Skip to main content

Electrical stimulation as a novel tool for regulating cell behavior in tissue engineering

Abstract

Recently, electrical stimulation as a physical stimulus draws lots of attention. It shows great potential in disease treatment, wound healing, and mechanism study because of significant experimental performance. Electrical stimulation can activate many intracellular signaling pathways, and influence intracellular microenvironment, as a result, affect cell migration, cell proliferation, and cell differentiation. Electrical stimulation is using in tissue engineering as a novel type of tool in regeneration medicine. Besides, with the advantages of biocompatible conductive materials coming into view, the combination of electrical stimulation with suitable tissue engineered scaffolds can well combine the benefits of both and is ideal for the field of regenerative medicine. In this review, we summarize the various materials and latest technologies to deliver electrical stimulation. The influences of electrical stimulation on cell alignment, migration and its underlying mechanisms are discussed. Then the effect of electrical stimulation on cell proliferation and differentiation are also discussed.

Background

Regenerative medicine and tissue engineering are new scientific and technological undertakings of biomedicine that combine aspects of medicine, cell, and molecular biology, materials science and bioengineering, to regenerate, repair or replace tissues or organs [1, 2]. The three key factors of regenerative medicine and tissue engineering include seed cell, scaffold, and stimulating factor [3]. Tissue engineering scaffold could deliver specific cells to the damaged site and as a medium to provide stimuli, with the help of an appropriate structure, similar composition to natural tissue. Scaffold mimic nature tissue’s mechanical properties and desired biological properties, to ensure in vivo support, optimum diffusion of nutrients, and encourage cellular communication [4, 5], which is an indispensable part of tissue engineering. Seed cell include non-stem cells and stem cells. The non-stem cells includes: Schwann cells (SCs), osteoblasts, fibroblast, and endothelial cells (ECs); the stem cells can be divided into two types: (1) adult stem cell: Adipose derived stem cells and muscle derived stem cells; and (2) non-adult stem cell: Embryonic stem cells (ESCs), neural stem cells (NSCs) and bone mesenchymal stem cells (BMSCs) [6]. Stem cells that possess strong self-renewal and multilineage differentiation potential [7], in cases better than non-stem cells. A cell-seeded scaffold is implanted in the patient, and then the cell will produce the new tissue, is the pillars of tissue engineering [8, 9]. Providing seed cell for damaged or lost organs and tissues are the core of tissue engineering [10]. Rapid and complete regeneration of tissue or organs is a very challenging problem because transplanted cells are easily lost in host tissues and have low survival rates [11, 12]. Furthermore, if defective cells migrating to the wound site will lead to a more severe condition [13]. The loss of cellular function at the donor site and uncontrollable differentiation are limitations for the use of stem cell transplantation in regenerative medicine [13,14,15]. Thus they require the manipulation of cell behavior in vitro and in vivo, including cell proliferation, migration, differentiation, and other cellular processes [16, 17]. The choice of scaffold material, the surface topography of the scaffold, and the additional stimulating factors can manipulate cell behavior [18,19,20,21].

To date, several studies have demonstrated that both biochemical and biophysical cues could influence cell behaviors [17, 22, 23]. Different forms of stimulating factors can induce cell proliferation, differentiation, to complete tissue repair, inappropriate ES could cause cell death or no effect [24]. As a result select suitable stimulate factor could maximize the repair effect [25]. Biochemical cues include supplying chemical reagents [26] and performing chemical surface modification on scaffolds [27]. On the one side, adding growth factors, surface-immobilised biosignals, cytokines, and small molecule drugs would be immediately diluted by blood or metabolized by organisms. On the other side, the surface fixation methods of chemical reagents is also not perfect, methods such as silanization or co-precipitation are complexity and low efficiency, requiring more surface treatment to improve connection efficiency [28] and increasing the deposition rate [29]. Biophysical cues involve surface topography, substrate stiffness, compression and stretching, electric or magnetic fields, ultrasound stimulation, and photostimulation [30,31,32,33,34]. Biophysical cues have the advantages of cost-effectiveness, long life, easy to characterize, and high reproducibility, which is facilitated for a large-scale operation. As a biophysical cue, electrical stimulation (ES) has been shown to effectively relieve pain, promote blood circulation, reduce vascular and skeletal muscle tension, and promote reabsorption of edema and joint fluid in the clinic [35]. Also, many studies show that ES could effectively manipulate cell behaviors in vitro and in vivo. For example, Jaatinen et al. [36] demonstrated that the mouse myoblast cell line undergoes dramatic changes in cell morphology, viability, cell structure, and cell adhesion under pulsed monophasic currents. Kumar et al. [37] showed that the cell proliferation and osteogenic differentiation of preosteoblast could favourably be regulated under dynamic electric field conditions. Besides, with the specific parameter of ES, the neurite outgrowth of NSCs could be improved [19], and the differentiation of NSCs could also be matipulated [38]. In the human body, every cell subjected to some form of stimulation, local bioelectrical signals affect cells in a variety of tissues [39]. Indeed, providing ES is a considerable method based on recent research. ES triggers the cells themselves to deliver signals through intrinsic pathways, consequently leads to direct cell activities, including migration, differentiation, and proliferation, etc. Meanwhile, ES can be used synergistically with other techniques, reducing the cost of the whole process, has the potential of alleviating some of the problems that currently prevail in tissue engineering and regenerative medicine [17, 40,41,42].

Here, we provide an overview of the recent developments of ES in tissue engineering, for better application in regenerative medicine and tissue engineering. The main part of the review is organized into three parts: (1) summarized the conductive materials that can be applied to tissue engineering scaffolds, including types and their advantages and disadvantages. (2) Three methods of providing ES include direct coupling, capacitive coupling, and using an electromagnetic field, and their advantages and disadvantages. (3) The specific regulation of ES on cell behaviors include cell alignment, migration and its underlying mechanisms, proliferation and differentiation. Comprehensively demonstrate the potential that ES has for tissue engineering.

Different materials and methods to deliver ES

Tissue engineering attempts to imitate the structure and function of the tissues or organs through the use of engineered scaffolds, which optimize the response of cell-biomaterials and mimic the native environment [43]. ES has received considerable attention to influence cellular or tissue behaviors as stimulation factor of tissue engineering [44]. As a bridge to deliver ES, scaffolds need to meet requirements: excellent biocompatibility, prominent electrochemical performance, and no byproduct generation [45].

Materials to deliver electrical stimulation

To provide ES for tissue regeneration or functional recovery, a number of materials have been developed to make scaffolds. To date, metallic biomaterials include platinum, and gold is increasingly used in medical applications due to their high mechanical strength, long-term stability, good conductivity, and biocompatibility [46]. However, except noble metals, most metallic materials are easily oxidized, and present weak corrosion resistance [47, 48], the release of metal ion may also cause allergies or carcinogenesis. Surface modification of metallic materials is considered to be an effective method to solve the above problems, include preparation of coatings, and covalent chemical conjugation of bioactive molecules [48]. Conducting polymers, such as poly (3,4-ethylene dioxythiophene), polyaniline, polypyrrole (PPy) have been studied as possible candidates [49]. Although the cracks or delamination of conducting polymers under long-term stimulation restrict the electrode performance, cross-link with a specific agent or in-situ polymerization of the conjugated polymer improving the physical stability while allowing the exploration of their superior properties [50, 51]. Carbon materials such as graphene, carbon nanotubes and carbon aerogels have excellent electrical properties and the ability to be easily bio-functionalized, as well as drug loadings [52,53,54]. However, the biocompatibility of carbon relative materials is still a significant challenge. Scientists have modified carbon relevant materials to eliminate defects. Depositing and fixing graphene can avoid the damage to cells due to direct contact interaction and wrapping mechanisms [55]. The direct exposure of the carbon nanotubes inside the body presents a high probability of detachment. The above issue was overcome by reinforcing the carbon nanotube patterns inside the biocompatible soft polymer matrix [56]. Table 1 lists the forms of different materials, the electrical conductivity and the advantages and disadvantages.

Table 1 The forms, electrical conductivity, advantages and disadvantages of different materials

Methods to deliver electrical stimulation and its parameters

The methods to deliver ES can mainly divide into three types: direct coupling, capacitive coupling, and using an electromagnetic field.

Direct coupling shows in Fig. 1a. The electrodes are inserted directly into the culture medium and attached to the scaffold to deliver ES. This method is most widely used because of its easy operation. However, drawbacks are apparent such as insufficient biocompatibility of the electrode, contact with the medium lead to temperature rise, pH changes, and the generation of harmful byproducts [59]. Prabhakaran et al. [19] demonstrated that ES applied to the NSCs cultured on the electrospun Poly-L-lactide/polyaniline fibers scaffold exhibited extended neurite outgrowth. Briefly, a silver electrode and a platinum electrode were inserted into the opposite ends of the nanofiber scaffold placed in the medium and connected to constant unipolar trapezoidal pulses. The average length of neurite extended from NSCs with stimulation were nearly 100% higher than the average length of those without ES. In other circumstances, the scaffold is used as substrate-cathode. Stewart et al. [60] induced human NSCs differentiation with a two-electrode device, placing a platinum mesh electrode on top, and using a PPy-coated Au-mylar surface as a working electrode. ES on PPy induced hNSC to differentiate into neurons expressing β-III Tubulin (Tuj1), and a small number of glial cells expressing glial fibrillary acidic proteins, accompanied with longer neurites and significant branches.

Fig. 1
figure 1

Three ways to deliver electrical stimulation: (a) direct coupling, (b) capacitive coupling, and (c) inductive Reprinted with permission from reference [40]

Capacitive coupling is more biologically safe compared with direct coupling. As shown in Fig. 1b, two electrodes are placed at opposite ends to provide a uniform electric field to the cells seeded on the scaffold which locates between the electrodes. This system is non-invasive and does not require a conductive scaffold to provide uniform ES. Vaca-Gonzalez et al. [61] used a capacitive coupling system to provide chondrocytes with a 4 mV/cm ES, which significantly increased the cell proliferation. When applied an 8 mV/cm ES, the glycosaminoglycan secreted by chondrocytes remained stable, which is essential to maintain tissue moisture, lubrication, and protection of articular surfaces.

The third type is inductive coupling, as shown in Fig. 1c. Inductive coupling usually uses a controllable electromagnetic field generated by a conductive coil placed around the cell culture system, called pulsed electromagnetic field stimulation (PEMF). The stimulus is transmitted by the pulse to mimic the natural potential transfer in the human body [40]. PEMF provide potential near the target cell, rather than directly apply ES to the cells. The main drawback of PEMF treatment is taking time and resource consumption. For example, some therapies need to last 10 h per day [62] or quite high voltage. Hess et al. [63] created a device using ferromagnetic core covered by the first current-carrying coil, and the culture chamber with connected silicone tube as the second coil. When applied an alternating voltage to the first coil, the second coil only influenced by the electrical potential without any interfering magnetic field or biochemical reactions. The combination of inductive coupling and high sulfated hyaluronan derivative substrate can synergistically stimulate BMSCs, the bone-related proteins (RUNX-2, ALP, OPN) in BMSCs were significantly improved then cells show an osteogenic differentiation. Table 2 lists the advantages and disadvantages of each methods mentioned above.

Table 2 The advantages and disadvantages of three methods of providing ES

The different parameters of ES are quite considerable for regulating cell behavior. ES can provide in the form of monophasic and biphasic, where the waveform has pulse, sinusoidal, square, triangular, and sawtooth pattern. The use of intermittent or continuous stimulation is another parameter. According to the stimulation parameters setup, monophasic stimulation is effective in polarizing the target tissue [65] but may produce reactive oxygen species through the oxidation-reduction process due to the Faradaic reaction at the surface of a metal electrode [66]. Especially in the case of large current pulses delivered at long durations and/or high frequencies, cell damage may result from the Joule heating effect [67]. However, the impact of localized ionic or electrochemical imbalance caused by the small Faradaic impact on the experiment cannot be ignored [66]. On the contrary, biphasic stimulation may be more advantageous since it prevents the charge accumulation, generates lower levels of electrolysis products at the electrodes, outside of the limitations of monophasic stimulation and can be applied for more extended periods and at higher voltages [68,69,70]. For instance, biphasic stimulation is usually selected in clinical application to stimulate neural tissue, because of the few charge accumulation and toxic by-products and is less likely to cause neuronal loss [69, 71]. Monophasic stimulation can be used for short-term experiments, but long-term applications require biphasic stimulation [72, 73]. Improper stimulation parameters can lead to results that are contrary to experimental expectations. The cell migration rate is positively correlated with the intensity of ES, and is discussed in next section. For stimulus frequency, ES frequency below 1 kHz augment cell proliferation through significantly affected the cell cycle, increasing the proportion of cells and synthesizing DNA [74]. Experiments show that ES frequency above 1 kHz can induce cell differentiation, but must maintain low intensity [75, 76]. High intensity above 100 V/cm can cause cell membrane electroporation, an immediate increase in intracellular Ca2+ and reactive oxygen species, then induce cell apoptosis [24, 77]. When high intensity is unavoidable, pulse electrical stimulation should be used to avoid continuous high intensity to reduce damage.

The effects of electrical stimulation on cell alignment and migration

By participating in tissue formation, tissue regeneration, and wound healing, directed cell migration and alignment can be very beneficial for regenerative medicine [78,79,80,81]. The ability to regulate cell migration and alignment would be an invaluable asset for regenerative medicine. The mechanisms underneath the cell alignment and migration are thought to be responsible for these effects, including voltage-gated ion channels, G-protein coupling receptors, integrins, cell polarization, and endogenous electric fields [82]. In literature, ES as the physical methods has attracted much attention because it can activate specific signaling pathways in cells near the cathode or anode and induce cell migration and alignment [83].

The effects of electrical stimulation on cell alignment

Numerous methods are capable of inducing cell migration and alignment by external factors, such as the addition of bioactive factors, providing individual surface patterns, and studies have shown that physical methods ES over all other techniques to guide cell migration and alignment [17, 41, 84].

ES has a significant effect on cell alignment and redirect random cells to be aligned, the direction of cell alignment changes gradually as the direction of the ES changes [84,85,86]. Some types of cell are aligned perpendicular to the direction of the electric field vectors to minimize the field gradient across the cell, such as cardiac adipose tissue-derived progenitor cells [18], endothelial progenitor cells [87], vascular ECs [77], BMSCs [88], adipose-derived stromal cells [89], etc. At the same time, some cells are aligned parallel to the field vectors due to the ES causes rearrangement of the cell cytoskeleton, such as ventricular myocytes [20], cardiomyocytes [86], myoblasts [85], PC-12 cells [90], obsteoblasts [84], etc. The ES intensity usually < 10 V/cm, and the cells aligned better with the ES intensity increased, but the cell activity is relatively decreased [77, 87, 91].

In the presence of 100 mA monophasic ES, the osteoblasts elongation and proliferation were mainly reliant on the ES, whereas the topographical features played a minor role [84]. In our present study [90], perpendicular electrical field vectors from the pattern may reorientate the PC-12 cell alignment, while parallel vectors could further increase neuritis extension compared with perpendicular vectors. On the contrary, some types of cell are aligned perpendicular to the vectors under ES. The murine adipose-derived stromal cells exhibited a constant perpendicular alignment with the field vectors [89]. BMSCs consistent perpendicular alignment to the field vectors [88]. The long axis of cobblestone-like ECs transformed into highly ordered, perpendicular to the field vectors with applied ES, then the cell morphology presents similarly to the inner surface topography of the blood vessel, which shows excellent potential in angiogenesis [77].

The effects of electrical stimulation on cell migration

As well as cell alignment, ES plays a significant role in cell migration. The directional migration of cells in response to ES is called electrotaxis [77]. The guiding effect of ES on cell migration also differs depending on the cell type. Among different types of cell, NSCs [92], macrophages [93], mouse neural precursor cells (NPCs) [94], osteoblasts [95] and endothelial progenitor cells [87] toward the cathode [20, 89, 92, 96], BMSCs [97], human dermal fibroblasts [98] and SCs [99, 100] toward the anode. Reversal of the ES polarity reversed the migration direction of cells [87, 92]. The ES intensity can stimulate cell migration from a minimum of 0.1 V/cm to a maximum of 12 V/cm and did not cause any significant damage to cells, did not affect cell phenotype or the differentiation potential [101,102,103]. At the same time, cells showed gradually increased migration rate and distance with higher ES strength [101, 104, 105].

As NPCs shows a cathodal migration, Liu et al. [103] demonstrated that reversing the ES direction resulted in a reversed direction of NPCs migration. This behavior is beneficial to long-distance migration of neural cells for nerve repair. ES also directs some types of cell migrate to the anode, such as BMSCs [97, 106], human dermal fibroblasts [37] and SCs [99, 100, 107]. Forciniti et al. [100] indicated an increase in average displacement of the SCs to anodal with 0.5 V to 1 V ES, which is of great value in optimizing conductive polymers for different biomedical applications such as nerve repair. Moreover, even different cell lines of the same specie have different responses to ES in the migration. Li et al. [108] demonstrated different cell lines of non-small cell lung cancer cells (H460, and H1299) showed great different in migration direction, while H460 migrated toward cathode, H1299 exhibited anodal migration. This discovery has led to a better understanding of the mechanism of cancer, which is beneficial for cancer therapy.

mASCs exhibited a positive correlation between cell migration rate and ES intensity (from 1 V/cm to 10 V/cm) [89]. The same performance showed in human BMSCs, as the ES intensity is above the physiological level, the migration rate increases while physiological levels of ES do not affect cell motility, this behavior may be essential for maintaining transplanted cells at the lesion site [88, 105]. Different research groups have conflict results in the same cell type. Forciniti et al. [100] showed ES conditions (from 0 to 0.5 V) produced a significant increase in SCs migration rate but did not affect migration direction. Different from others, Li et al. [99] demonstrated the ES did not significantly affect the SCs migration speed, but as the ES intensity increases from 50 mV/mm to 200 mV/mm, the orientation and displacement of the anode migration increase significantly.

Mechanism of ES on cell migration

The specific mechanism leading to electrotaxis is still unclear. Many factors may be included, such as endogenous microenvironment, ion channels, membrane receptors, transport proteins, and competing signal pathways such as Wnt/GSK3β and TGFβ1/ERK/NF-κB [109,110,111]. According to the literature, the main mechanism involved in ES induced cell migration is shown in Fig. 2.

Fig. 2
figure 2

Mechanism of ES-induced cell migration

The importance of epidermal growth factor receptors (EGFR) in ES induced cell migration has attracted the attention of numerous scientists [77, 87, 112]. ES regulates downstream growth factors and cytokine production through a feedback mechanism by EGFR. Asymmetric expression of EGFR in front of and behind cells activates signaling pathways and leads to cell migration [112, 113]. Pilipos et al. [114] demonstrated that blockade of EGFR via erlotinib significantly attenuate NPCs electrotaxis. EGFR activation stimulates several signaling pathways, including MAPK-ERK1/2 and PI3K/Akt pathway. MAPK-ERK1/2 are generally involved in signal transduction from extracellular sources in different signaling pathways. Downstream protein ERK1 and ERK2 are fully activated by phosphorylation of MEK, which is involved in cell migration. Hammerick et al. [89] proved that inhibiting the MAPK pathway leading to migration inhibited suggests the participation of this pathway in mASCs electrotaxis. It has been reported that static monophasic ES can regulate epithelial cell proliferation and migration by activating the MAPK signaling cascade ERK1/2 [115].

The other and the most critical pathway is the PI3K/Akt signaling pathway, which has been mostly explored in the cellular response to ES. ES significantly increase the expression of downstream protein PIP3 and phosphorylation of Akt, at the same time, induce the asymmetric distribution of PIP3 and cytoskeletal proteins migrate toward cathode [113, 116]. Meng et al. [116] exhibited the cathode migration of NPCs under ES needed the activation of the PI3K signal pathway. Pharmacological inhibition or genetic disruption of PI3K/Akt pathway will inhibit the electrotaxis, then applied ES increases Akt phosphorylation and PIP3 fluorescence, demonstrate the importance of the PI3K/Akt pathway in ES driven directed migration of NPCs. PTEN is a phosphatase that inhibits PI3K signaling transduction including Akt. Zhao et al. [117] demonstrated genetic disruption of PTEN within keratinocyte enhanced ES induced ERK and Akt phosphorylation, cell performed a significantly higher directionality compared with controls. This result supports the antagonistic relationship between PI3K and PTEN.

In addition to signal pathways, ion channels such as voltage-gated Ca2+ channels is a vital part of membrane polarization and cell response during ES [118]. All living cells have a transmembrane potential. The current induces a flow of ions (Na+, Cl, K+, Ca2+, etc.) by ion channels and transporters [16]. In response to ES, intracellular molecular and transport channel polarization, then ion flow takes place and trigger cytoskeleton changes to direct cell migration, in consequence, the calcium influx contribute to persistent cell migration towards cathode [101]. In the review of Balint et al. [40], the ES effect will be impaired or completely blocked if block the calcium channels, intracellular store and the calmodulin by verapamil and nifedipine, TMB-8, W-7, respectively. Zhao et al. [102] reported that a 115 V/m monophasic ES could cause a calcium influx mediated NPCs mobility improved and cathodal migration.

The effects of electrical stimulation on cell proliferation and differentiation

In addition to the impact on the migration, ES also plays a significant role in influencing proliferation and guiding differentiation [50, 119, 120]. A major challenge in regenerative medicine is to compensate cells that lost as a result of injury or disease [15]. For instance, after excessive cell loss, most tissues produce scars, such as fibrous collagen scars in the heart, causing an ischemic environment and limiting oxygen delivery [121]. However, the inevitable problems are the difficulty in harvesting sufficient cells for implantation. In these circumstances, the capabilities of proliferation and multilineage differentiation of seed cells have attracted much attention [122]. Table 1 summarizes the specific conditions of ES to induce cell proliferation and differentiation.

The effects of electrical stimulation on cell proliferation

Proper ES can promote cell proliferation, usually under continuous stimulation of < 1 V/cm [37, 95, 120]. Within the ES intensity range the cell proliferation rate increases with increasing intensity [95]. Preosteoblasts [123], obsteoblasts [37], unrestricted somatic human stem cells [124], human umbilical vein ECs [125], NSCs [126], human dermal fibroblasts [127] exhibited 0.2 to 1.5 times proliferation, with increasing cellular metabolic activity, and do not affect cell phenotype [123]. High-intensity ES of > 100 V/cm is also favorable for cell proliferation in a short period (< 1 ms) single stimulation, but excessively high intensity leads to cell death [128].

Shao et al. [17] demonstrated that in the presence of 100 mA direct current (DC), the viability of osteoblasts did not affect, and the degree of proliferation increased significantly compared with no stimuli group and other current stimuli groups. This finding indicates that proper ES can be used for bone regeneration and fracture healing. Zhu et al. [126] revealed that compared with the control group, NSCs proliferation was increased 35% with ES, which condition is 100 μA and pulse rates of 100 Hz with 100 μsec duration for 24 h. A possible cause is that ES upregulates the proliferation of cell nuclear antigen, which is a DNA polymerase-associated protein, and activate extracellular signal-regulated kinases 1 and 2 pathways involved in transduction of proliferative signals. Meanwhile, the differentiation potential of NSCs has not been affected by the promotion of proliferation. In the study of Sun et al. [129] SCs underwent a remarkable degree of proliferation under ES (100 mV/cm, 1 h per day) than without ES, especially in day 5. It might because ES promoted SC myelin gene expression and neurotrophin secretion, then led to SCs proliferation. This result indicated the promising potential of ES on peripheral nerve repair and regeneration. Kapeller et al. [119] pointed out that with 1 μA DC stimuli, cardiomyocytes showed better proliferation behavior with no morphological changes in vitro. At the same time, ES also affects matrix metalloproteinases (MMP) in cardiomyocytes, which involved to keep the balance between matrix synthesis and degradation. It shows that ES has a certain positive effect on heart repair.

The effects of electrical stimulation on cell differentiation

Cell differentiation based therapy provides a promising approach for regeneration medicine. Short-term (several minutes), low-intensity (0.06~6 V/cm) ES can promote cardiac differentiation of human induced pluripotent stem cells, muscle precursor cells, with the formation of embryoid bodies [130, 131]. Neural progenitor cells, neural precursor cells, and NSCs can differentiate into neurons instead of glial cells under ES [102, 106, 132]. The stimulation intensity is basically < 2 V/cm and sustain more than 7 days. The differential medium can appropriately added with FBS, retinoic acid, and nerve growth factor. ES can induce bone marrow stromal cells, BMSCs, MC3T3-E1 cells osteogenic differentiation instead of cartilage, stimulation intensity should be < 2 V/cm and sustain 14~28 days, the medium need to add dexamethasone in most cases [6, 63, 120, 133,134,135]. Therefore, the application of ES could provide a valid approach to induce cell differentiation in tissue engineering.

Hernández et al. [131] demonstrated human induced pluripotent stem cells (Foreskin)-2 cell line exhibit cardiac differentiation under ES (65 mV/mm or 200 mV/mm, biphasic current pulses), by significantly increased the expression of cardiac transcription factors nkx2.5 and tbx5, as well as the cardiac contractile muscle proteins ACTC1, TNNT2, MYH7, and MYL7, compared with the unstimulated control group. In the study of Chan et al. [130], human ESCs cultured in the presence of ES (6.6 V/cm, 1 Hz, and 2 ms pulses) showed an increase in the proportion of ventricular-like cardiomyocytes, by significant upregulation of cardiac-specific gene expression including hcn1, mlc2v, scn5a, serca, kv4.3, and gata4. Matsumoto et al. [135] induced the differentiation of mouse bone marrow stromal cells into neural cells using ES (10 Hz, 100 mV, 2.0 ms, 30 min, rectangular pulse). Neurogenin2 was detected as increased expression through the β-catenin signaling pathway after ES, which is involved in neural differentiation and inhibits astrocytic differentiation. Sun et al. [129] demonstrated ES (100 mV/cm, 1 h) could induce PC-12 cell differentiation into SCs and synaptic elongation even without nerve growth factor treatment, which is a kind of nerve growth regulator that promotes both neuron nutrition and neurite outgrowth.

In the study of Wang et al. [120] ES (200 mV/cm, 1 Hz to 100 kHz, 30 min) promote MC3T3-E1 cells osteogenic differentiation via examined alkaline phosphatase activity. During stimulation, 100 Hz could up-regulate the mRNA levels of collagen I, collagen II, and RUNX2, which are osteosis-related genes. On the other hand, 1 Hz to 10 Hz could improve calcium deposition in the intracellular matrix, which contribute to treat the bone fracture and bone nonunion. Hronik-Tupaj et al. [6] demonstrated ES (20 mV/cm, 60 kHz) improved hMSC osteogenic differentiation potential based on calcium deposition, because of the two-fold increase of alkaline phosphatase and collagen type 1.

Table 3 The specific conditions of ES to induce cell proliferation and differentiation.

Table 3 The specific conditions of ES to induce cell proliferation and differentiation

Summary

In this review, we have highlighted the effects of ES as a physical stimulator on cellular behavior for the purposes of applying to regenerative medicine and tissue engineering. In most cases, ES facilitates cell proliferation and differentiation, enhance cell cathode migration and alignment to field vectors, and mainly through EGFR, PI3K and Ca2+ related mechanism. ES can be delivered through tissue engineering scaffolds made of metallic biomaterials, conducting polymers, or carbon materials, the main methods are direct coupling, capacitive coupling, and using an electromagnetic field.

In the future, the combination of specific material/structure and ES will offer many advantages over other types of stimulations and allow for precise cellular regulation. Developing safe and effective partition-type scaffold combine with ES that can distinguish different areas to perform different stimulation still requires some challenges to be overcome. Although still in its early stages, the field of ES is rapidly evolving, and new next-generation regenerative medicine and tissue engineering will make it possible to take advantage of ES. Studying the low risk ES method for wearable application is also a future direction. With increasing research, electrical stimuli have the potential to play a significant role in tissue engineering and regenerative medicine.

Availability of data and materials

Not applicable.

References

  1. Tatullo M, Marrelli M, Paduano F. The regenerative medicine in oral and maxillofacial surgery: the most important innovations in the clinical application of mesenchymal stem cells. Int J Med Sci. 2015;2(1):72–7.

    Article  CAS  Google Scholar 

  2. Mason C, Dunnill P. A brief definition of regenerative medicine. Regen Med. 2008;3(1):1–5.

    Article  Google Scholar 

  3. Bhardwaj N, Devi D, Mandal BB. Tissue-engineered cartilage: the crossroads of biomaterials, cells and stimulating factors. Macromol Biosci. 2015;15(2):153–82.

    Article  CAS  Google Scholar 

  4. Gupta P, Adhikary M, Kumar M, Bhardwaj N, Mandal BB. Biomimetic, osteoconductive non-mulberry silk fiber reinforced tricomposite scaffolds for bone tissue engineering. ACS Appl Mat Interfaces. 2016;8(45):30797–810.

    Article  CAS  Google Scholar 

  5. Mondschein RJ, Kanitkar A, Williams CB, Verbridge SS, Long TE. Polymer structure-property requirements for stereolithographic 3d printing of soft tissue engineering scaffolds. Biomaterials. 2017;140:170–88.

    Article  CAS  Google Scholar 

  6. Hronik-Tupaj M, Rice WL, Cronin-Golomb M, Kaplan DL, Georgakoudi I. Osteoblastic differentiation and stress response of human mesenchymal stem cells exposed to alternating current electric fields. Biomed Eng Online. 2011;10(1):9.

    Article  Google Scholar 

  7. Li C, Wei G, Gu Q, Wen G, Qi B, Xu L, et al. Donor age and cell passage affect osteogenic ability of rat bone marrow mesenchymal stem cells. Cell Biochem Biophys. 2015;72(2):543–9.

    Article  CAS  Google Scholar 

  8. Guillot PV, Cui WN, Polak DJ. Stem cell differentiation and expansion for clinical applications of tissue engineering. J Cell Mol Med. 2010;11(5):935–44.

    Article  CAS  Google Scholar 

  9. Breuls RG, Jiya TU, Smit TH. Scaffold stiffness influences cell behavior: opportunities for skeletal tissue engineering. Open Orthop J. 2008;2(2):103–9.

    Article  Google Scholar 

  10. Bianco P, Robey PG. Stem cells in tissue engineering. Nature. 2014;414(6859):639–53.

    Google Scholar 

  11. Ma F, Xiao Z, Chen B, Hou X, Han J, Zhao Y, et al. Accelerating proliferation of neural stem/progenitor cells in collagen sponges immobilized with engineered basic fibroblast growth factor for nervous system tissue engineering. Biomacromolecules. 2014;15(3):1062–8.

    Article  CAS  Google Scholar 

  12. Chen G, Qi Y, Niu L, Di T, Zhong J, Fang T, et al. Application of the cell sheet technique in tissue engineering. Biomed Rep. 2015;3(6):749–57.

    Article  Google Scholar 

  13. Min Z. Electrical fields in wound healing-an overriding signal that directs cell migration. Semin Cell Dev Biol. 2009;20(6):674–82.

    Article  CAS  Google Scholar 

  14. Li W, Guan T, Zhang X, Wang Z, Wang M, Zhong W, et al. The effect of layer-by-layer assembly coating on the proliferation and differentiation of neural stem cells. ACS Appl Mater Interfaces. 2015;7(5):3018–29.

    Article  CAS  Google Scholar 

  15. Kobolak J, Dinnyes A, Memic A, Khademhosseini A, Mobasheri A. Mesenchymal stem cells: identification, phenotypic characterization, biological properties and potential for regenerative medicine through biomaterial micro-engineering of their niche. Methods. 2016;99:62–8.

    Article  CAS  Google Scholar 

  16. Love MR, Palee S, Chattipakorn SC, Chattipakorn N. Effects of electrical stimulation on cell proliferation and apoptosis. J Cell Physiol. 2018;233(3):1860–76.

    Article  CAS  Google Scholar 

  17. Ding S, Kingshott P, Thissen H, Pera M, Wang PY. Modulation of human mesenchymal and pluripotent stem cell behavior using biophysical and biochemical cues: a review. Biotechnol Bioeng. 2017;114(2):260–80.

    Article  CAS  Google Scholar 

  18. Llucià-Valldeperas A, Sanchez B, Soler-Botija C, Gálvez-Montón C, Prat-Vidal C, Roura S, et al. Electrical stimulation of cardiac adipose tissue-derived progenitor cells modulates cell phenotype and genetic machinery. J Tissue Eng Regen Med. 2015;9(11):E76–83.

    Article  CAS  Google Scholar 

  19. Prabhakaran MP, Ghasemimobarakeh L, Jin G, Ramakrishna S. Electrospun conducting polymer nanofibers and electrical stimulation of nerve stem cells. J Biosci Bioeng. 2011;112(5):501–7.

    Article  CAS  Google Scholar 

  20. Radisic M, Park H, Shing H, Consi T, Schoen FJ, Langer R, et al. Functional assembly of engineered myocardium by electrical stimulation of cardiac myocytes cultured on scaffolds. Proc Natl Acad Sci U S A. 2004;101(52):18129–34.

    Article  CAS  Google Scholar 

  21. Frelinger AL 3rd, Torres AS, Caiafa A, Morton CA, Berny-Lang MA, Gerrits AJ, et al. Platelet-rich plasma stimulated by pulse electric fields: platelet activation, procoagulant markers, growth factor release and cell proliferation. Platelets. 2016;27(2):128–35.

    CAS  Google Scholar 

  22. Jeon O, Lee K, Alsberg E. Spatial micropatterning of growth factors in 3d hydrogels for location-specific regulation of cellular behaviors. Small. 2018;14(25):1800579.

    Article  CAS  Google Scholar 

  23. Wang Z, Qing Q, Chen X, Liu CJ, Luo JC, Hu JL, et al. Effects of scaffold surface morphology on cell adhesion and survival rate in vitreous cryopreservation of tenocyte-scaffold constructs. Appl Surf Sci. 2016;388:223–7.

    Article  CAS  Google Scholar 

  24. Nuccitelli R, Lui K, Kreis M, Athos B, Nuccitelli P. Nanosecond pulsed electric field stimulation of reactive oxygen species in human pancreatic cancer cells is ca (2+)-dependent. Biochem Biophys Res Commun. 2013;435(4):580–5.

    Article  CAS  Google Scholar 

  25. Ji WC, Zhang XW, Qiu YS. Selected suitable seed cell, scaffold and growth factor could maximize the repair effect using tissue engineering method in spinal cord injury. World J of Exp Med. 2016;6(3):58–62.

    Article  Google Scholar 

  26. Choi JS, Yoon HI, Lee KS, Choi YC, Yang SH, Kim I-S, Cho YW. Exosomes from differentiating human skeletal muscle cells trigger myogenesis of stem cells and provide biochemical cues for skeletal muscle regeneration. J Control Release. 2016;222:107–15.

    Article  CAS  Google Scholar 

  27. Prabhakaran MP, Venugopal J, Chan CK, Ramakrishna S. Surface modified electrospun nanofibrous scaffolds for nerve tissue engineering. Nanotechnology. 2008;19(45):455102.

    Article  CAS  Google Scholar 

  28. Shori D, Pandey S, Kubde R, Rathod Y, Atara R, Rathi S. To evaluate and compare the effect of different post surface treatments on the tensile bond strength between fiber posts and composite resin. J Int Oral Health. 2013;5(5):27–32.

    Google Scholar 

  29. Teker D, Muhaffel F, Menekse M, Karaguler NG, Baydogan M, Cimenoglu H. Characteristics of multi-layer coating formed on commercially pure titanium for biomedical applications. Mater Sci Eng C Mater Biol Appl. 2015;48:579–85.

    Article  CAS  Google Scholar 

  30. Morgan JT, Wood JA, Shah NM, Hughbanks ML, Russell P, Barakat AI, et al. Integration of basal topographic cues and apical shear stress in vascular endothelial cells. Biomaterials. 2012;33(16):4126–35.

    Article  CAS  Google Scholar 

  31. Yang Y, Wang K, Gu X, Leong KW. Biophysical regulation of cell behavior—cross talk between substrate stiffness and nanotopography. Engineering (Beijing). 2017;3(1):36–54.

    CAS  Google Scholar 

  32. Zhang SX, Liu L, Zhao W. Targeting biophysical cues: a niche approach to study, diagnose, and treat cancer. Trends Cancer. 2018;4(4):268–71.

    Article  Google Scholar 

  33. Gao J, Raghunathan VK, Reid B, Wei D, Diaz RC, Russell P, et al. Biomimetic stochastic topography and electric fields synergistically enhance directional migration of corneal epithelial cells in a mmp-3-dependent manner. Acta Biomater. 2015;12:102–12.

    Article  CAS  Google Scholar 

  34. Huang X, Das R, Patel A, Nguyen TD. Physical stimulations for bone and cartilage regeneration. Regen Eng Transl Med. 2018;4(4):216–37.

    Article  Google Scholar 

  35. Hernandez-Bule ML, Trillo MA, Ubeda A. Molecular mechanisms underlying antiproliferative and differentiating responses of hepatocarcinoma cells to subthermal electric stimulation. PLoS One. 2014;9(1):e84636.

    Article  CAS  Google Scholar 

  36. Jaatinen L, Young E, Hyttinen J, Vörös J, Zambelli T, Demkó L. Quantifying the effect of electric current on cell adhesion studied by single-cell force spectroscopy. Biointerphases. 2016;11(1):011004.

    Article  CAS  Google Scholar 

  37. Kumar A, Nune KC, Misra R. Understanding the response of pulsed electric field on osteoblast functions in three-dimensional mesh structures. J Biomater Appl. 2016;31(4):594–605.

    Article  CAS  Google Scholar 

  38. Li N, Zhang Q, Gao S, Song Q, Huang R, Wang L, et al. Three-dimensional graphene foam as a biocompatible and conductive scaffold for neural stem cells. Sci Rep. 2013;3:1064.

    Article  CAS  Google Scholar 

  39. Rando TA, Ambrosio F. Regenerative rehabilitation: applied biophysics meets stem cell therapeutics. Cell Stem Cell. 2018;22(3):306–9.

    Article  CAS  Google Scholar 

  40. Balint R, Cassidy NJ, Cartmell SH. Electrical stimulation: a novel tool for tissue engineering. Tissue Eng Part B Rev. 2013;19(1):48–57.

    Article  CAS  Google Scholar 

  41. Wang ET, Zhao M. Regulation of tissue repair and regeneration by electric fields. Chin J Traumatol. 2010;13(1):55–61.

    Google Scholar 

  42. Chen C, Kong XD, Lee IS. Modification of surface/neuron interfaces for neural cell-type specific responses: a review. Biomed Mater. 2015;11(1):014108.

    Article  CAS  Google Scholar 

  43. Deepthi S, Jeevitha K, Sundaram MN, Chennazhi K, Jayakumar R. Chitosan–hyaluronic acid hydrogel coated poly (caprolactone) multiscale bilayer scaffold for ligament regeneration. Chem Eng J. 2015;260:478–85.

    Article  CAS  Google Scholar 

  44. Wang J, Tian L, Chen N, Ramakrishna S, Mo X. The cellular response of nerve cells on poly-l-lysine coated plga-mwcnts aligned nanofibers under electrical stimulation. Mater Sci Eng C Mater Biol Appl. 2018;91:715–26.

    Article  CAS  Google Scholar 

  45. Li J, Hu H, Li H, Yao C. Recent developments in electrochemical sensors based on nanomaterials for determining glucose and its byproduct H2O2. J Mater Sci. 2017;52(17):10455–69.

    Article  CAS  Google Scholar 

  46. Colic M, Stamenkovic D, Anzel I, Lojen G, Rudolf R. The influence of the microstructure of high noble gold-platinum dental alloys on their corrosion and biocompatibility in vitro. Gold Bull. 2009;42(1):34–47.

    Article  CAS  Google Scholar 

  47. Nazeer MA, Yilgor E, Yilgor I. Intercalated chitosan/hydroxyapatite nanocomposites: promising materials for bone tissue engineering applications. Carbohydr Polym. 2017;175:38–46.

    Article  CAS  Google Scholar 

  48. Hu GQ, Guan K, Lu LB, Zhang JR, Lu N, Guan YC. Engineered functional surfaces by laser microprocessing for biomedical applications. Engineering. 2018;4(6):822–30.

    Article  CAS  Google Scholar 

  49. Ramasamy MS, Nikolakapoulou A, Raptis D, Dracopoulos V, Paterakis G, Lianos P. Reduced graphene oxide/polypyrrole/pedot composite films as efficient pt-free counter electrode for dye-sensitized solar cells. Electrochimi Acta. 2015;173:276–81.

    Article  CAS  Google Scholar 

  50. Pires F, Ferreira Q, Rodrigues CAV, Morgado J, Ferreira FC. Neural stem cell differentiation by electrical stimulation using a cross-linked pedot substrate: expanding the use of biocompatible conjugated conductive polymers for neural tissue engineering. Biochim Biophys Acta. 2015;1850(6):1158–68.

    Article  CAS  Google Scholar 

  51. DeVolder RJ, Seo Y, Kong H. Proangiogenic alginate-g-pyrrole hydrogel with decoupled control of mechanical rigidity and electrically conductivity. Biomater Res. 2017;21(1):24.

    Article  CAS  Google Scholar 

  52. Zhang LL, Zhao X. Carbon-based materials as supercapacitor electrodes. Chem Soc Rev. 2009;38(9):2520–31.

    Article  CAS  Google Scholar 

  53. Lee JR, Ryu S, Kim S, Kim BS. Behaviors of stem cells on carbon nanotube. Biomater Res. 2015;19(1):3.

    Article  CAS  Google Scholar 

  54. Kim J, Kim HD, Park J, Lee ES, Kim E, Lee SS, et al. Enhanced osteogenic commitment of murine mesenchymal stem cells on graphene oxide substrate. Biomater Res. 2018;22(1):1.

    Article  CAS  Google Scholar 

  55. Akhavan O. Graphene scaffolds in progressive nanotechnology/stem cell-based tissue engineering of the nervous system. J Mater Chem B. 2016;4(19):3169–90.

    Article  CAS  Google Scholar 

  56. Gupta P, Lahiri D. Aligned carbon nanotube containing scaffolds for neural tissue regeneration. Neural Regen Res. 2016;11(7):1062–3.

    Article  Google Scholar 

  57. Ghasemi-Mobarakeh L, Prabhakaran MP, Morshed M, Nasr-Esfahani MH, Baharvand H, Kiani S, et al. Application of conductive polymers, scaffolds and electrical stimulation for nerve tissue engineering. J Tissue Eng Regen Med. 2011;5(4):e17–35.

    Article  CAS  Google Scholar 

  58. Liu S, Wang J, Zhang D, Zhang P, Ou J, Liu B, et al. Investigation on cell biocompatible behaviors of polyaniline film fabricated via electroless surface polymerization. Appl Surf Sci. 2010;256(11):3427–31.

    Article  CAS  Google Scholar 

  59. Samadian H, Zakariaee SS, Adabi M, Mobasheri H, Azami M, Faridi-Majidi R. Effective parameters on conductivity of mineralized carbon nanofibers: an investigation using artificial neural networks. RSC Adv. 2016;6(113):11908–18.

    Article  CAS  Google Scholar 

  60. Stewart E, Kobayashi NR, Higgins MJ, Quigley AF, Jamali S, Moulton SE, et al. Electrical stimulation using conductive polymer polypyrrole promotes differentiation of human neural stem cells: a biocompatible platform for translational neural tissue engineering. Tissue Eng Part C Methods. 2015;21(4):385–93.

    Article  CAS  Google Scholar 

  61. Vaca-Gonzalez JJ, Guevara JM, Vega JF, Garzon-Alvarado DA. An in vitro chondrocyte electrical stimulation framework: a methodology to calculate electric fields and modulate proliferation, cell death and glycosaminoglycan synthesis. Cell Mol Bioeng. 2016;9(1):116–26.

    Article  CAS  Google Scholar 

  62. Fu YC, Lin CC, Chang JK, Chen CH, Tai IC, Wang GJ, et al. A novel single pulsed electromagnetic field stimulates osteogenesis of bone marrow mesenchymal stem cells and bone repair. PLoS One. 2014;9(3):e91581.

    Article  CAS  Google Scholar 

  63. Hess R, Jaeschke A, Neubert H, Hintze V, Moeller S, Schnabelrauch M, et al. Synergistic effect of defined artificial extracellular matrices and pulsed electric fields on osteogenic differentiation of human mscs. Biomaterials. 2012;33(35):8975–85.

    Article  CAS  Google Scholar 

  64. Meng S, Rouabhia M, Zhang Z, De D, De F, Laval U. Electrical stimulation in tissue regeneration. Appl Biomed Eng. 2011;23:37–62.

    Google Scholar 

  65. Sommer M, Kamm T, Tergau F, Ulm G, Paulus W. Repetitive paired-pulse transcranial magnetic stimulation affects corticospinal excitability and finger tapping in parkinson's disease. Clin Neurophysiol. 2002;113(6):944–50.

    Article  Google Scholar 

  66. Waataja JJ, Tweden KS, Honda CN. Effects of high-frequency alternating current on axonal conduction through the vagus nerve. J Neural Eng. 2011;8(5):056013.

    Article  Google Scholar 

  67. Kim J, Yang HJ, Cho TH, Lee SE, Yong DP, Kim HM, et al. Enhanced regeneration of rabbit mandibular defects through a combined treatment of electrical stimulation and rhbmp-2 application. Med Biol Eng Comput. 2013;51(12):1339–48.

    Article  Google Scholar 

  68. Babona-Pilipos R, Pritchard-Oh A, Popovic MR, Morshead CM. Biphasic monopolar electrical stimulation induces rapid and directed galvanotaxis in adult subependymal neural precursors. Stem Cell Res Ther. 2015;6(1):67.

    Article  CAS  Google Scholar 

  69. Siegel D, Baca SM, Thompson DC, Huntsman MM, Mower MM, Ross D. The effect of anodal/cathodal biphasic electrical stimulation on insulin release. J Cell Physiol. 2019;234(9):16389–99.

    Article  CAS  Google Scholar 

  70. Ercan B, Webster TJ. The effect of biphasic electrical stimulation on osteoblast function at anodized nanotubular titanium surfaces. Biomaterials. 2010;31(13):3684–93.

    Article  CAS  Google Scholar 

  71. Redouté JM, Browne D, Fitrio D, Lowery A, Kleeman L. A reduced data bandwidth integrated electrode driver for visual intracortical neural stimulation in 0.35 μ m mathcontainer loading mathjax high voltage CMOS. Microelectron J. 2013;44(4):277–82.

    Article  Google Scholar 

  72. van dan Honert C, Mortimer JT. A technique for collision block of peripheral nerve: frequency dependence. IEEE Trans Biomed Eng. 1981;28(5):379–82.

    Article  Google Scholar 

  73. Celik ME, Karagoz I. Comparison of monophasic and biphasic electrical stimulation by using temporal analysis for different inter-electrode spacings in the hexagonal arrays. Arab J Sci Eng. 2018;43(6):2889–98.

    Article  Google Scholar 

  74. Yadollahpour A, Jalilifar M. Electromagnetic fields in the treatment of wound: a review of current techniques and future perspective. J Pure Appl Microbio. 2014;8(4):2863–77.

    Google Scholar 

  75. Esfandiari E, Roshankhah S, Mardani M, Hashemibeni B, Naghsh E, Kazemi M, et al. The effect of high frequency electric field on enhancement of chondrogenesis in human adipose-derived stem cells. Iran J Basic Med Sci. 2014;17(8):571–6.

    Google Scholar 

  76. Mardani M, Roshankhah S, Hashemibeni B, Salahshoor M, Naghsh E, Esfandiari E. Induction of chondrogenic differentiation of human adipose-derived stem cells by low frequency electric field. Adv Biomed Res. 2016;5(1):97.

    Article  CAS  Google Scholar 

  77. Tzoneva R. Influence of electric field on cell behavior. Electrotreatment of cells for biomedical applications. Asian J Phys. 2014;23:789–814.

    Google Scholar 

  78. Ross CL. The use of electric, magnetic, and electromagnetic field for directed cell migration and adhesion in regenerative medicine. Biotechnol Prog. 2017;33(1):5–16.

    Article  CAS  Google Scholar 

  79. Zohar B, Blinder Y, Mooney DJ, Levenberg S. Flow-induced vascular network formation and maturation in three-dimensional engineered tissue. ACS Biomater Sci Eng. 2017;4(4):1265–71.

    Article  CAS  Google Scholar 

  80. Colosi C, Shin SR, Manoharan V, Massa S, Costantini M, Barbetta A, et al. Microfluidic bioprinting of heterogeneous 3d tissue constructs using low-viscosity bioink. Adv Mater. 2016;28(4):677–84.

    Article  CAS  Google Scholar 

  81. Min SK, Mi HL, Kwon BJ, Seo HJ, Koo MA, You KE, et al. Effects of direct current electric-field using ITO plate on breast cancer cell migration. Biomater Res. 2014;18(1):1–6.

    Article  Google Scholar 

  82. Li YH, Huang GY, Zhang XH, Wang L, Du YA, Lu TJ, et al. Engineering cell alignment in vitro [J]. Biotechnol Adv. 2014;32(2):347–65.

    Article  CAS  Google Scholar 

  83. Tai G, Tai M, Zhao M. Electrically stimulated cell migration and its contribution to wound healing [J]. Burns Trauma. 2018;6(1):20.

    Article  Google Scholar 

  84. Shao SJ, Zhou SB, Li L, Li JR, Luo C, Wang JX, et al. Osteoblast function on electrically conductive electrospun pla/mwcnts nanofibers. Biomaterials. 2011;32(11):2821–33.

    Article  CAS  Google Scholar 

  85. Ko UH, Park S, Bang H, Kim M, Shin H, Shin JH. Promotion of myogenic maturation by timely application of electric field along the topographical alignment. Tissue Eng Part A. 2018;24(9–10):752–60.

    Article  CAS  Google Scholar 

  86. Ganji Y, Li Q, Quabius ES, Böttner M, Selhuber-Unkel C, Kasra M. Cardiomyocyte behavior on biodegradable polyurethane/gold nanocomposite scaffolds under electrical stimulation. Mater Sci Eng C Mater Biol Appl. 2016;59:10–8.

    Article  CAS  Google Scholar 

  87. Zhao Z, Qin L, Reid B, Pu J, Hara T, Zhao M. Directing migration of endothelial progenitor cells with applied dc electric fields. Stem Cell Res. 2012;8(1):38–48.

    Article  CAS  Google Scholar 

  88. Banks TA, Luckman PSB, Frith JE, Cooper-White JJ. Effects of electric fields on human mesenchymal stem cell behaviour and morphology using a novel multichannel device. Integr Biol (Camb). 2015;7(6):693–712.

    Article  CAS  Google Scholar 

  89. Hammerick KE, Longaker MT, Prinz FB. In vitro effects of direct current electric fields on adipose-derived stromal cells. Biochem Biophys Res Commun. 2010;397(1):12–7.

    Article  CAS  Google Scholar 

  90. Chen C, Ruan S, Bai X, Lin C, Xie C, Lee I-S. Patterned iridium oxide film as neural electrode interface: biocompatibility and improved neurite outgrowth with electrical stimulation. Mater Sci Eng C. 2019;103:109865.

    Article  CAS  Google Scholar 

  91. Long H, Yang G, Ma K, Xiao Z, Ren X. Effect of different electrical stimulation waves on orientation and alignment of adipose derived mesenchymal stem cells. Zhongguo Xiu Fu Chong Jian Wai Ke Za Zhi. 2017;31(7):853–61.

    Google Scholar 

  92. Feng JF, Liu J, Zhang XZ, Zhang L, Jiang JY, Nolta J, et al. Guided migration of neural stem cells derived from human embryonic stem cells by an electric field. Stem Cells. 2012;30(2):349–55.

    Article  CAS  Google Scholar 

  93. Hoare JI, Rajnicek AM, McCaig CD, Barker RN, Wilson HM. Electric fields are novel determinants of human macrophage functions. J Leukoc Biol. 2016;99(6):1141–51.

    Article  CAS  Google Scholar 

  94. Hayashi H, Edin F, Li H, Liu W, Rask-Andersen H. The effect of pulsed electric fields on the electrotactic migration of human neural progenitor cells through the involvement of intracellular calcium signaling. Brain Res. 2016;1652:195–203.

    Article  CAS  Google Scholar 

  95. Kumar A, Nune KC, Misra RD. Electric field-mediated growth of osteoblasts - the significant impact of dynamic flow of medium. Biomater Sci. 2016;4(1):136–44.

    Article  CAS  Google Scholar 

  96. Canillas M, Moreno B, Chinarro E, Rajnicek AM. TiO2 surfaces support neuron growth during electric field stimulation. Mater Sci Eng C. 2017;79:1–8.

    Article  CAS  Google Scholar 

  97. Zhao Z, Watt C, Karystinou A, Roelofs AJ, McCaig CD, Gibson IR, et al. Directed migration of human bone marrow mesenchymal stem cells in a physiological direct current electric field. Eur Cells Mater. 2011;22:344–58.

    Article  CAS  Google Scholar 

  98. Guo A, Song B, Reid B, Gu Y, Forrester JV, Jahoda CA, et al. Effects of physiological electric fields on migration of human dermal fibroblasts. J Invest Dermatol. 2010;130(9):2320–7.

    Article  CAS  Google Scholar 

  99. Yao L, Li YC, Knapp J, Smith P. Exploration of molecular pathways mediating electric field-directed schwann cell migration by RNA-seq. J Cell Physiol. 2015;230(7):1515–24.

    Article  CAS  Google Scholar 

  100. Forciniti L, Ybarra J, Zaman MH, Schmidt CE. Schwann cell response on polypyrrole substrates upon electrical stimulation. Acta Biomater. 2014;10(6):2423–33.

    Article  CAS  Google Scholar 

  101. Iwasa SN, Babona-Pilipos R, Morshead CM. Environmental factors that influence stem cell migration: an “electric field”. Stem Cells Int. 2017;2017:4276927.

    Article  CAS  Google Scholar 

  102. Zhao H, Steiger A, Nohner M, Ye H. Specific intensity direct current (DC) electric field improves neural stem cell migration and enhances differentiation towards βiii-tubulin+ neurons. PLoS One. 2015;10(6):e0129625.

    Article  CAS  Google Scholar 

  103. Liu J, Zhu B, Zhang G, Wang J, Tian W, Ju G, et al. Electric signals regulate directional migration of ventral midbrain derived dopaminergic neural progenitor cells via wnt/gsk3β signaling. Exp Neurol. 2015;263:113–21.

    Article  CAS  Google Scholar 

  104. Lo KY, Wu SY, Sun YS. A microfluidic device for studying the production of reactive oxygen species and the migration in lung cancer cells under single or coexisting chemical/electrical stimulation. Microfluid Nanofluid. 2016;20(1):15.

    Article  CAS  Google Scholar 

  105. Jezierska-Wozniak K, Lipinski S, Huflejt M, Grabarczyk L, Barczewska M, Habich A, et al. Migration of human mesenchymal stem cells stimulated with pulsed electric field and the dynamics of the cell surface glycosylation. Adv Clin Exp Med. 2018;27(9):1181–93.

    Article  Google Scholar 

  106. Dong ZY, Pei Z, Li Z, Wang YL, Khan A, Meng XT. Electric field stimulation induced neuronal differentiation of filum terminale derived neural progenitor cells. Neurosci Lett. 2017;651:109–15.

    Article  CAS  Google Scholar 

  107. McKasson MJ, Huang L, Robinson KR. Chick embryonic schwann cells migrate anodally in small electrical fields. Exp Neurol. 2008;211(2):585–7.

    Article  Google Scholar 

  108. Li YP, Xu T, Chen XM, Lin S, Cho M, Sun D, et al. Effects of direct current electric fields on lung cancer cell electrotaxis in a PMMA-based microfluidic device. Anal Bioanal Chem. 2017;409(8):2163–78.

    Article  CAS  Google Scholar 

  109. Liu Q, Song B. Electric field regulated signaling pathways. Int J Biochem Cell B. 2014;55:264–8.

    Article  CAS  Google Scholar 

  110. Gao RC, Zhang XD, Sun YH, Kamimura Y, Mogilner A, Devreotes PN, et al. Different roles of membrane potentials in electrotaxis and chemotaxis of dictyostelium cells. Eukaryot Cell. 2011;10(9):1251–6.

    Article  CAS  Google Scholar 

  111. Zhao M, Penninger J, Isseroff RR. Electrical activation of wound-healing pathways. Adv Skin Wound Care. 2010;1(1):567–73.

    Google Scholar 

  112. Bai H, Forrester JV, Zhao M. DC electric stimulation upregulates angiogenic factors in endothelial cells through activation of VEGF receptors. Cytokine. 2011;55(1):110–5.

    Article  CAS  Google Scholar 

  113. Cortese B, Palama IE, D'Amone S, Gigli G. Influence of electrotaxis on cell behaviour. Integr Biol (Camb). 2014;6(9):817–30.

    Article  CAS  Google Scholar 

  114. Babona-Pilipos R, Droujinine IA, Popovic MR, Morshead CM. Adult subependymal neural precursors, but not differentiated cells, undergo rapid cathodal migration in the presence of direct current electric fields. PLoS One. 2011;6(8):e23808.

    Article  CAS  Google Scholar 

  115. Hernández-Bule ML, Paíno CL, Trillo MÁ, Úbeda A. Electric stimulation at 448 kHz promotes proliferation of human mesenchymal stem cells. Cell Physiol Biochem. 2014;34(5):1741–55.

    Article  CAS  Google Scholar 

  116. Meng XT, Arocena M, Penninger J, Gage FH, Zhao M, Song B. PI3K mediated electrotaxis of embryonic and adult neural progenitor cells in the presence of growth factors. Exp Neurol. 2011;227(1):210–7.

    Article  CAS  Google Scholar 

  117. Zhao M, Song B, Pu J, Wada T, Reid B, Tai G, et al. Electrical signals control wound healing through phosphatidylinositol-3-OH kinase-γ and PTEN. Nature. 2006;442(7101):457–60.

    Article  CAS  Google Scholar 

  118. Babona-Pilipos R, Liu N, Pritchard-Oh A, Mok A, Badawi D, Popovic M, et al. Calcium influx differentially regulates migration velocity and directedness in response to electric field application. Exp Cell Res. 2018;368(2):202–14.

    Article  CAS  Google Scholar 

  119. Kapeller B, Mueller J, Losert U, Podesser BK, Macfelda K. Microcurrent stimulation promotes reverse remodelling in cardiomyocytes. Esc Heart Failure. 2016;3(2):122–30.

    Article  Google Scholar 

  120. Wang Y, Cui HT, Wu ZX, Wu NP, Wang ZL, Chen XS, et al. Modulation of osteogenesis in MC3T3-E1 cells by different frequency electrical stimulation. PLoS One. 2016;11(5):e0154924.

    Article  CAS  Google Scholar 

  121. Naumova AV, Michel M, Anna M, Murry CE, Frank JA. Clinical imaging in regenerative medicine. Nat Biotechnol. 2014;32(8):804–18.

    Article  CAS  Google Scholar 

  122. Zhao C, Tan A, Pastorin G, Han KH. Nanomaterial scaffolds for stem cell proliferation and differentiation in tissue engineering. Biotechnol Adv. 2013;31(5):654–68.

    Article  CAS  Google Scholar 

  123. Santos N, Cicuéndez M, Holz T, Silva V, Fernandes A, Vila M, et al. Diamond-graphite nanoplatelet surfaces as conductive substrates for the electrical stimulation of cell functions. ACS Appl Mat Interfaces. 2017;9(2):1331–42.

    Article  CAS  Google Scholar 

  124. Dodel M, Nejad NH, Bahrami SH, Soleimani M, Amirabad LM, Hanaee-Ahvaz H, et al. Electrical stimulation of somatic human stem cells mediated by composite containing conductive nanofibers for ligament regeneration. Biologicals. 2017;46:99–107.

    Article  CAS  Google Scholar 

  125. Li Y, Li X, Zhao R, Wang C, Qiu F, Sun B, et al. Enhanced adhesion and proliferation of human umbilical vein endothelial cells on conductive pani-pcl fiber scaffold by electrical stimulation. Mater Sci Eng C Mater Biol Appl. 2017;72:106–12.

    Article  CAS  Google Scholar 

  126. Zhu W, Ye T, Lee SJ, Cui H, Miao S, Zhou X, et al. Enhanced neural stem cell functions in conductive annealed carbon nanofibrous scaffolds with electrical stimulation. Nanomedicine. 2018;14(7):2485–94.

    Article  CAS  Google Scholar 

  127. Snyder S, DeJulius C, Willits RK. Electrical stimulation increases random migration of human dermal fibroblasts. Ann Biomed Eng. 2017;45(9):2049–60.

    Article  Google Scholar 

  128. Mamman HB, Adon M, Abdul JM. Optimization of electric field parameters for ht29 cell line towards wound healing application. Indian J Sci Technol. 2016;9(46):1–7.

    Google Scholar 

  129. Sun B, Tong W, Wang J, Li D, Jing W, Qiang G, et al. Polypyrrole-coated poly(l-lactic acid-co-ε-caprolactone)/silk fibroin nanofibrous membranes promoting neural cells proliferation and differentiation with electrical stimulation. J Mater Chem B. 2016;4(41):6670–9.

    Article  CAS  Google Scholar 

  130. Chan YC, Ting S, Lee YK, Ng KM, Zhang J, Chen Z, et al. Electrical stimulation promotes maturation of cardiomyocytes derived from human embryonic stem cells. J Cardiovasc Transl Res. 2013;6(6):989–99.

    Article  Google Scholar 

  131. Hernández D, Millard R, Sivakumaran P, Wong RCB, Crombie DE, Hewitt AW, et al. Electrical stimulation promotes cardiac differentiation of human induced pluripotent stem cells. Stem Cells Int. 2016;2016:1718041.

    Article  CAS  Google Scholar 

  132. Park SY, Park J, Sim SH, Sung MG, Kim KS, Hong BH, et al. Enhanced differentiation of human neural stem cells into neurons on graphene. Adv Mater. 2011;23(36):H263–7.

    Article  CAS  Google Scholar 

  133. Creecy CM, O'Neill CF, Arulanandam BP, Sylvia VL, Navara CS, Bizios R. Mesenchymal stem cell osteodifferentiation in response to alternating electric current. Tissue Eng Part A. 2013;19(3–4):467–74.

    Article  CAS  Google Scholar 

  134. Zhang J, Neoh KG, Kang ET. Electrical stimulation of adipose-derived mesenchymal stem cells and endothelial cells co-cultured in a conductive scaffold for potential orthopaedic applications. J Tissue Eng Regen Med. 2018;12(4):878–89.

    Article  CAS  Google Scholar 

  135. Matsumoto M, Imura T, Fukazawa T, Sun Y, Takeda M, Kajiume T, et al. Electrical stimulation enhances neurogenin2 expression through β-catenin signaling pathway of mouse bone marrow stromal cells and intensifies the effect of cell transplantation on brain injury. Neurosci Lett. 2013;533:71–6.

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This research was supported by Natural Science Foundation of Zhejiang Province (LY20E010006, LQ16E020006), National Natural Science Foundation of China (51502265), Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education, Republic of Korea (2015R1D1A1A09058875), Opening Project of Zhejiang Provincial Top Key Discipline of Pharmaceutical Sciences, and the Project of Administration of Traditional Chinese Medicine of Zhejiang Province of China (No. 2018ZB011).

Funding

Not applicable.

Author information

Authors and Affiliations

Authors

Contributions

CC is responsible for modifying the manuscript. XB was responsible for initial literature review and writing of the manuscript. YHD made some pertinent suggestions. ISL contributed to the completion of the manuscript. Both authors read and approved the final manuscript.

Corresponding author

Correspondence to In-Seop Lee.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chen, C., Bai, X., Ding, Y. et al. Electrical stimulation as a novel tool for regulating cell behavior in tissue engineering. Biomater Res 23, 25 (2019). https://doi.org/10.1186/s40824-019-0176-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40824-019-0176-8

Keywords