Next Article in Journal
Interaction between Gut Microbiota and Dendritic Cells in Colorectal Cancer
Next Article in Special Issue
Dynamic Pupillary Response in Multiple Sclerosis Patients with and without Optic Neuritis
Previous Article in Journal
KDF1 Promoted Proliferation, Migration and Invasion of Lung Adenocarcinoma Cells through Activating STAT3 and AKT Pathway
Previous Article in Special Issue
Understanding the Episodic Memory and Executive Functioning Axis Impairment in MCI Patients: A Multicenter Study in Comparison with CSF Biomarkers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Review of Piezoelectrical Materials Potentially Useful for Peripheral Nerve Repair

1
Departamento de Anatomia, NOVA Medical School|Faculdade de Ciências Médicas (NMS|FCM), Universidade NOVA de Lisboa, 1169-056 Lisboa, Portugal
2
Plastic and Reconstructive Surgery Department and Burn Unit, Centro Hospitalar Universitário de Lisboa Central, Rua José António Serrano, 1169-045 Lisbon, Portugal
3
Centro de Ciências e Tecnologias Nucleares (C2TN), Instituto Superior Técnico (IST), Universidade de Lisboa, 2695-066 Bobadela, Portugal
4
Departamento de Engenharia e Ciências Nucleares (DECN), Instituto Superior Técnico (IST), Universidade de Lisboa, 2695-066 Bobadela, Portugal
5
Centro de Química Estrutural (CQE), Institute of Molecular Sciences (IMS), Instituto Superior Técnico (IST), Universidade de Lisboa, 2695-066 Bobadela, Portugal
6
Centro de Ecologia, Evolução e Alterações Ambientais (cE3c) & CHANGE—Global Change and Sustainability Institute, Departamento de Biologia Animal, Faculdade de Ciências, Universidade de Lisboa (FCUL), 1749-016 Lisboa, Portugal
7
Gynaecology and Obstetrics Department, Maternidade Alfredo da Costa, Centro Hospitalar Universitário de Lisboa Central, R. Viriato 1, 2890-495 Lisboa, Portugal
8
Anatomy Institute and Orthopedics Department, Faculty of Medicine, University of Coimbra, 3004-504 Coimbra, Portugal
9
Spine Unit, Orthopedics Department, Coimbra University Hospital, 3000-602 Coimbra, Portugal
10
Departamento de Biologia Animal, Faculdade de Ciências, Universidade de Lisboa (FCUL), 1749-016 Lisboa, Portugal
*
Author to whom correspondence should be addressed.
Biomedicines 2023, 11(12), 3195; https://doi.org/10.3390/biomedicines11123195
Submission received: 1 October 2023 / Revised: 21 November 2023 / Accepted: 29 November 2023 / Published: 1 December 2023
(This article belongs to the Special Issue Neurodegenerative Diseases: Recent Advances and Future Perspectives)

Abstract

:
It has increasingly been recognized that electrical currents play a pivotal role in cell migration and tissue repair, in a process named “galvanotaxis”. In this review, we summarize the current evidence supporting the potential benefits of electric stimulation (ES) in the physiology of peripheral nerve repair (PNR). Moreover, we discuss the potential of piezoelectric materials in this context. The use of these materials has deserved great attention, as the movement of the body or of the external environment can be used to power internally the electrical properties of devices used for providing ES or acting as sensory receptors in artificial skin (e-skin). The fact that organic materials sustain spontaneous degradation inside the body means their piezoelectric effect is limited in duration. In the case of PNR, this is not necessarily problematic, as ES is only required during the regeneration period. Arguably, piezoelectric materials have the potential to revolutionize PNR with new biomedical devices that range from scaffolds and nerve-guiding conduits to sensory or efferent components of e-skin. However, much remains to be learned regarding piezoelectric materials, their use in manufacturing of biomedical devices, and their sterilization process, to fine-tune their safe, effective, and predictable in vivo application.

1. Introduction

The piezoelectrical effect was discovered in 1880 by the French scientists Jacques and Pierre Curie. They observed that piezoelectrical materials can transfer electrons when pressed and/or twisted, and receive back these electrons when distended or relaxed, allowing the generation of electric dipoles. The observation that pressure on certain materials (single-crystal quartz in the original description) generated an electrical charge led to the term “piezoelectricity” being coined (“piezo” means “pressure” in Greek). Piezoelectricity results from the conversion of mechanical energy (mechanical strain and vibration) into electric polarization without the need of applying an external voltage. Since its original description, its use has become extensively widespread, and there now exist numerous applications of this effect in industrial, military, domestic and health care settings [1,2,3].
The qualification of the piezoelectric effect can be made through the Piezoelectric Coefficient, which can be defined as the charge (expressed in Coulombs) developed on the surface of the piezoelectric material per unit force applied on it (expressed in Newtons). Hence, in the SI system, the unit becomes coulomb/Newton. Notwithstanding, since the charge developed per unit of force is small, the Piezoelectric Coefficient is more conveniently expressed as pC/N [4].
Interestingly, more recently it was noted that piezoelectricity was ubiquitous in nature and particularly in living beings. It can be observed in the various degrees of nature’s hierarchical organization, from the amino acid and protein levels to DNA molecules, viruses, tissues, organs, skeletons, and even jungles and seashores [5,6,7,8,9].
Moreover, it has increasingly been realized that electrical currents may play a pivotal key role in cell migration and tissue repair. Hence, the heightened interest in piezoelectrical materials, particularly in the realm of peripheral nerve repair (PNR), where the contemporary reconstructive strategies present frequently dismaying results, leaving those affected with permanent motor, sensory and/or autonomic disability. Furthermore, these patients are often stricken with neuropathic pain whose treatment is often difficult and incomplete [10,11,12,13]. Therefore, peripheral nerve injuries (PNI) exert a significant psychosocial and economic burden on both individuals and society [13,14].
This is all the more important taking into consideration that PNI are relatively common from birth to old age, occurring in multiple contexts, namely compressive neuropathies, multiple types of trauma, as a result of tumor treatment, in anesthetic procedures, infections or degenerative diseases [14,15,16,17,18,19]. Up to one in every 1000 children are born with a significant PNI (brachial plexus palsy) [20]. It is estimated that in Sweden alone, each year there are 13.9 new cases for each 100,000 people of serious PNI mandating hospitalization [14]. In the USA, the mean annual incidence of PNI is 16.9 per 100,000 for the upper extremity, and 13.3 cases per million for the lower extremity [10]. In this country, more than 200,000 trauma-related nerve injuries occur each year [21]. Worldwide, the incidence of PNI in the head, neck and trunk regions is also significant, although difficult to quantify [22].
In this review, we will summarize the current evidence supporting the potential benefits of electric stimulation (ES) in the physiology of PNR. Subsequently, we will discuss the materials with piezoelectrical properties available for producing devices with potential use for PNR. Next, we will review the devices already described using piezoelectrical properties for PNR. Finally, we will discuss future perspectives concerning the use of piezoelectrical materials in this context.

2. Role of Electrical Stimulation in the Physiology of Peripheral Nerve Repair

It has been known for more than 150 years, since the seminal works of Luigi Galvani and later on of those of Emil Du-Bois Reymond, that after tissue injury there is a local disturbance in electric charges, generating endogenous electrical fields and electric current [23,24,25]. These, in turn, create electric dipoles that guide and promote the migration of numerous cell types in process named “galvanotaxis” or “electrotaxis” [24,26,27,28]. Numerous animal studies have shown that galvanotaxis is initiated immediately after injury, helping coordinate all the processes (hemostasis, inflammation, proliferation and remodeling) that lead to definitive tissue repair [24,26,27,28]. Under the influence of electric fields, peripheral nervous system neurons extend protrusions and migrate towards the cathodic pole [26,29]. This type of behavior has also been observed in human fibroblasts, lymphocytes, macrophages [30], and endothelial cells, all of which are known to be important in peripheral nerve repair [30,31,32,33]. Piezoelectric materials can generate electrical charges in response to mechanical strain, thus stimulating axonal regeneration by galvanotaxis following nerve injury [34,35]. To increase the amount of electricity produced by piezoelectric materials, some authors have applied external ultrasound sources to internally placed devices that are used in PNR [36,37,38].
In 1952, Hoffman noted enhanced peripheral nerve regeneration after applying ES stimulation in nerve roots [13,39]. Subsequent studies on rabbit and rat hindlimb models confirmed the regenerative-inducing potential of ES in the peripheral nervous system [40,41,42,43].
The reason why the application of ES either intra-operatively or post-operatively has not yet become broadly accepted is probably because its mechanisms of action have remained largely elusive until recently. In fact, only in the past years have several technological advances allowed accurate electrophysiological measurements close to the injury site and contributed to the unravelling of the underlying physiological mechanisms behind enhanced PNS recovery after ES [13,23,26,44,45].
However, even today multiple questions remain unanswered. The most primordial and pressing question probably pertains to the mechanisms that allow cells to sense electrical charges [23]. Notwithstanding, several studies have suggested that asymmetrically distributed cell receptors, namely, integrins and receptors of acetylcholine, epidermal growth factor and of concanavalin A, probably play a role in the electrotactic response [23,34,35,46,47,48,49,50,51,52,53].
Experimental data suggest that ES is transduced by the second messenger molecules cyclic AMP, Rho-associated protein kinase and phosphoinositide-3 kinase [29]. Additionally, ES causes up-regulation of brain-derived neurotrophic factor, T alpha-1 tubulin, growth-associated protein 43 (GAP-43), as well as other regeneration-associated genes, resulting in axon regeneration [35,45,54,55,56]. Globally, all these events lead to increased neuronal cell adhesion, proliferation, migration, and protein synthesis, particularly of neuronal cytoskeletal proteins, hastening the outgrowth of PNS axons across the injury site [11,12,23,35,43,45,54,57,58]. Additionally, ES promotes remyelination of elongating axons by Schwan cells [45,59]. Furthermore, in a mouse model, ES has been shown to induce differentiation of neural stem cells and progenitor cells into neurons and glial cells [26,60].
Clinically, ES has been studied sparsely for the past decades. Most studies are related to its percutaneous application for prevention of muscular atrophy after PNS injury [61]. There have been four randomized clinical trials on the use of ES in the clinical setting, all presenting positive results. Two of them report postoperative ES after carpal tunnel and cubital tunnel surgical release [62,63]. Another study describes the prevention of accessory nerve dysfunction after oncologic neck dissection using intraoperative ES [64]. Lastly and most revealingly, there is paper on the brief post-surgical low frequency ES of surgically repaired digital nerves which had been accidentally sectioned. This work showed accelerated axon outgrowth across the repair site and hastened target reinnervation [54].
All these data spurred the recent enthusiasm over the use of electrical currents to treat different types of pathologies, including PNS lesions. This, in turn, led to the term “electroceuticals” being coined [26,34,35]. Lack of sound data regarding the treatment of PNI with ES, namely concerning the best method of delivery of ES, its frequency, duration or intensity, the occasional discomfort associated with its use in awake patients, the need for multiple interventions, and its feasibility in critical nerve gap injury have generally been enumerated as reasons for the lack of general acceptance of this method [12,23,26,32,34,35].
In a significant number of PNI cases, there is a gap between the nerve stumps that precludes their surgical suture. In these cases, it is necessary to apply a conduit to “bridge” the nerve defect. In fact, after a nerve section there is a latency period of up to 30 days in which the proximal axons do not elongate in the direction of the distal nerve stump [32]. If the nerve gap was not protected and bridged, the surrounding connective tissue would proliferate in this period and physically block most of the elongating axons from the proximal stump from reaching the distal nerve stump. Traditionally, autologous dispensable nerves or veins are used to bridge nerve defects [65,66,67]. More recently, allogenic nerve grafts and artificial nerve-guiding conduits (NGC) were introduced, and are in widespread use [13,32]. However, none of these options is perfect. Autologous alternatives entail non-negligible donor site morbidity. Allogenic nerve grafts have been associated with inflammatory reactions akin to rejection responses. Generally, both allogenic nerve grafts and NGC have been associated with worse functional results than autologous nerve grafts or flaps, particularly for longer nerve defects [13,16,32,35,65,66,67,68].
Hence, great effort has been put into developing better NGCs, ideally with ES properties. In this context, the use of piezoelectric materials has deserved great attention, as the movement of the body could be used to power internally the electrical properties of the device, avoiding toxic batteries which eventually need to be removed or exchanged surgically [69,70,71,72,73,74].
The authors propose that there is enough evidence to believe that piezoelectric materials may play a significant role in the treatment of PNI. In this paper, the authors will try to provide a critical appraisal of the literature on this subject.

3. Piezoelectrical Materials

There is a plethora of both inorganic and organic strongly piezoelectric materials available to construct NGCs and other PNI repair devices [5,74]. When choosing these materials, it is fundamental to have a sound grasp of not only their piezoelectric properties, but also their safety profile, their biocompatibility, their biostability, and their degradation products inside the body. Moreover, it is also important to understand the techniques available to shape them into the desired geometrical configurations. Finally, their ductility, resistance and softness should also be controlled, in order for the devices to be fixed with sutures and tolerated inside the body [35,74] (Table 1).

4. Inorganic

Aluminum nitride (AlN), Barium titanate (BaTiO3), Lead zirconate titanate (PZT-5H) are examples of biocompatible ceramic materials with a high piezoelectric response. However, they are rigid, brittle, and contain non-degradable and toxic compounds that limit their potential for constructing implantable devices [5,99].
One of the most commonly used inorganic polymers in PNR is polyvinylidene fluoride (PVDF). This compound is flexible, has excellent piezoelectric properties and is biocompatible, allowing for direct contact with biological tissues. However, it is not degradable, requiring a removal surgery with all the inconvenience and the risks it entails [5,74].
Graphene is another compound that, despite having piezoelectric properties, a high surface area and high electrical conductivity, can be hazardous when inserted into the body, as it breaks up and its fragments can accumulate inside various organs, potentially causing severe cellular damage and disease [5,74,100].
Metals are considered too rigid to be used alone in ES devices. However, metal nanoparticles, namely of gold, silver, and copper, can be used to increase the mechanical strength and electrical conductivity of composite materials. Some metals can progressively dissolve, such as magnesium, zinc, tungsten, iron and molybdenum, allowing the construction of “transient electronics” [101]. However, due to the consumption of oxygen and release of byproducts in the corrosion of these metals, which can lead to adjacent tissue necrosis, biosafety studies are warranted before clinical trials are implemented [5,35,74,102,103,104,105].

5. Organic

Organic piezoelectric biomaterials can be of different classes, such as natural occurring amino acids (e.g., glycine, cysteine, alanine, threonine, diphenylalanine), proteins (e.g., collagen, silk), and polysaccharides (e.g., cellulose, chitin, chitosan, alginate), or synthetic polymeric compounds, such as poly-lactic acid (PLA), Glycine-Polyvinyl alcohol (PVA), Polycaprolactone (PCL), Polyamide (PA), and Polypyrrole (PPy) [106,107,108]. Since synthetic polymers have greater mechanical qualities than natural polymers and can be readily synthesized into 3D structures, they are frequently employed to fabricate NGCs for PNR [109,110].
Although with a lower piezoelectric effect compared to many inorganic materials, these organic compounds present a much more favorable biocompatibility, biosafety, and biodegradability profile. In fact, these organic compounds are readily recognized and naturally degraded by host cells and/or microbiome enzymes, allowing for recipient cell invasion and progressive replacement of the device with endogenous tissues. Hence, living tissues can easily tolerate these compounds without triggering unfavorable immunological reactions. PA and PCL, for example, are frequently used when strength, flexibility and durability are required [111]. Arguably, these properties make these materials the most obvious candidates for the construction of implantable ES devices [69,70,71,110,112].
β-Glycine has received great attention, due to its high piezoelectric constant in its crystalline form. Unfortunately, glycine salts are readily soluble in body fluids and are difficult to handle, as they are hard and brittle. To circumvent these limitations, β-Glycine crystals have been associated with Polycaprolactone to form a soft and resistant material with a significant piezoelectric effect. This composite material has already been used with success as devices placed inside rodents’ brains, and to produce NGC [5,74,113].
Collagen is the most common extracellular protein in animals. It makes for a good material for PNR as it promotes cell adhesion and development and presents excellent biocompatibility, hydrophilicity, and low antigenicity. Moreover, it has significant piezoelectric properties [78,114,115].
Silk fibroin is extracted from silkworm silk and has exceptional mechanical strength, flexibility, and biocompatibility. These characteristics enable the production of silk fibroin in a variety of shapes, including films, fibers, and scaffolds, all of which can be adjusted to closely resemble the mechanical characteristics of peripheral nerves. Silk fibroin is a perfect substance for nerve tissue engineering, since it has also been demonstrated to support cell adhesion, proliferation, and differentiation [79].
Cellulose is a biocompatible and biodegradable polymer, with piezoelectric properties, that has been increasingly used in PNR. Cellulose has exceptional mechanical qualities, such as high strength and stiffness, which are essential for supporting the structure of the nerve during regeneration. This is crucial for PNR, since the material needs to be able to endure mechanical stresses and maintain its integrity until axonal elongation is concluded. Furthermore, cellulose has good biocompatibility. This quality is necessary to support cell adhesion, differentiation, and proliferation—all processes that are necessary for effective neuron regeneration. Moreover, cellulose is a sustainable and renewable substance, which makes it an attractive choice from an environmental standpoint. Its abundance in nature and ability to be derived from various sources, such as plants or bacteria, further contribute to its appeal as a piezoelectric material for peripheral nerve repair [80].
Chitin can be obtained from the exoskeletons of insects, arthropods, and crustacean shells. After cellulose, chitin is the most common natural polysaccharide. Chitosan can be found in some fungi or be derived from chitin through the partial deacetylation of the latter [79,116].
Chitin and chitosan possess several advantageous features, namely biocompatibility, biodegradability, amenability to create various geometrical forms (porous scaffolds, hydrogels, fibers, sponges, films, etc.), chemical and enzymatic modifiability, antimicrobial characteristics, potential for controlled release of cytokines, antibiotics and extracellular matrix constituents, the ability to promote cell adherence and viability [82]. The characteristics mentioned above led to multiple studies on the potential use of chitosan and chitin in reconstruction of peripheral nerves [117].
Another interesting natural polysaccharide is alginate. This compound is present in the cell walls of brown algae. It is characterized by its hydrophilicity, retaining large volumes of water and forming gels in the process. Alginate is also characterized by its biocompatibility and low toxicity. Its main disadvantages are uncontrollable and relatively fast degradation, inadequate mechanical strength, and inadequate cell signaling [118]. Furthermore, alginate piezoelectric properties are minimal [83].
Polyvinylidene fluoride (PVDF) is a frequently utilized synthetic organic piezoelectric material for PNR, due to its piezoelectric qualities. It possesses adequate mechanical flexibility and strength, electrical conductivity, and biocompatibility for the production of NGC. However, this material is naturally strongly hydrophobic, which limits its usage in biomedical device fabrication, unless chemically modified or associated with other more hydrophilic materials [119,120].
Alternatively, poly(lactic-co-glycolic acid) (PLGA), is a biodegradable and biocompatible polymer, which possesses piezoelectric properties when aligned in a certain direction. Its biodegradability has been exploited for the controlled release of drugs or growth factors, further enhancing the regenerative process [121].
Additional synthetic organic piezoelectric materials used in the realm of PNR are polyurethane (PU) and poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) (PHBHHx). PU exhibits good mechanical properties, biocompatibility, and electrical conductivity. PHBHHx, on the other hand, is a biodegradable and biocompatible polymer that can be electrospun into nanofiber scaffolds for nerve regeneration [86,87,122].
Poly-γ-benzyl-L-glutamate (PBLG) is another synthetic peptide. It is a resilient and flexible material with good mechanical qualities, simplicity of processing, and flexibility, which make it a good fit for applications involving PNR. In animal models, PBLG showed enhanced axonal outgrowth and nerve regeneration. Furthermore, scaffolds based on PBLG have been created, offering structural support to direct nerve cells and promote their proliferation. By imitating the natural extracellular matrix, these scaffolds can be made to encourage cell adhesion, proliferation, and differentiation [5].
The fact that most organic materials sustain spontaneous degradation inside the body means their piezoelectric effect is also limited in duration. In the case of PNR, this is not necessarily problematic, as ES is only required during the regeneration period of the PNS [69,70,71,72,74,102].
However, in general, it can be argued that hardly any given material possesses ideal biological, piezoelectric and structural properties. For instance, naturally occurring materials with piezoelectric qualities present better biocompatibility and bioactivity. Conversely, synthetic piezoelectric materials such as PVDF and lead zirconate titanate (PZT) can be engineered to have specific mechanical and piezoelectric features. However, synthetic materials tend to be long-lasting, often behaving as a foreign body that may entrap the growing nerve and/or increase the risk of infection. These in turn may mandate a second surgery for scaffold extraction, increasing cost and potential morbidity [70]. Finally, graphene and other highly piezoelectric materials have been added to other more readily rebsorbable materials to enhance the piezoelectric properties of the composite materials [123].
Thus, combining different natural compounds, natural and synthetic materials, or different synthetic compounds in composite materials may allow the fine-tuning of the composition of devices for optimal PNR. For example, a recent study showed improved neurite outgrowth and nerve regeneration with a composite scaffold NGC made of PVDF and collagen. While the natural collagen aided in cell attachment and tissue integration, the synthetic PVDF supplied the required mechanical support and enhanced piezoelectric qualities. Similar studies have been performed with a myriad of natural and/or synthetic compounds [70,85,124,125,126].

6. Processing Piezoelectrical Materials

The piezoelectric effect of materials results from the regular alignment of molecular dipoles [127,128]. In some cases, it may be useful to increment this effect, particularly when the processing of the material diminishes the original piezoelectric effect. There are several methods to obtain this increment, namely stretching of the material (drawing), thermal annealing (heat treatment), application of a high external electrical field (electrical poling) or maximizing the macroscopic alignment of fibers in the case of materials obtained through electrospinning processes [5,74].
Stretching materials, especially at high temperatures, multiple times, promotes the alignment of dipoles. Heat treatment increases the crystalline content of amorphous materials, increasing their piezoelectric properties. Electrical poling consists of applying a high-voltage electrical field (commonly 1 to 10 kV) to ferroelectric materials (i.e., materials that have spontaneous electric polarization that can be reversed by the application of an external electric field), in order to align dipoles, which, in turn, will increase their piezoelectric properties and ensure the desired polarization [5,68,129].
Some authors have augmented the piezoelectric effect of devices by externally applying ultrasound to drive this effect [130]. Notwithstanding, detractors of this process argue that the frequent need for acoustic streaming may have deleterious effects in tissues, namely through cavitation and local heat generation [35,36,131].

7. Biomedical Devices

Several biomedical devices with piezoelectric properties have been applied experimentally for PNR [53].
Piezoelectric materials can be used to produce tridimensional scaffolds amenable to the implantation of cells important in the process of PNR, such as elongating axons, Schwan cells or mesenchymal cells. Bio-printed devices may even include the latter two cell types in their composition. Additionally, substances required for chemically stimulating cell growth and differentiation may be added. Finally, some of these chemotactic substances may also have piezoelectric properties, such as chitosan [5,68,116,132,133,134,135,136]. Organic bioresorbable piezoelectric materials are good options for tissue engineering, due to their biocompatibility, minimal toxicity and galvanotaxis effect [5,132].
Nerve-guiding conduits (NGC) are the most reported devices. They are used to bridge nerve gaps, which occur relatively frequently in clinical practice. Traditionally, these gaps have been reconstructed with resort to autologous nerve grafts, which entails variable donor site morbidity and limited supply [137].
Their potential association with specific cell therapy, growth factors, gene therapy alone or in combination, has been placing NGC as a promising alternative to nerve autografts [35,71,138,139,140]. NGC can be produced by various methods, namely 3D printing, mold casting, electrospinning, and roll-up sheeting [35,112]. Three-dimensional printing of NGC allows design freedom, as well as the possibility to replicate complex nerve anatomy in monolithic devices without any assembly requirements [112,141].
A self-powered patch composed of a flexible piezoelectric generator applied over a wound bed has been shown to promote skin nerve regeneration and sensation [142].
Electronic skin, also known as e-skin, is a broad term used to refer to artificial skin that emulates human skin, not only for covering and protective purposes, but also for providing haptic, thermal and humidity sensations [5,143,144]. It has a wide range of potential applications, namely robotics, prosthetics, virtual reality, human/machine interfacing, monitoring vital signs, detecting environmental pollutants, and human skin replacement [144,145].
E-skin typically consists of three main components: a flexible substrate, functional materials, and sensors. The flexible substrate provides the base for the electronic circuitry and ensures its mechanical flexibility. Various materials, such as polymers or nanomaterials, are used to achieve the desired properties of flexibility, stretchability, and durability. In several conceptions of e-skin, receptors are piezoelectric, although capacitive and resistive receptors have also been described [144,145].
Despite its multiple potential advantages, e-skin also presents several challenges. One limitation is the difficulty in achieving long-term stability and reliability of the electronic components embedded within the skin-like material. The mechanical and electrical properties of the e-skin need to be carefully optimized to ensure durability and performance over time. Additionally, the scalability and manufacturing processes of e-skin need to be further developed to enable mass production at a low cost. Moreover, the integration of power sources, such as batteries or energy harvesters, remains a challenge for e-skin devices [146].

8. Discussion

Even today, despite countless surgical and technological advancements, the clinical results after PNR remain unsatisfactory. This is surprising and certainly unrelated to the immense time scientists have devoted to research in this field. In fact, visionary surgeons like Paul of Aegina apparently were performing nerve sutures as far back as 600 AD [67,147,148,149,150,151].
Hence, the use of piezoelectric materials in the realm of PNR holds great promise as a breakthrough technology that may improve clinical results. In fact, by allowing the conversion of mechanical energy from normal movements of the body into electrical gradients, or the translation of the contact with normal external stimuli into electrical potentials, these materials can be used to produce scaffolds, NGC or e-skin. The ES they produce can be channeled to drive PNR. In fact, these materials have been shown in some experimental and clinical studies to promote the growth and alignment of nerve fibers that are regenerating. Moreover, many of these materials, particularly organic ones, are biocompatible, lowering the possibility of rejection or inflammation and ensuring compatibility with biological tissues [34,35].
Additionally, several of the devices produced with piezoelectric materials can be 3D-printed [112,152,153]. This manufacturing technique presents an enormous versatility and room for creativity. Being based on CAD (Computer Aided Design) files and not having the normal constraints of traditional manufacturing methods, it allows a wide range of geometric forms, including highly complex organic structures. Additionally, CAD files can be further refined using generative design. This process can be defined as a set of computational methods, including artificial intelligence algorithms and machine learning, designed to maximize structural performance requirements with the minimal amount of material, and a faster printing speed. Generative design can be used to perform topological optimization, reinforcing structures in the regions where greater forces are applied, without the need to create continuous objects or surfaces. In medical devices, this lattice structure reduces weight and facilitates native tissue invasion and integration. Simultaneously, this architecture promotes survival of cells by simple diffusion initially and by neo angiogenesis subsequently. Therefore, 3D printing has the potential to greatly improve the ergonomics and efficiency of medical devices used in PNR. Moreover, 3D-printing biomedical devices allows design freedom and the possibility to replicate complex nerve anatomy in monolithic devices without any assembly requirements. For example, printing NGCs before surgeries could diminish resort to autologous nerve grafts and their associated morbidity [137,154,155,156].
The inner structure of NGC can be printed in a compartmentalized fashion, providing additional physical clues to guide elongating axons [157,158].
Finally, 3D-printed piezoelectrical devices may even be associated with growth factors and/or cells, further boosting PNR [112,152,159,160,161,162,163].
However, there are still significant hurdles and difficulties related to the restoration of peripheral nerves using piezoelectric materials. Fabricating scaffolds or conduits with the mechanical strength, pliability, biocompatibility, nontoxicity, durability, and piezoelectric characteristics (ES dosage and polarity) required for implantation in the human body is one of the key issues. Advanced production processes and exact material composition and structure control are needed for this [34,35,85].
A potential caveat of using piezoelectric materials to produce NGC is the requirement of an adequate orientation of the electric polarization of the device. In fact, as mentioned above, neurons, fibroblasts, macrophages, and endothelial cells, paramount in PNR, have been shown experimentally to migrate towards the cathodic pole [26,29,30,31,32,33,164]. Electrical poling and stretching of materials can be used to align dipoles [5,68,129]. Characterization of piezoelectric features with Piezoresponse Force Microscopy and/or Atomic Force Microscopy measurements will help establish the efficacy of these techniques in specific devices [68,145,165,166].
Additionally, a largely overlooked aspect in this field is the effect of the required sterilizing processes required prior to in vivo implantation of biomedical devices. One of the most common ways to sterilize medical devices is using gamma radiation. However, its effects on the structure, physical and biological properties of commonly used 3D-printed devices is largely unknown [167]. There is some evidence that gamma radiation may cause weakening of mechanical properties, namely of tensile strength and elongation, which could limit the practical use of the 3D-printed devices. This knowledge is therefore of paramount importance to uphold the requirements for medical devices’ safety and usefulness [168,169,170].
Furthermore, it has been shown that PNR is dependent on local blood supply [66,67]. Hence, long and/or wide NGCs may provide inadequate blood supply to the elongating nerve. Therefore, increasing devices’ porosity or producing prefabricated vascularized NGCs have been proposed to try to mitigate these limitations [66,67,171,172]. In general, these modifications follow the trend of trying to replicate an optimized neuronal microenvironment, including structural, biochemical, electrical, vascular, and biological clues that support and promote PNR [173].
In the future, devices must be thoroughly studied not only in hindlimb models, but also in forelimb models where PNI are more common, and for which data are not yet available [65,67]. Finally, further studies are warranted to confirm or dismiss the promising experimental data and the scarce clinical data on the use of piezoelectric materials for PNR [35,53,65,67].

9. Conclusions

Arguably, piezoelectric materials have the potential to revolutionize the somewhat stalled field of peripheral nerve repair with new biomedical devices that range from scaffolds and NGC to sensory or efferent components of artificial skin (e-skin).
However, much remains to be learned regarding the piezoelectric materials, the manufacturing of the biomedical devices, and their sterilization process to fine-tune its safe, effective and predictable in vivo application [35].

Author Contributions

Conceptualization, D.C., M.H.C., L.M.F., J.P.L., G.R., R.L., D.L.M., L.G., J.B.L., D.P. and P.M.P.S.; Methodology, D.C., M.H.C., L.M.F. and P.M.P.S.; Data curation, G.R., R.L., D.L.M., L.G., J.B.L. and D.P.; Writing—original draft, D.C., M.H.C., L.M.F., G.R., R.L. and P.M.P.S.; Writing—review and editing, D.C., G.R., R.L., M.H.C., J.P.L., L.M.F., J.B.L. and P.M.P.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the International Atomic Energy Agency: Research Contract No. 24415 (CRP F23035). Authors acknowledge FCT—Fundação para a Ciência e a Tecnologia I.P. for the national funds in the scope of the project UIDB/04349/2020+UIDP/04349/2020, UIDB/00100/2020+LA/P/0056/2020 and UIDB/00329/2020.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the data described are present in the article.

Conflicts of Interest

The authors declare no conflict of interests.

References

  1. Javidi, H.; Ramazani Saadatabadi, A.; Sadrnezhaad, S.K.; Najmoddin, N. Conductive nerve conduit with piezoelectric properties for enhanced PC12 differentiation. Sci. Rep. 2023, 13, 12004. [Google Scholar] [CrossRef] [PubMed]
  2. Katzir, S. The Beginnings of Piezoelectricity: A Study in Mundane Physics; Springer: Berlin/Heidelberg, Germany, 2006; Volume 246. [Google Scholar]
  3. Wang, R.; Sui, J.; Wang, X. Natural piezoelectric biomaterials: A biocompatible and sustainable building block for biomedical devices. ACS Nano 2022, 16, 17708–17728. [Google Scholar] [CrossRef] [PubMed]
  4. Priya, S.; Song, H.-C.; Zhou, Y.; Varghese, R.; Chopra, A.; Kim, S.-G.; Kanno, I.; Wu, L.; Ha, D.S.; Ryu, J. A review on piezoelectric energy harvesting: Materials, methods, and circuits. Energy Harvest. Syst. 2017, 4, 3–39. [Google Scholar] [CrossRef]
  5. Ali, M.; Bathaei, M.J.; Istif, E.; Karimi, S.N.H.; Beker, L. Biodegradable Piezoelectric Polymers: Recent Advancements in Materials and Applications. Adv. Heal. Mater. 2023, 12, e2300318. [Google Scholar] [CrossRef] [PubMed]
  6. Ghosh, S.K.; Mandal, D. Efficient natural piezoelectric nanogenerator: Electricity generation from fish swim bladder. Nano Energy 2016, 28, 356–365. [Google Scholar] [CrossRef]
  7. Lee, B.Y.; Zhang, J.; Zueger, C.; Chung, W.J.; Yoo, S.Y.; Wang, E.; Meyer, J.; Ramesh, R.; Lee, S.W. Virus-based piezoelectric energy generation. Nat. Nanotechnol. 2012, 7, 351–356. [Google Scholar] [CrossRef]
  8. Ghosh, S.K.; Mandal, D. Bio-assembled, piezoelectric prawn shell made self-powered wearable sensor for non-invasive physiological signal monitoring. Appl. Phys. Lett. 2017, 110, 123701. [Google Scholar] [CrossRef]
  9. Fathizadeh, S.; Behnia, S. Control of a DNA based piezoelectric biosensor. J. Phys. Soc. Jpn. 2020, 89, 24004. [Google Scholar] [CrossRef]
  10. Maita, K.C.; Garcia, J.P.; Avila, F.R.; Torres-Guzman, R.A.; Ho, O.; Chini, C.C.S.; Chini, E.N.; Forte, A.J. Evaluation of the Aging Effect on Peripheral Nerve Regeneration: A Systematic Review. J. Surg. Res. 2023, 288, 329–340. [Google Scholar] [CrossRef]
  11. Panagopoulos, G.N.; Megaloikonomos, P.D.; Mavrogenis, A.F. The Present and Future for Peripheral Nerve Regeneration. Orthopedics 2017, 40, e141–e156. [Google Scholar] [CrossRef]
  12. Wang, M.L.; Rivlin, M.; Graham, J.G.; Beredjiklian, P.K. Peripheral nerve injury, scarring, and recovery. Connect. Tissue Res. 2019, 60, 3–9. [Google Scholar] [CrossRef] [PubMed]
  13. Kubiak, C.A.; Kung, T.A.; Brown, D.L.; Cederna, P.S.; Kemp, S.W.P. State-of-the-Art Techniques in Treating Peripheral Nerve Injury. Plast. Reconstr. Surg. 2018, 141, 702–710. [Google Scholar] [CrossRef]
  14. Asplund, M.; Nilsson, M.; Jacobsson, A.; von Holst, H. Incidence of Traumatic Peripheral Nerve Injuries and Amputations in Sweden between 1998 and 2006. Neuroepidemiology 2009, 32, 217–228. [Google Scholar] [CrossRef]
  15. Missios, S.; Bekelis, K.; Spinner, R.J. Traumatic peripheral nerve injuries in children: Epidemiology and socioeconomics. J. Neurosurg. Pediatr. 2014, 14, 688–694. [Google Scholar] [CrossRef] [PubMed]
  16. Kornfeld, T.; Vogt, P.M.; Radtke, C. Nerve grafting for peripheral nerve injuries with extended defect sizes. Wien. Med. Wochenschr. 2019, 169, 240–251. [Google Scholar] [CrossRef] [PubMed]
  17. Kouyoumdjian, J.A.; Graca, C.R.; Ferreira, V.F.M. Peripheral nerve injuries: A retrospective survey of 1124 cases. Neurol. India 2017, 65, 551–555. [Google Scholar] [CrossRef] [PubMed]
  18. Li, N.Y.; Onor, G.I.; Lemme, N.J.; Gil, J.A. Epidemiology of Peripheral Nerve Injuries in Sports, Exercise, and Recreation in the United States, 2009–2018. Phys. Sport. 2021, 49, 355–362. [Google Scholar] [CrossRef] [PubMed]
  19. Welch, M.B.; Brummett, C.M.; Welch, T.D.; Tremper, K.K.; Shanks, A.M.; Guglani, P.; Mashour, G.A. Perioperative peripheral nerve injuries: A retrospective study of 380,680 cases during a 10-year period at a single institution. J. Am. Soc. Anesthesiol. 2009, 111, 490–497. [Google Scholar] [CrossRef]
  20. Van der Looven, R.; Le Roy, L.; Tanghe, E.; Samijn, B.; Roets, E.; Pauwels, N.; Deschepper, E.; De Muynck, M.; Vingerhoets, G.; Van den Broeck, C. Risk factors for neonatal brachial plexus palsy: A systematic review and meta-analysis. Dev. Med. Child. Neurol. 2020, 62, 673–683. [Google Scholar] [CrossRef]
  21. Padovano, W.M.; Dengler, J.; Patterson, M.M.; Yee, A.; Snyder-Warwick, A.K.; Wood, M.D.; Moore, A.M.; Mackinnon, S.E. Incidence of nerve injury after extremity trauma in the United States. Hand 2022, 17, 615–623. [Google Scholar] [CrossRef]
  22. Stankovic, P.; Wittlinger, J.; Georgiew, R.; Dominas, N.; Hoch, S.; Wilhelm, T. Continuous intraoperative neuromonitoring (cIONM) in head and neck surgery-a review. HNO 2020, 68, 86–92. [Google Scholar] [CrossRef] [PubMed]
  23. Naixin, J.; Jinrui, Y.; Jie, L.; Zhang, J. Electric field: A key signal in wound healing. Chin. J. Plast. Reconstr. Surg. 2021, 3, 95–102. [Google Scholar]
  24. Nuccitelli, R.; Nuccitelli, P.; Ramlatchan, S.; Sanger, R.; Smith, P.J. Imaging the electric field associated with mouse and human skin wounds. Wound Repair. Regen. 2008, 16, 432–441. [Google Scholar] [CrossRef]
  25. Piccolino, M. Luigi Galvani’s path to animal electricity. Comptes Rendus Biol. 2006, 329, 303–318. [Google Scholar] [CrossRef] [PubMed]
  26. Tai, G.; Tai, M.; Zhao, M. Electrically stimulated cell migration and its contribution to wound healing. Burn. Trauma 2018, 6, 20. [Google Scholar] [CrossRef] [PubMed]
  27. Tai, G.; Reid, B.; Cao, L.; Zhao, M. Electrotaxis and wound healing: Experimental methods to study electric fields as a directional signal for cell migration. In Chemotaxis: Methods and Protocols; Humana Press: Totowa, NJ, USA, 2009; pp. 77–97. [Google Scholar]
  28. Messerli, M.A.; Graham, D.M. Extracellular electrical fields direct wound healing and regeneration. Biol. Bull. 2011, 221, 79–92. [Google Scholar] [CrossRef] [PubMed]
  29. Yao, L.; Shanley, L.; McCaig, C.; Zhao, M. Small applied electric fields guide migration of hippocampal neurons. J. Cell. Physiol. 2008, 216, 527–535. [Google Scholar] [CrossRef]
  30. Hoare, J.I.; Rajnicek, A.M.; McCaig, C.D.; Barker, R.N.; Wilson, H.M. Electric fields are novel determinants of human macrophage functions. J. Leukoc. Biol. 2016, 99, 1141–1151. [Google Scholar] [CrossRef]
  31. Simpson, M.J.; Lo, K.Y.; Sun, Y.S. Quantifying the roles of random motility and directed motility using advection-diffusion theory for a 3T3 fibroblast cell migration assay stimulated with an electric field. BMC Syst. Biol. 2017, 11, 39. [Google Scholar] [CrossRef]
  32. Liu, X.; Duan, X. Mechanisms and Treatments of Peripheral Nerve Injury. Ann. Plast. Surg. 2023, 91, 313–318. [Google Scholar] [CrossRef]
  33. Li, J.; Nandagopal, S.; Wu, D.; Romanuik, S.F.; Paul, K.; Thomson, D.J.; Lin, F. Activated T lymphocytes migrate toward the cathode of DC electric fields in microfluidic devices. Lab. Chip 2011, 11, 1298–1304. [Google Scholar] [CrossRef] [PubMed]
  34. Majid, A. Electroceuticals. In Advance in Electrostimulation Therapies; Springer: Berlin/Heidelberg, Germany, 2017. [Google Scholar]
  35. Maeng, W.Y.; Tseng, W.L.; Li, S.; Koo, J.; Hsueh, Y.Y. Electroceuticals for peripheral nerve regeneration. Biofabrication 2022, 14, 42002. [Google Scholar] [CrossRef] [PubMed]
  36. Wu, P.; Chen, P.; Xu, C.; Wang, Q.; Zhang, F.; Yang, K.; Jiang, W.; Feng, J.; Luo, Z. Ultrasound-driven in vivo electrical stimulation based on biodegradable piezoelectric nanogenerators for enhancing and monitoring the nerve tissue repair. Nano Energy 2022, 102, 107707. [Google Scholar] [CrossRef]
  37. Cafarelli, A.; Marino, A.; Vannozzi, L.; Puigmarti-Luis, J.; Pane, S.; Ciofani, G.; Ricotti, L. Piezoelectric Nanomaterials Activated by Ultrasound: The Pathway from Discovery to Future Clinical Adoption. ACS Nano 2021, 15, 11066–11086. [Google Scholar] [CrossRef] [PubMed]
  38. Yang, S.; Wang, Y.; Liang, X. Piezoelectric Nanomaterials Activated by Ultrasound in Disease Treatment. Pharmaceutics 2023, 15, 1338. [Google Scholar] [CrossRef] [PubMed]
  39. Hoffman, H. Acceleration and retardation of the process of axon-sprouting in partially denervated muscles. Aust. J. Exp. Biol. Med. Sci. 1952, 30, 541–566. [Google Scholar] [CrossRef] [PubMed]
  40. Nix, W.A.; Hopf, H.C. Electrical stimulation of regenerating nerve and its effect on motor recovery. Brain Res. 1983, 272, 21–25. [Google Scholar] [CrossRef]
  41. Pockett, S.; Gavin, R.M. Acceleration of peripheral nerve regeneration after crush injury in rat. Neurosci. Lett. 1985, 59, 221–224. [Google Scholar] [CrossRef]
  42. Russo, T.L.; Peviani, S.M.; Freria, C.M.; Gigo-Benato, D.; Geuna, S.; Salvini, T.F. Electrical stimulation based on chronaxie reduces atrogin-1 and myoD gene expressions in denervated rat muscle. Muscle Nerve 2007, 35, 87–97. [Google Scholar] [CrossRef]
  43. Gordon, T.; English, A.W. Strategies to promote peripheral nerve regeneration: Electrical stimulation and/or exercise. Eur. J. Neurosci. 2016, 43, 336–350. [Google Scholar] [CrossRef]
  44. Kong, L.; Gao, X.; Qian, Y.; Sun, W.; You, Z.; Fan, C. Biomechanical microenvironment in peripheral nerve regeneration: From pathophysiological understanding to tissue engineering development. Theranostics 2022, 12, 4993–5014. [Google Scholar] [CrossRef] [PubMed]
  45. Zuo, K.J.; Gordon, T.; Chan, K.M.; Borschel, G.H. Electrical stimulation to enhance peripheral nerve regeneration: Update in molecular investigations and clinical translation. Exp. Neurol. 2020, 332, 113397. [Google Scholar] [CrossRef] [PubMed]
  46. Scarpa, E.; Mayor, R. Collective cell migration in development. J. Cell Biol. 2016, 212, 143–155. [Google Scholar] [CrossRef] [PubMed]
  47. Pullar, C.E.; Baier, B.S.; Kariya, Y.; Russell, A.J.; Horst, B.A.; Marinkovich, M.P.; Isseroff, R.R. beta4 integrin and epidermal growth factor coordinately regulate electric field-mediated directional migration via Rac1. Mol. Biol. Cell 2006, 17, 4925–4935. [Google Scholar] [CrossRef] [PubMed]
  48. Zhao, M.; Pu, J.; Forrester, J.V.; McCaig, C.D. Membrane lipids, EGF receptors, and intracellular signals colocalize and are polarized in epithelial cells moving directionally in a physiological electric field. FASEB J. 2002, 16, 857–859. [Google Scholar] [CrossRef]
  49. Fang, K.S.; Ionides, E.; Oster, G.; Nuccitelli, R.; Isseroff, R.R. Epidermal growth factor receptor relocalization and kinase activity are necessary for directional migration of keratinocytes in DC electric fields. J. Cell Sci. 1999, 112(Pt. 12), 1967–1978. [Google Scholar] [CrossRef]
  50. Orida, N.; Poo, M.M. Electrophoretic movement and localisation of acetylcholine receptors in the embryonic muscle cell membrane. Nature 1978, 275, 31–35. [Google Scholar] [CrossRef]
  51. Poo, M.; Robinson, K.R. Electrophoresis of concanavalin A receptors along embryonic muscle cell membrane. Nature 1977, 265, 602–605. [Google Scholar] [CrossRef]
  52. Palza, H.; Zapata, P.A.; Angulo-Pineda, C. Electroactive smart polymers for biomedical applications. Materials 2019, 12, 277. [Google Scholar] [CrossRef]
  53. Rahman, M.; Mahady Dip, T.; Padhye, R.; Houshyar, S. Review on electrically conductive smart nerve guide conduit for peripheral nerve regeneration. J. Biomed. Mater. Res. Part A 2023, 111, 1916–1950. [Google Scholar] [CrossRef]
  54. Chan, K.M.; Curran, M.W.; Gordon, T. The use of brief post-surgical low frequency electrical stimulation to enhance nerve regeneration in clinical practice. J. Physiol. 2016, 594, 3553–3559. [Google Scholar] [CrossRef] [PubMed]
  55. Zhao, M.; Penninger, J.; Isseroff, R.R. Electrical Activation of Wound-Healing Pathways. Adv. Ski. Wound Care 2010, 1, 567–573. [Google Scholar] [CrossRef]
  56. Nakajima, K.I.; Zhu, K.; Sun, Y.H.; Hegyi, B.; Zeng, Q.; Murphy, C.J.; Small, J.V.; Chen-Izu, Y.; Izumiya, Y.; Penninger, J.M.; et al. KCNJ15/Kir4.2 couples with polyamines to sense weak extracellular electric fields in galvanotaxis. Nat. Commun. 2015, 6, 8532. [Google Scholar] [CrossRef]
  57. Al-Majed, A.A.; Tam, S.L.; Gordon, T. Electrical stimulation accelerates and enhances expression of regeneration-associated genes in regenerating rat femoral motoneurons. Cell. Mol. Neurobiol. 2004, 24, 379–402. [Google Scholar] [CrossRef] [PubMed]
  58. Sharma, N.; Marzo, S.J.; Jones, K.J.; Foecking, E.M. Electrical stimulation and testosterone differentially enhance expression of regeneration-associated genes. Exp. Neurol. 2010, 223, 183–191. [Google Scholar] [CrossRef]
  59. Huang, J.; Ye, Z.; Hu, X.; Lu, L.; Luo, Z. Electrical stimulation induces calcium-dependent release of NGF from cultured Schwann cells. Glia 2010, 58, 622–631. [Google Scholar] [CrossRef]
  60. Chang, H.F.; Lee, Y.S.; Tang, T.K.; Cheng, J.Y. Pulsed DC Electric Field-Induced Differentiation of Cortical Neural Precursor Cells. PLoS ONE 2016, 11, e0158133. [Google Scholar] [CrossRef]
  61. Willand, M.P. Electrical Stimulation Enhances Reinnervation After Nerve Injury. Eur. J. Transl. Myol. 2015, 25, 243–248. [Google Scholar] [CrossRef]
  62. Gordon, T.; Amirjani, N.; Edwards, D.C.; Chan, K.M. Brief post-surgical electrical stimulation accelerates axon regeneration and muscle reinnervation without affecting the functional measures in carpal tunnel syndrome patients. Exp. Neurol. 2010, 223, 192–202. [Google Scholar] [CrossRef]
  63. Power, H.A.; Morhart, M.J.; Olson, J.L.; Chan, K.M. Postsurgical Electrical Stimulation Enhances Recovery Following Surgery for Severe Cubital Tunnel Syndrome: A Double-Blind Randomized Controlled Trial. Neurosurgery 2020, 86, 769–777. [Google Scholar] [CrossRef]
  64. Barber, B.; Seikaly, H.; Ming Chan, K.; Beaudry, R.; Rychlik, S.; Olson, J.; Curran, M.; Dziegielewski, P.; Biron, V.; Harris, J.; et al. Intraoperative Brief Electrical Stimulation of the Spinal Accessory Nerve (BEST SPIN) for prevention of shoulder dysfunction after oncologic neck dissection: A double-blinded, randomized controlled trial. J. Otolaryngol. Head. Neck Surg. 2018, 47, 7. [Google Scholar] [CrossRef] [PubMed]
  65. Casal, D.; Mota-Silva, E.; Iria, I.; Pais, D.; Farinho, A.; Alves, S.; Pen, C.; Mascarenhas-Lemos, L.; Ferreira-Silva, J.; Ferraz-Oliveira, M.; et al. Functional and Physiological Methods of Evaluating Median Nerve Regeneration in the Rat. J. Vis. Exp. 2020, 158, e59767. [Google Scholar] [CrossRef]
  66. Casal, D.; Pais, D.; Mota-Silva, E.; Pelliccia, G.; Iria, I.; Videira, P.A.; Mendes, M.M.; Goyri-O’Neill, J.; Mouzinho, M.M. Reconstruction of a long defect of the ulnar artery and nerve with an arterialized neurovenous free flap in a teenager: A case report and literature review. Microsurgery 2018, 38, 209–217. [Google Scholar] [CrossRef] [PubMed]
  67. Casal, D.; Mota-Silva, E.; Iria, I.; Alves, S.; Farinho, A.; Pen, C.; Lourenco-Silva, N.; Mascarenhas-Lemos, L.; Silva-Ferreira, J.; Ferraz-Oliveira, M.; et al. Reconstruction of a 10-mm-long median nerve gap in an ischemic environment using autologous conduits with different patterns of blood supply: A comparative study in the rat. PLoS ONE 2018, 13, e0195692. [Google Scholar] [CrossRef]
  68. Alvarez-Lorenzo, C.; Zarur, M.; Seijo-Rabina, A.; Blanco-Fernandez, B.; Rodriguez-Moldes, I.; Concheiro, A. Physical stimuli-emitting scaffolds: The role of piezoelectricity in tissue regeneration. Mater. Today Bio 2023, 22, 100740. [Google Scholar] [CrossRef] [PubMed]
  69. Anderson, M.; Shelke, N.B.; Manoukian, O.S.; Yu, X.; McCullough, L.D.; Kumbar, S.G. Peripheral Nerve Regeneration Strategies: Electrically Stimulating Polymer Based Nerve Growth Conduits. Crit. Rev. Biomed. Eng. 2015, 43, 131–159. [Google Scholar] [CrossRef]
  70. Sarker, M.D.; Naghieh, S.; McInnes, A.D.; Schreyer, D.J.; Chen, X. Regeneration of peripheral nerves by nerve guidance conduits: Influence of design, biopolymers, cells, growth factors, and physical stimuli. Prog. Neurobiol. 2018, 171, 125–150. [Google Scholar] [CrossRef]
  71. Vijayavenkataraman, S. Nerve guide conduits for peripheral nerve injury repair: A review on design, materials and fabrication methods. Acta Biomater. 2020, 106, 54–69. [Google Scholar] [CrossRef]
  72. Ferrigno, B.; Bordett, R.; Duraisamy, N.; Moskow, J.; Arul, M.R.; Rudraiah, S.; Nukavarapu, S.P.; Vella, A.T.; Kumbar, S.G. Bioactive polymeric materials and electrical stimulation strategies for musculoskeletal tissue repair and regeneration. Bioact. Mater. 2020, 5, 468–485. [Google Scholar] [CrossRef]
  73. Koo, J.; MacEwan, M.R.; Kang, S.K.; Won, S.M.; Stephen, M.; Gamble, P.; Xie, Z.; Yan, Y.; Chen, Y.Y.; Shin, J.; et al. Wireless bioresorbable electronic system enables sustained nonpharmacological neuroregenerative therapy. Nat. Med. 2018, 24, 1830–1836. [Google Scholar] [CrossRef]
  74. Chorsi, M.T.; Le, T.T.; Lin, F.; Vinikoor, T.; Das, R.; Stevens, J.F.; Mundrane, C.; Park, J.; Tran, K.T.M.; Liu, Y.; et al. Highly piezoelectric, biodegradable, and flexible amino acid nanofibers for medical applications. Sci. Adv. 2023, 9, eadg6075. [Google Scholar] [CrossRef] [PubMed]
  75. Lee, H.M.; Bharathi, K.; Kim, D.K. Processing and characterization of aluminum nitride ceramics for high thermal conductivity. Advanced Engineering Materials. 2014, 16, 655–669. [Google Scholar] [CrossRef]
  76. Mei, T.; Dai, Q.; Zheng, W.; Chen, T. Strain properties and piezoelectric constant of lead-free barium titanate ceramics. Materials Research Express. 2019, 6, 106301. [Google Scholar] [CrossRef]
  77. Lay, R.; Deijs, G.S.; Malmström, J. The intrinsic piezoelectric properties of materials–a review with a focus on biological materials. RSC advances. 2021, 11, 30657–30673. [Google Scholar] [CrossRef] [PubMed]
  78. Goonoo, N.; Bhaw-Luximon, A. Piezoelectric polymeric scaffold materials as biomechanical cellular stimuli to enhance tissue regeneration. Mater. Today Commun. 2022, 31, 103491. [Google Scholar] [CrossRef]
  79. Farokhi, M.; Mottaghitalab, F.; Shokrgozar, M.A.; Kaplan, D.L.; Kim, H.-W.; Kundu, S.C. Prospects of peripheral nerve tissue engineering using nerve guide conduits based on silk fibroin protein and other biopolymers. Int. Mater. Rev. 2017, 62, 367–391. [Google Scholar] [CrossRef]
  80. Jabbari, F.; Babaeipour, V.; Bakhtiari, S. Bacterial cellulose-based composites for nerve tissue engineering. Int. J. Biol. Macromol. 2022, 217, 120–130. [Google Scholar] [CrossRef]
  81. Kim, K.; Ha, M.; Choi, B.; Joo, S.H.; Kang, H.S.; Park, J.H.; Gu, B.; Park, C.; Park, C.; Kim, J.; et al. Biodegradable, electro-active chitin nanofiber films for flexible piezoelectric transducers. Nano Energy 2018, 48, 275–283. [Google Scholar] [CrossRef]
  82. Zhang, M.; An, H.; Zhang, F.; Jiang, H.; Wan, T.; Wen, Y.; Han, N.; Zhang, P. Prospects of Using Chitosan-Based Biopolymers in the Treatment of Peripheral Nerve Injuries. Int. J. Mol. Sci. 2023, 24, 12956. [Google Scholar] [CrossRef]
  83. Ccorahua, R.; Huaroto, J.; Luyo, C.; Quintana, M.; Vela, E.A. Enhanced-performance bio-triboelectric nanogenerator based on starch polymer electrolyte obtained by a cleanroom-free processing method. Nano Energy 2019, 59, 610–618. [Google Scholar] [CrossRef]
  84. Farahani, A.; Zarei-Hanzaki, A.; Abedi, H.R.; Haririan, I.; Akrami, M.; Aalipour, Z.; Tayebi, L. An investigation into the polylactic acid texturization through thermomechanical processing and the improved d33 piezoelectric outcome of the fabricated scaffolds. J. Mater. Res. Technol. 2021, 15, 6356–6366. [Google Scholar] [CrossRef] [PubMed]
  85. Wu, L.; Gao, H.; Han, Q.; Guan, W.; Sun, S.; Zheng, T.; Liu, Y.; Wang, X.; Huang, R.; Li, G. Piezoelectric materials for neuroregeneration: A review. Biomater. Sci. 2023, 11, 7296–7310. [Google Scholar] [CrossRef] [PubMed]
  86. Yao, X.; Qian, Y.; Fan, C. Electroactive nanomaterials in the peripheral nerve regeneration. J. Mater. Chem. B 2021, 9, 6958–6972. [Google Scholar] [CrossRef] [PubMed]
  87. Pryadko, A.; Surmeneva, M.A.; Surmenev, R.A. Review of hybrid materials based on polyhydroxyalkanoates for tissue engineering applications. Polymers 2021, 13, 1738. [Google Scholar] [CrossRef] [PubMed]
  88. King, I.I.I.W.E.; Bowlin, G.L. Near-field electrospinning and melt electrowriting of biomedical polymers—Progress and limitations. Polymers 2021, 13, 1097. [Google Scholar] [CrossRef] [PubMed]
  89. Frayssinet, A.; Petta, D.; Illoul, C.; Haye, B.; Markitantova, A.; Eglin, D.; Mosser, G.; D’Este, M.; Hélary, C. Extracellular matrix-mimetic composite hydrogels of cross-linked hyaluronan and fibrillar collagen with tunable properties and ultrastructure. Carbohydr. Polym. 2020, 236, 116042. [Google Scholar] [CrossRef] [PubMed]
  90. Yao, X.; Zou, S.; Fan, S.; Niu, Q.; Zhang, Y. Bioinspired silk fibroin materials: From silk building blocks extraction and reconstruction to advanced biomedical applications. Materials Today Bio 2022, 16, 100381. [Google Scholar] [CrossRef] [PubMed]
  91. Xing, X.; Han, Y.; Cheng, H. Biomedical applications of chitosan/silk fibroin composites: A review. Int. J. Biol. Macromol. 2023, 240, 124407. [Google Scholar] [CrossRef]
  92. Takeya, H.; Itai, S.; Kimura, H.; Kurashina, Y.; Amemiya, T.; Nagoshi, N.; Iwamoto, T.; Sato, K.; Shibata, S.; Matsumoto, M.; et al. Schwann cell-encapsulated chitosan-collagen hydrogel nerve conduit promotes peripheral nerve regeneration in rodent sciatic nerve defect models. Sci. Rep. 2023, 13, 11932. [Google Scholar] [CrossRef]
  93. Mohamadi, F.; Ebrahimi-Barough, S.; Nourani, M.R.; Ahmadi, A.; Ai, J. Use new poly (ε-caprolactone/collagen/NBG) nerve conduits along with NGF for promoting peripheral (sciatic) nerve regeneration in a rat. Artif. Cells Nanomed. Biotechnol. 2018, 46, 34–45. [Google Scholar] [CrossRef]
  94. Nawrotek, K.; Kubicka, M.; Gatkowska, J.; Wieczorek, M.; Michlewska, S.; Bekier, A.; Wach, R.; Rudnicka, K. Controlling the spatiotemporal release of nerve growth factor by chitosan/polycaprolactone conduits for use in peripheral nerve regeneration. Int. J. Mol. Sci. 2022, 23, 2852. [Google Scholar] [CrossRef] [PubMed]
  95. Chen, C.; Xi, Y.; Weng, Y. Progress in the development of graphene-based biomaterials for tissue engineering and regeneration. Materials. 2022, 15, 2164. [Google Scholar] [CrossRef] [PubMed]
  96. Stocco, E.; Barbon, S.; Emmi, A.; Tiengo, C.; Macchi, V.; De Caro, R.; Porzionato, A. Bridging Gaps in Peripheral Nerves: From Current Strategies to Future Perspectives in Conduit Design. Int. J. Mol. Sci. 2023, 24, 9170. [Google Scholar] [CrossRef] [PubMed]
  97. Cheng, Y.; Xu, Y.; Qian, Y.; Chen, X.; Ouyang, Y.; Yuan, W.-E. 3D structured self-powered PVDF/PCL scaffolds for peripheral nerve regeneration. Nano Energy 2020, 69, 104411. [Google Scholar] [CrossRef]
  98. Javidi, H.; Ramazani Saadatabadi, A.; Sadrnezhaad, S.K.; Najmoddin, N. Preparation and characterization of self-stimuli conductive nerve regeneration conduit using co-electrospun nanofibers filled with gelatin-chitosan hydrogels containing polyaniline-graphene-ZnO nanoparticles. Int. J. Polym. Mater. Polym. Biomater. 2022, 73, 165–175. [Google Scholar] [CrossRef]
  99. Guy, I.L.; Muensit, S.; Goldys, E.M. Extensional piezoelectric coefficients of gallium nitride and aluminum nitride. Appl. Phys. Lett. 1999, 75, 4133–4135. [Google Scholar] [CrossRef]
  100. Wang, K.; Ruan, J.; Song, H.; Zhang, J.; Wo, Y.; Guo, S.; Cui, D. Biocompatibility of Graphene Oxide. Nanoscale Res. Lett. 2011, 6, 8. [Google Scholar] [CrossRef]
  101. Zhang, Y.; Chen, S.; Xiao, Z.; Liu, X.; Wu, C.; Wu, K.; Liu, A.; Wei, D.; Sun, J.; Zhou, L.; et al. Magnetoelectric Nanoparticles Incorporated Biomimetic Matrix for Wireless Electrical Stimulation and Nerve Regeneration. Adv. Heal. Mater. 2021, 10, e2100695. [Google Scholar] [CrossRef]
  102. Dong, R.; Ma, P.X.; Guo, B. Conductive biomaterials for muscle tissue engineering. Biomaterials 2020, 229, 119584. [Google Scholar] [CrossRef]
  103. Yin, L.; Cheng, H.; Mao, S.; Haasch, R.; Liu, Y.; Xie, X.; Hwang, S.W.; Jain, H.; Kang, S.K.; Su, Y. Dissolvable metals for transient electronics. Adv. Funct. Mater. 2014, 24, 645–658. [Google Scholar] [CrossRef]
  104. Choi, Y.; Koo, J.; Rogers, J.A. Inorganic materials for transient electronics in biomedical applications. MRS Bull. 2020, 45, 103–112. [Google Scholar] [CrossRef]
  105. Mei, D.; Lamaka, S.V.; Lu, X.; Zheludkevich, M.L. Selecting medium for corrosion testing of bioabsorbable magnesium and other metals–a critical review. Corros. Sci. 2020, 171, 108722. [Google Scholar] [CrossRef]
  106. Wu, J. Advances in Lead-Free Piezoelectric Materials, 1st ed.; Springer: Berlin/Heidelberg, Germany, 2018. [Google Scholar]
  107. Liu, J.; Sun, L.; Xu, W.; Wang, Q.; Yu, S.; Sun, J. Current advances and future perspectives of 3D printing natural-derived biopolymers. Carbohydr. Polym. 2019, 207, 297–316. [Google Scholar] [CrossRef] [PubMed]
  108. Pina, S.; Ribeiro, V.P.; Marques, C.F.; Maia, F.R.; Silva, T.H.; Reis, R.L.; Oliveira, J.M. Scaffolding Strategies for Tissue Engineering and Regenerative Medicine Applications. Materials 2019, 12, 1824. [Google Scholar] [CrossRef] [PubMed]
  109. Jiang, H.; Qian, Y.; Fan, C.; Ouyang, Y. Polymeric Guide Conduits for Peripheral Nerve Tissue Engineering. Front. Bioeng. Biotechnol. 2020, 8, 582646. [Google Scholar] [CrossRef] [PubMed]
  110. Gregory, H.; Phillips, J.B. Materials for peripheral nerve repair constructs: Natural proteins or synthetic polymers? Neurochem. Int. 2021, 143, 104953. [Google Scholar] [CrossRef] [PubMed]
  111. Sharma, S.; Goel, S.A. 3D printing and its future in medical world. J. Med. Res. Innov. 2019, 3, e000141. [Google Scholar] [CrossRef]
  112. Zennifer, A.; Thangadurai, M.; Sundaramurthi, D.; Sethuraman, S. Additive manufacturing of peripheral nerve conduits—Fabrication methods, design considerations and clinical challenges. SLAS Technol. 2023, 28, 102–126. [Google Scholar] [CrossRef]
  113. Pan, X.; Sun, B.; Mo, X. Electrospun polypyrrole-coated polycaprolactone nanoyarn nerve guidance conduits for nerve tissue engineering. Front. Mater. Sci. 2018, 12, 438–446. [Google Scholar] [CrossRef]
  114. Najjari, A.; Mehdinavaz Aghdam, R.; Ebrahimi, S.S.; Suresh, K.S.; Krishnan, S.; Shanthi, C.; Ramalingam, M. Smart piezoelectric biomaterials for tissue engineering and regenerative medicine: A review. Biomed. Eng. Biomed. Tech. 2022, 67, 71–88. [Google Scholar] [CrossRef]
  115. Wang, H. A Review of the Effects of Collagen Treatment in Clinical Studies. Polymers 2021, 13, 3868. [Google Scholar] [CrossRef] [PubMed]
  116. Casimiro, M.H.; Ferreira, L.M.; Santos, P.M.P.; Leal, J.P.; Rodrigues, G.; Iria, I.; Alves, S.; Pais, D.; Casal, D. Chitosan-Based Membranes for Skin Wound Repair in a Dorsal Fold Chamber Rat Model. Pharmaceutics 2022, 14, 2736. [Google Scholar] [CrossRef]
  117. Jiang, Z.; Zhang, Y.; Wang, Y.; Wang, S.; Chang, J.; Liu, W.; Han, B. Multichannel nerve conduit based on chitosan derivates for peripheral nerve regeneration and Schwann cell survival. Carbohydr. Polym. 2023, 301, 120327. [Google Scholar] [CrossRef] [PubMed]
  118. Silva, S.S.; Fernandes, E.M.; Pina, S.; Silva-Correia, J.; Vieira, S.; Oliveira, J.M.; Reis, R.L. 2.11 Polymers of Biological Origin. In Comprehensive Biomaterials II; Ducheyne, P., Ed.; Elsevier: Oxford, UK, 2017. [Google Scholar] [CrossRef]
  119. Mohammadpourfazeli, S.; Arash, S.; Ansari, A.; Yang, S.; Mallick, K.; Bagherzadeh, R. Future prospects and recent developments of polyvinylidene fluoride (PVDF) piezoelectric polymer; fabrication methods, structure, and electro-mechanical properties. RSC Adv. 2022, 13, 370–387. [Google Scholar] [CrossRef] [PubMed]
  120. Orkwis, J.A.; Wolf, A.K.; Mularczyk, Z.J.; Bryan, A.E.; Smith, C.S.; Brown, R.; Krutko, M.; McCann, A.; Collar, R.M.; Esfandiari, L. Mechanical stimulation of a bioactive, functionalized PVDF-TrFE scaffold provides electrical signaling for nerve repair applications. Biomater. Adv. 2022, 140, 213081. [Google Scholar] [CrossRef] [PubMed]
  121. Lu, P.; Wang, G.; Qian, T.; Cai, X.; Zhang, P.; Li, M.; Shen, Y.; Xue, C.; Wang, H. The balanced microenvironment regulated by the degradants of appropriate PLGA scaffolds and chitosan conduit promotes peripheral nerve regeneration. Mater. Today Bio 2021, 12, 100158. [Google Scholar] [CrossRef] [PubMed]
  122. Wang, Q.; Wang, H.; Ma, Y.; Cao, X.; Gao, H. Effects of electroactive materials on nerve cell behaviors and applications in peripheral nerve repair. Biomater. Sci. 2022, 10, 6061–6076. [Google Scholar] [CrossRef]
  123. Zare, P.; Aleemardani, M.; Seifalian, A.; Bagher, Z.; Seifalian, A.M. Graphene oxide: Opportunities and challenges in biomedicine. Nanomaterials 2021, 11, 1083. [Google Scholar] [CrossRef]
  124. Du, L.; Li, T.; Jin, F.; Wang, Y.; Li, R.; Zheng, J.; Wang, T.; Feng, Z.-Q. Design of high conductive and piezoelectric poly (3, 4-ethylenedioxythiophene)/chitosan nanofibers for enhancing cellular electrical stimulation. J. Colloid. Interface Sci. 2020, 559, 65–75. [Google Scholar] [CrossRef]
  125. Askari, S.; Bozcheloei, Z.A. Piezoelectric composites in neural tissue engineering: Material and fabrication techniques. J. Compos. Compd. 2022, 4, 37–46. [Google Scholar] [CrossRef]
  126. Wu, H.; Zhang, J.; Luo, Y.; Wan, Y.; Sun, S. Mechanical properties and permeability of porous chitosan–poly (p-dioxanone)/silk fibroin conduits used for peripheral nerve repair. J. Mech. Behav. Biomed. Mater. 2015, 50, 192–205. [Google Scholar] [CrossRef] [PubMed]
  127. Chorsi, M.T.; Curry, E.J.; Chorsi, H.T.; Das, R.; Baroody, J.; Purohit, P.K.; Ilies, H.; Nguyen, T.D. Piezoelectric Biomaterials for Sensors and Actuators. Adv. Mater. 2019, 31, e1802084. [Google Scholar] [CrossRef] [PubMed]
  128. Li, T.; Qu, M.; Carlos, C.; Gu, L.; Jin, F.; Yuan, T.; Wu, X.; Xiao, J.; Wang, T.; Dong, W.; et al. High-Performance Poly(vinylidene difluoride)/Dopamine Core/Shell Piezoelectric Nanofiber and Its Application for Biomedical Sensors. Adv. Mater. 2021, 33, e2006093. [Google Scholar] [CrossRef] [PubMed]
  129. Smith, M.; Kar-Narayan, S. Piezoelectric polymers: Theory, challenges and opportunities. Int. Mater. Rev. 2022, 67, 65–88. [Google Scholar] [CrossRef]
  130. Pi, W.; Rao, F.; Cao, J.; Zhang, M.; Chang, T.; Han, Y.; Zheng, Y.; Liu, S.; Li, Q.; Sun, X. Sono-electro-mechanical therapy for peripheral nerve regeneration through piezoelectric nanotracts. Nano Today 2023, 50, 101860. [Google Scholar] [CrossRef]
  131. Padilla, F.; Puts, R.; Vico, L.; Raum, K. Stimulation of bone repair with ultrasound: A review of the possible mechanic effects. Ultrasonics 2014, 54, 1125–1145. [Google Scholar] [CrossRef] [PubMed]
  132. Tandon, B.; Blaker, J.J.; Cartmell, S.H. Piezoelectric materials as stimulatory biomedical materials and scaffolds for bone repair. Acta Biomater. 2018, 73, 1–20. [Google Scholar] [CrossRef]
  133. Dai, Y.; Lu, T.; Shao, M.; Lyu, F. Recent advances in PLLA-based biomaterial scaffolds for neural tissue engineering: Fabrication, modification, and applications. Front. Bioeng. Biotechnol. 2022, 10, 1011783. [Google Scholar] [CrossRef]
  134. Gryshkov, O.; Al Halabi, F.; Kuhn, A.I.; Leal-Marin, S.; Freund, L.J.; Forthmann, M.; Meier, N.; Barker, S.A.; Haastert-Talini, K.; Glasmacher, B. PVDF and P(VDF-TrFE) Electrospun Scaffolds for Nerve Graft Engineering: A Comparative Study on Piezoelectric and Structural Properties, and In Vitro Biocompatibility. Int. J. Mol. Sci. 2021, 22, 11373. [Google Scholar] [CrossRef]
  135. Li, Y.; Liao, C.; Tjong, S.C. Electrospun Polyvinylidene Fluoride-Based Fibrous Scaffolds with Piezoelectric Characteristics for Bone and Neural Tissue Engineering. Nanomaterials 2019, 9, 952. [Google Scholar] [CrossRef]
  136. Wang, X.F.; Li, M.L.; Fang, Q.Q.; Zhao, W.Y.; Lou, D.; Hu, Y.Y.; Chen, J.; Wang, X.Z.; Tan, W.Q. Flexible electrical stimulation device with Chitosan-Vaseline(R) dressing accelerates wound healing in diabetes. Bioact. Mater. 2021, 6, 230–243. [Google Scholar] [CrossRef] [PubMed]
  137. Dixon, A.R.; Jariwala, S.H.; Bilis, Z.; Loverde, J.R.; Pasquina, P.F.; Alvarez, L.M. Bridging the gap in peripheral nerve repair with 3D printed and bioprinted conduits. Biomaterials 2018, 186, 44–63. [Google Scholar] [CrossRef] [PubMed]
  138. Fornasari, B.E.; Carta, G.; Gambarotta, G.; Raimondo, S. Natural-Based Biomaterials for Peripheral Nerve Injury Repair. Front. Bioeng. Biotechnol. 2020, 8, 554257. [Google Scholar] [CrossRef]
  139. Li, R.; Li, D.-H.; Zhang, H.-Y.; Wang, J.; Li, X.-K.; Xiao, J. Growth factors-based therapeutic strategies and their underlying signaling mechanisms for peripheral nerve regeneration. Acta Pharmacol. Sin. 2020, 41, 1289–1300. [Google Scholar] [CrossRef] [PubMed]
  140. Xia, B.; Lv, Y. Dual-delivery of VEGF and NGF by emulsion electrospun nanofibrous scaffold for peripheral nerve regeneration. Mater. Sci. Eng. C Mater. Biol. Appl. 2018, 82, 253–264. [Google Scholar] [CrossRef] [PubMed]
  141. Joshi, A.; Choudhury, S.; Baghel, V.S.; Ghosh, S.; Gupta, S.; Lahiri, D.; Ananthasuresh, G.K.; Chatterjee, K. 4D Printed Programmable Shape-Morphing Hydrogels as Intraoperative Self-Folding Nerve Conduits for Sutureless Neurorrhaphy. Adv. Heal. Mater. 2023, 12, e2300701. [Google Scholar] [CrossRef] [PubMed]
  142. Tan, M.-H.; Xu, X.-H.; Yuan, T.-J.; Hou, X.; Wang, J.; Jiang, Z.-H.; Peng, L.-H. Self-powered smart patch promotes skin nerve regeneration and sensation restoration by delivering biological-electrical signals in program. Biomaterials 2022, 283, 121413. [Google Scholar] [CrossRef]
  143. Dolbashid, A.S.; Mokhtar, M.S.; Muhamad, F.; Ibrahim, F. Potential applications of human artificial skin and electronic skin (e-skin): A review. Bioinspired Biomim. Nanobiomaterials 2018, 7, 53–64. [Google Scholar] [CrossRef]
  144. Sun, Q.J.; Lai, Q.T.; Tang, Z.; Tang, X.G.; Zhao, X.H.; Roy, V.A. Advanced Functional Composite Materials toward E-Skin for Health Monitoring and Artificial Intelligence. Adv. Mater. Technol. 2023, 8, 2201088. [Google Scholar] [CrossRef]
  145. Wang, W.; Jiang, Y.; Zhong, D.; Zhang, Z.; Choudhury, S.; Lai, J.C.; Gong, H.; Niu, S.; Yan, X.; Zheng, Y.; et al. Neuromorphic sensorimotor loop embodied by monolithically integrated, low-voltage, soft e-skin. Science 2023, 380, 735–742. [Google Scholar] [CrossRef]
  146. Yang, J.C.; Mun, J.; Kwon, S.Y.; Park, S.; Bao, Z.; Park, S. Electronic Skin: Recent Progress and Future Prospects for Skin-Attachable Devices for Health Monitoring, Robotics, and Prosthetics. Adv. Mater. 2019, 31, e1904765. [Google Scholar] [CrossRef] [PubMed]
  147. Goodrich, J.T.; Kliot, M. History of the peripheral and cranial nerves. In Nerves and Nerve Injuries, 1st ed.; Tubbs, R.S., Rizk, E., Shoja, M.M., Loukas, M., Barbarom, N., Spinner, R.J., Eds.; Elsevier: New York, NY, USA, 2015; Volume 1, pp. 5–6. [Google Scholar]
  148. Missios, S.; Bekelis, K.; Roberts, D.W. Neurosurgery in the Byzantine Empire: The contributions of Paul of Aegina (625-690 AD). J. Neurosurg. 2014, 120, 244–249. [Google Scholar] [CrossRef] [PubMed]
  149. Geuna, S.; Tos, P.; Titolo, P.; Ciclamini, D.; Beningo, T.; Battiston, B. Update on nerve repair by biological tubulization. J. Brachial Plex. Peripher. Nerve Inj. 2014, 9, 3. [Google Scholar] [CrossRef] [PubMed]
  150. Desouches, C.; Alluin, O.; Mutaftschiev, N.; Dousset, E.; Magalon, G.; Boucraut, J.; Feron, F.; Decherchi, P. Peripheral nerve repair: 30 centuries of scientific research. Rev. Neurol. 2005, 161, 1045–1059. [Google Scholar] [CrossRef] [PubMed]
  151. Wills, A. Herophilus, Erasistratus, and the birth of neuroscience. Lancet 1999, 354, 1719–1720. [Google Scholar] [CrossRef] [PubMed]
  152. Song, S.; Wang, X.; Wang, T.; Yu, Q.; Hou, Z.; Zhu, Z.; Li, R. Additive Manufacturing of Nerve Guidance Conduits for Regeneration of Injured Peripheral Nerves. Front. Bioeng. Biotechnol. 2020, 8, 590596. [Google Scholar] [CrossRef] [PubMed]
  153. Qian, Y.; Cheng, Y.; Song, J.; Xu, Y.; Yuan, W.E.; Fan, C.; Zheng, X. Mechano-informed biomimetic polymer scaffolds by incorporating self-powered zinc oxide nanogenerators enhance motor recovery and neural function. Small 2020, 16, 2000796. [Google Scholar] [CrossRef]
  154. Noor, A.K. AI and the Future of the Machine Design. Mech. Eng. 2017, 139, 38–43. [Google Scholar] [CrossRef]
  155. Yu, X.; Zhang, T.; Li, Y. 3D Printing and Bioprinting Nerve Conduits for Neural Tissue Engineering. Polymers 2020, 12, 1637. [Google Scholar] [CrossRef]
  156. Liu, K.; Yan, L.; Li, R.; Song, Z.; Ding, J.; Liu, B.; Chen, X. 3D Printed Personalized Nerve Guide Conduits for Precision Repair of Peripheral Nerve Defects. Adv. Sci. 2022, 9, e2103875. [Google Scholar] [CrossRef]
  157. Ye, W.; Li, H.; Yu, K.; Xie, C.; Wang, P.; Zheng, Y.; Zhang, P.; Xiu, J.; Yang, Y.; Zhang, F. 3D printing of gelatin methacrylate-based nerve guidance conduits with multiple channels. Mater. Des. 2020, 192, 108757. [Google Scholar] [CrossRef]
  158. Zhan, L.; Deng, J.; Ke, Q.; Li, X.; Ouyang, Y.; Huang, C.; Liu, X.; Qian, Y. Grooved fibers: Preparation principles through electrospinning and potential applications. Adv. Fiber Mater. 2022, 203–213. [Google Scholar] [CrossRef]
  159. Xu, X.; Tao, J.; Wang, S.; Yang, L.; Zhang, J.; Zhang, J.; Liu, H.; Cheng, H.; Xu, J.; Gou, M. 3D printing of nerve conduits with nanoparticle-encapsulated RGFP966. Appl. Mater. Today 2019, 16, 247–256. [Google Scholar] [CrossRef]
  160. Tao, J.; Zhang, J.; Du, T.; Xu, X.; Deng, X.; Chen, S.; Liu, J.; Chen, Y.; Liu, X.; Xiong, M.; et al. Rapid 3D printing of functional nanoparticle-enhanced conduits for effective nerve repair. Acta Biomater. 2019, 90, 49–59. [Google Scholar] [CrossRef]
  161. Tao, J.; Liu, H.; Wu, W.; Zhang, J.; Liu, S.; Zhang, J.; Huang, Y.; Xu, X.; He, H.; Yang, S. 3D-printed nerve conduits with live platelets for effective peripheral nerve repair. Adv. Funct. Mater. 2020, 30, 2004272. [Google Scholar] [CrossRef]
  162. Khan, H.M.; Liao, X.; Sheikh, B.A.; Wang, Y.; Su, Z.; Guo, C.; Li, Z.; Zhou, C.; Cen, Y.; Kong, Q. Smart biomaterials and their potential applications in tissue engineering. J. Mater. Chem. B 2022, 10, 6859–6895. [Google Scholar] [CrossRef] [PubMed]
  163. Deng, R.; Luo, Z.; Rao, Z.; Lin, Z.; Chen, S.; Zhou, J.; Zhu, Q.; Liu, X.; Bai, Y.; Quan, D. Decellularized extracellular matrix containing electrospun fibers for nerve regeneration: A comparison between core–shell structured and preblended composites. Adv. Fiber Mater. 2022, 4, 503–519. [Google Scholar] [CrossRef]
  164. Zhao, M.; Bai, H.; Wang, E.; Forrester, J.V.; McCaig, C.D. Electrical stimulation directly induces pre-angiogenic responses in vascular endothelial cells by signaling through VEGF receptors. J. Cell Sci. 2004, 117, 397–405. [Google Scholar] [CrossRef]
  165. Buragohain, P.; Lu, H.; Richter, C.; Schenk, T.; Kariuki, P.; Glinsek, S.; Funakubo, H.; Iniguez, J.; Defay, E.; Schroeder, U.; et al. Quantification of the Electromechanical Measurements by Piezoresponse Force Microscopy. Adv. Mater. 2022, 34, e2206237. [Google Scholar] [CrossRef]
  166. Shimizu, M.; Okamoto, C.; Umeda, K.; Watanabe, S.; Ando, T.; Kodera, N. An ultrafast piezoelectric Z-scanner with a resonance frequency above 1.1 MHz for high-speed atomic force microscopy. Rev. Sci. Instrum. 2022, 93, 13701. [Google Scholar] [CrossRef]
  167. Kačarević, Ž.P.; Rider, P.M.; Alkildani, S.; Retnasingh, S.; Smeets, R.; Jung, O.; Ivanišević, Z.; Barbeck, M. An introduction to 3D bioprinting: Possibilities, challenges and future aspects. Materials 2018, 11, 2199. [Google Scholar] [CrossRef] [PubMed]
  168. Butovskaya, G.; Fedorenko, A.; Roginets, L.; Sal’nikova, I. Gamma-and Electron Beam Radiation-Induced Degradation of Poly-L-Lactide. High Energy Chem. 2020, 54, 136–141. [Google Scholar]
  169. Adamus-Wlodarczyk, A.; Wach, R.A.; Ulanski, P.; Rosiak, J.M.; Socka, M.; Tsinas, Z.; Al-Sheikhly, M. On the mechanisms of the effects of ionizing radiation on diblock and random copolymers of poly (lactic acid) and poly (trimethylene carbonate). Polymers 2018, 10, 672. [Google Scholar] [CrossRef] [PubMed]
  170. Ferro, W.P.; e Silva, L.G.A. Ionizing radiation effect studies on polyamide 6.6 properties. Radiat. Phys. Chem. 2004, 71, 269–271. [Google Scholar] [CrossRef]
  171. Thibodeau, A.; Galbraith, T.; Fauvel, C.M.; Khuong, H.T.; Berthod, F. Repair of peripheral nerve injuries using a prevascularized cell-based tissue-engineered nerve conduit. Biomaterials 2022, 280, 121269. [Google Scholar] [CrossRef]
  172. Qian, Y.; Xu, Y.; Yan, Z.; Jin, Y.; Chen, X.; Yuan, W.-E.; Fan, C. Boron nitride nanosheets functionalized channel scaffold favors microenvironment rebalance cocktail therapy for piezocatalytic neuronal repair. Nano Energy 2021, 83, 105779. [Google Scholar] [CrossRef]
  173. Qian, Y.; Lin, H.; Yan, Z.; Shi, J.; Fan, C. Functional nanomaterials in peripheral nerve regeneration: Scaffold design, chemical principles and microenvironmental remodeling. Mater. Today 2021, 51, 165–187. [Google Scholar] [CrossRef]
Table 1. Summary of features of different piezoelectric materials used in peripheral nerve repair.
Table 1. Summary of features of different piezoelectric materials used in peripheral nerve repair.
Type MaterialsBiocompatibilityBiodegradabilityMechanical PropertiesPyezolectric PropertiesReferences
Inorganic
Aluminum Nitride (AlN), ++++, rigid, brittle+[75]
Barium titanate (BaTiO3)+++++, hard, fracture resistant+++[76]
Lead zirconate titanate (PZT-5H)++++++, hard, fracture resistant+++[77]
Polyvinylidene fluoride (PVDF)+++++, flexible++[74]
Graphene (G)
Organic
Natural
Amino acidsGlycine +++++++, readily soluble in the body; hard and brittle++[5]
Cysteine+++++++, readily soluble in the body; hard and brittle+[5]
Alanine+++++++, readily soluble in the body; hard and brittle+[5]
Threonine+++++++, readily soluble in the body; hard and brittle+[5]
Diphenylalanine+++++++, readily soluble in the body; hard and brittle+[5]
ProteinsCollagen+++++++++[78]
Silk++++++++, exceptional mechanical strength and flexibility+[79]
PolysaccharidesCellulose++++++++, excellent strength and flexibility+[80]
Chitin++++++++, high strength and stiffness+[81]
Chitosan+++++++, pliable++[82]
Alginate++++++, fragile+[83]
Synthetic Poly-lactic acid (PLA), +++++++, rigid and brittle+[84]
Polyvinyl alcohol (PVA)+++++++++, soft+[85]
Polycaprolactone (PCL)+++++++++, soft+[86]
Polyamide (PA)++++++++, flexible, resistant+[85]
Polypyrrole (PPy)+++++, pyroelectric properties+[86]
Polyurethane (PU) ++++, flexible, resistant+[85]
Poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) (PHBHHx)++++++, flexible, resistant++[87]
Poly-γ-benzyl-L-glutamate (PBLG)+++++++++, flexible, resistant+++[88]
Composites 1
Natural Collagen/Tyramine Hyaluronic Acid derivative (HA-Tyr)
Hydrogel
++++++++++++[89]
Silk fibroin/Alginate (SF/Alg)++++++++++++[90]
Chitosan/Silk fibroin++++++++++++[91]
Chitosan/Collagen++++++++++++[92]
Natural/Synthetic Collagen/PCL++++++++++++[93]
Chitosan/PCL++++++++++++[94]
Synthetic G/PCL ++++++++++++[95]
G/PPy/PLA ++++++++++++[96]
PVDF/PCL ++++++++++++[97]
PVDF/G++++++++++++[98]
1 Composite materials have tunable compositions which allow a better match of the biological and physical properties of the devices. +, poor; ++, good; +++, excellent.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Casal, D.; Casimiro, M.H.; Ferreira, L.M.; Leal, J.P.; Rodrigues, G.; Lopes, R.; Moura, D.L.; Gonçalves, L.; Lago, J.B.; Pais, D.; et al. Review of Piezoelectrical Materials Potentially Useful for Peripheral Nerve Repair. Biomedicines 2023, 11, 3195. https://doi.org/10.3390/biomedicines11123195

AMA Style

Casal D, Casimiro MH, Ferreira LM, Leal JP, Rodrigues G, Lopes R, Moura DL, Gonçalves L, Lago JB, Pais D, et al. Review of Piezoelectrical Materials Potentially Useful for Peripheral Nerve Repair. Biomedicines. 2023; 11(12):3195. https://doi.org/10.3390/biomedicines11123195

Chicago/Turabian Style

Casal, Diogo, Maria Helena Casimiro, Luís M. Ferreira, João Paulo Leal, Gabriela Rodrigues, Raquel Lopes, Diogo Lino Moura, Luís Gonçalves, João B. Lago, Diogo Pais, and et al. 2023. "Review of Piezoelectrical Materials Potentially Useful for Peripheral Nerve Repair" Biomedicines 11, no. 12: 3195. https://doi.org/10.3390/biomedicines11123195

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