, 22, 4112. [Google Scholar] [CrossRef]
  146. Baldewig, M.; Goldbaum, O.; Richter-Landsberg, C.; Weyland, A.; Bantel, C. Short-term incubation of gabapentin or pregabalin does not affect chemically induced injury in neuronal cell models in vitro. JPR 2018, 11, 1181–1190. [Google Scholar] [CrossRef] [Green Version]
  147. Guo, L.; Hamre, J.; Eldridge, S.; Behrsing, H.P.; Cutuli, F.M.; Mussio, J.; Davis, M. Multiparametric image analysis of rat dorsal root ganglion cultures to evaluate peripheral neuropathy- inducing chemotherapeutics. Toxicol. Sci. 2017, 156, 275–288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Eldridge, S.; Scuteri, A.; Jones, E.M.C.; Cavaletti, G.; Guo, L.; Glaze, E. Considerations for a reliable in vitro model of chemotherapy-induced peripheral neuropathy. Toxics 2021, 9, 300. [Google Scholar] [CrossRef]
  149. Eldridge, S.; Guo, L.; Hamre, J. A comparative review of chemotherapy-induced peripheral neuropathy in in vivo and in vitro models. Toxicol. Pathol. 2020, 48, 190–201. [Google Scholar] [CrossRef]
  150. Silva, A.; Wang, Q.; Wang, M.; Ravula, S.K.; Glass, J.D. Evidence for direct axonal toxicity in vincristine neuropathy. J. Peripher. Nerv. Syst. 2006, 11, 211–216. [Google Scholar] [CrossRef]
  151. Yang, I.H.; Siddique, R.; Hosmane, S.; Thakor, N.; Höke, A. Compartmentalized microfluidic culture platform to study mechanism of paclitaxel-induced axonal degeneration. Exp. Neurol. 2009, 218, 124–128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Geisler, S.; Doan, R.A.; Cheng, G.C.; Cetinkaya-Fisgin, A.; Huang, S.X.; Höke, A.; Milbrandt, J.; DiAntonio, A. Vincristine and bortezomib use distinct upstream mechanisms to activate a common SARM1-dependent axon degeneration program. JCI Insight 2019, 4, e129920. [Google Scholar] [CrossRef]
  153. Imai, S.; Koyanagi, M.; Azimi, Z.; Nakazato, Y.; Matsumoto, M.; Ogihara, T.; Yonezawa, A.; Omura, T.; Nakagawa, S.; Wakatsuki, S.; et al. Taxanes and platinum derivatives impair schwann cells via distinct mechanisms. Sci. Rep. 2017, 7, 5947. [Google Scholar] [CrossRef] [Green Version]
  154. Werdehausen, R.; Braun, S.; Essmann, F.; Schulze-Osthoff, K.; Walczak, H.; Lipfert, P.; Stevens, M.F. Lidocaine induces apoptosis via the mitochondrial pathway independently of death receptor signaling. Anesthesiology 2007, 107, 136–143. [Google Scholar] [CrossRef]
  155. El-Boghdadly, K.; Pawa, A.; Chin, K.J. Local anesthetic systemic toxicity: Current perspectives. Local Reg. Anesth. 2018, 11, 35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Hermanns, H.; Hollmann, M.W.; Stevens, M.F.; Lirk, P.; Brandenburger, T.; Piegeler, T.; Werdehausen, R. Molecular mechanisms of action of systemic lidocaine in acute and chronic pain: A narrative review. Br. J. Anaesth. 2019, 123, 335–349. [Google Scholar] [CrossRef] [PubMed]
  157. Kim, E.-J.; Kim, H.Y.; Ahn, J.-H. Neurotoxicity of local anesthetics in dentistry. J. Dent. Anesth. Pain Med. 2020, 20, 55. [Google Scholar] [CrossRef]
  158. Zhang, H.; Wang, K. Downregulation of microRNA-33-5p protected bupivacaine-induced apoptosis in murine dorsal root ganglion neurons through GDNF. Neurotox. Res. 2019, 35, 860–866. [Google Scholar] [CrossRef]
  159. Guo, Z.; Liu, Y.; Cheng, M. Resveratrol Protects bupivacaine-induced neuro-apoptosis in dorsal root ganglion neurons via activation on tropomyosin receptor kinase A. Biomed. Pharmacother. 2018, 103, 1545–1551. [Google Scholar] [CrossRef]
  160. Brull, R.; Hadzic, A.; Reina, M.A.; Barrington, M.J. Pathophysiology and etiology of nerve injury following peripheral nerve blockade. Reg. Anesth. Pain Med. 2015, 40, 479–490. [Google Scholar] [CrossRef] [PubMed]
  161. Zhang, K.; Yang, S.; Luo, C. TNF-alpha and TNF-R1 regulate bupivacaine-induced apoptosis in spinal cord dorsal root ganglion neuron. Eur. J. Pharmacol. 2018, 833, 63–68. [Google Scholar] [CrossRef] [PubMed]
  162. Smith, J.E.; Garner, J. Pathophysiology of primary blast injury. J. R. Army Med. Corps 2019, 165, 57–62. [Google Scholar] [CrossRef] [PubMed]
  163. Antonine Jerusalem, M.D. Continuum modelling of neuronal cell under blast loading. Acta Biomater. 2012, 8, 3360–3371. [Google Scholar] [CrossRef] [Green Version]
  164. Cernak, I. Understanding blast-induced neurotrauma: How far have we come? Concussion 2017, 2, CNC42. [Google Scholar] [CrossRef] [Green Version]
  165. Hicks, R.R.; Fertig, S.J.; Desrocher, R.E.; Koroshetz, W.J.; Pancrazio, J.J. Neurological effects of blast injury. J. Trauma Inj. Infect. Crit. Care 2010, 68, 1257–1263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Suneson, A.; Hansson, H.A.; Lycke, E.; Seeman, T. Pressure wave injuries to rat dorsal root ganglion cells in culture caused by high-energy missiles. J. Trauma Inj. Infect. Crit. Care 1989, 29, 10–18. [Google Scholar] [CrossRef] [PubMed]
  167. Zander, N.E.; Piehler, T.; Boggs, M.E.; Banton, R.; Benjamin, R. In vitro studies of primary explosive blast loading on neurons. J. Neurosci. Res. 2015, 93, 1353–1363. [Google Scholar] [CrossRef]
  168. Chen, P.; Piao, X.; Bonaldo, P. Role of macrophages in wallerian degeneration and axonal regeneration after peripheral nerve injury. Acta Neuropathol. 2015, 130, 605–618. [Google Scholar] [CrossRef]
  169. Madhusudanan, P.; Reade, S.; Shankarappa, S.A. Neuroglia as targets for drug delivery systems: A review. Nanomed. Nanotechnol. Biol. Med. 2017, 13, 667–679. [Google Scholar] [CrossRef]
  170. Sanders, V.M.; Jones, K.J. Role of immunity in recovery from a peripheral nerve injury. J. Neuroimmune Pharmacol. 2006, 1, 11–19. [Google Scholar] [CrossRef]
Figure 1. Schematic overview depicting various in vitro peripheral nerve injury models and peripheral cellular responses to injuries.
Figure 1. Schematic overview depicting various in vitro peripheral nerve injury models and peripheral cellular responses to injuries.
Ijms 23 00816 g001
Figure 2. Aligned scratches on PDL-coated culture dish using rake pins for axonal guidance. Representative image of micron-sized scratches applied on poly-D-lysine-coated cell culture plates. Each scratch was approximately 14 μm in width and spaced approximately 150 μm apart. Inset shows the pin rake made by cementing together 20 individual insect pins, with each pin placed approximately 200 μm apart. Scale bar = 25 μm.
Figure 2. Aligned scratches on PDL-coated culture dish using rake pins for axonal guidance. Representative image of micron-sized scratches applied on poly-D-lysine-coated cell culture plates. Each scratch was approximately 14 μm in width and spaced approximately 150 μm apart. Inset shows the pin rake made by cementing together 20 individual insect pins, with each pin placed approximately 200 μm apart. Scale bar = 25 μm.
Ijms 23 00816 g002
Figure 3. Laser-induced transection injury of an axon. Image showing a laser beam focused on neurite extension from a differentiated RGC-5 cell. Laser pulse was set to 7.5 nJ at 76 MHz, with exposure of 5 s (reproduced with permission from Santiago Costantino, Department of Ophthalmology, University of Montreal, Canada, PLoS ONE; 2011, 6(11), e26832.
Figure 3. Laser-induced transection injury of an axon. Image showing a laser beam focused on neurite extension from a differentiated RGC-5 cell. Laser pulse was set to 7.5 nJ at 76 MHz, with exposure of 5 s (reproduced with permission from Santiago Costantino, Department of Ophthalmology, University of Montreal, Canada, PLoS ONE; 2011, 6(11), e26832.
Ijms 23 00816 g003
Figure 4. Axonal stretching from DRG explants. Image showing neurons from DRG explants cultured on a stationary substrate (SS) and a mobile towing substrate (TS). Neurons cultured on both SS and TS when exposed to controlled stretching force yielded axons with bi-directional polarity (A), while neurons cultured exclusively on TS yielded axons with unidirectional polarity (B). Scale bar = 2 mm (reproduced with permission from Bryan J. Pfister, Department of Biomedical Engineering, New Jersey Institute of Technology, Newark, NJ, USA. Journal of Neurotrauma, 28 (11), 2011, 2389–2403, Mary Ann Liebert, Inc.).
Figure 4. Axonal stretching from DRG explants. Image showing neurons from DRG explants cultured on a stationary substrate (SS) and a mobile towing substrate (TS). Neurons cultured on both SS and TS when exposed to controlled stretching force yielded axons with bi-directional polarity (A), while neurons cultured exclusively on TS yielded axons with unidirectional polarity (B). Scale bar = 2 mm (reproduced with permission from Bryan J. Pfister, Department of Biomedical Engineering, New Jersey Institute of Technology, Newark, NJ, USA. Journal of Neurotrauma, 28 (11), 2011, 2389–2403, Mary Ann Liebert, Inc.).
Ijms 23 00816 g004
Figure 5. Stretching an individual axon. Image showing an unstretched (A) and a stretched (B) axon. The lateral stretching force was applied using an optical fiber cantilever. Scale bar: 20 μm (reproduced with permission from Pramod Pullarkat, Soft Condensed Matter Group, Raman Research Institute, Bengaluru, India, eLife; 2020, 9, 1–22).
Figure 5. Stretching an individual axon. Image showing an unstretched (A) and a stretched (B) axon. The lateral stretching force was applied using an optical fiber cantilever. Scale bar: 20 μm (reproduced with permission from Pramod Pullarkat, Soft Condensed Matter Group, Raman Research Institute, Bengaluru, India, eLife; 2020, 9, 1–22).
Ijms 23 00816 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Varier, P.; Raju, G.; Madhusudanan, P.; Jerard, C.; Shankarappa, S.A. A Brief Review of In Vitro Models for Injury and Regeneration in the Peripheral Nervous System. Int. J. Mol. Sci. 2022, 23, 816. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23020816

AMA Style

Varier P, Raju G, Madhusudanan P, Jerard C, Shankarappa SA. A Brief Review of In Vitro Models for Injury and Regeneration in the Peripheral Nervous System. International Journal of Molecular Sciences. 2022; 23(2):816. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23020816

Chicago/Turabian Style

Varier, Parvathi, Gayathri Raju, Pallavi Madhusudanan, Chinnu Jerard, and Sahadev A. Shankarappa. 2022. "A Brief Review of In Vitro Models for Injury and Regeneration in the Peripheral Nervous System" International Journal of Molecular Sciences 23, no. 2: 816. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23020816

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop