Open access peer-reviewed chapter

Electrospun Polymeric Substrates for Tissue Engineering: Viewpoints on Fabrication, Application, and Challenges

Written By

Azadeh Izadyari Aghmiuni, Arezoo Ghadi, Elmira Azmoun, Niloufar Kalantari, Iman Mohammadi and Hossein Hemati Kordmahaleh

Submitted: 24 November 2021 Reviewed: 11 January 2022 Published: 08 July 2022

DOI: 10.5772/intechopen.102596

From the Edited Volume

Electrospinning - Material Technology of the Future

Edited by Tomasz Tański and Paweł Jarka

Chapter metrics overview

121 Chapter Downloads

View Full Metrics

Abstract

Electrospinning is the technique for producing nonwoven fibrous structures, to mimic the fabrication and function of the native extracellular matrix (ECM) in tissue. Prepared fibrous with this method can act as potential polymeric substrates for proliferation and differentiation of stem cells (with the cellular growth pattern similar to damaged tissue cells) and facilitation of artificial tissue remodeling. Moreover, such substrates can improve biological functions, and lead to a decrease in organ transplantation. In this chapter, we focus on the fundamental parameters and principles of the electrospinning technique to generate natural ECM-like substrates, in terms of structural and functional complexity. In the following, the application of these substrates in regenerating various tissues and the role of polymers (synthetic/natural) in the formation of such substrates is evaluated. Finally, challenges of this technique (such as cellular infiltration and inadequate mechanical strength) and solutions to overcome these limitations are studied.

Keywords

  • electrospinning process
  • parameters of electrospinning technique
  • synthetic/natural polymers
  • electrospun scaffolds
  • Tissue engineering

1. Introduction

Nowadays, tissue engineering is known as a multi-disciplinary science that leads to the regeneration of damaged/lost tissues via combining the cell and scaffold [1, 2]. In this technique, the engineered scaffolds act as micro/nano/smart environments for improving the interactions of cell-scaffold and cellular functions (such as proliferation, adhesion, differentiation, and growth) [3]. The methods that can be led to the design of scaffolds similar to extracellular matrices (ECM), with the properties of mechanical/biochemical support from the cells and imitation of architecture/structure and function of the tissues, play an important role in this field [4, 5, 6, 7].

Electrospinning is one of the unique approaches in this field that lead to the production of polymeric fibers with interconnecting pores and the opportunity to control the network and morphology of scaffolds, especially in bio-polymers processing. Such that, the polymeric solutions with low viscosity lead to shorter and finer scaffolds/substrates, while, more viscous solutions provide a relatively continuous scaffold. However, the morphology and diameter of the produced fibers depend on the processing conditions and the type of the polymer [8].

These conditions can be provided via controlling parameters of the solution, process, and ambient, such as an electric-field application, distance between the needle and collector, needle diameter, and flow rate, solution conductivity, the concentration of the polymeric solution, materials molecular weight, and solution viscosity [9, 10, 11, 12, 13].

In recent years, many studies have been carried out on electrospun substrates and their functions for tissue engineering applications, such as the regeneration of the blood vessels, skin tissue, cartilage, and bone, as well as muscle [14, 15, 16, 17, 18, 19, 20, 21, 22, 23]. However, a better understanding of cellular responses to sophisticated structures derived from this technique can be effective in reaching ECM-liked substrates.

Hence, in this chapter, we discuss the principles of the electrospinning technique to analyze these sophisticated structures and generate natural ECM-like substrates. In the following, the application of these substrates in regenerating various tissues and the role of polymers (synthetic/natural) in the formation of such substrates is evaluated. Finally, challenges of this technique and solutions to overcome these limitations are studied.

Advertisement

2. Electrospinning process

In this simple technique (spinning technique), a very high electrical field (high voltage), in the range of 10–50 kV, is applied for accelerating the charged polymer jet and producing ultrafine fibers (Figure 1A). Moreover, the electrospinning machine includes a syringe pump along with a syringe (with a metallic needle) that is loaded with the polymeric solutions, so that the tip needle is attached to one of the negative or positive terminals of the high-voltage electrical field, and pendant-shaped droplets of the polymer solution are held through surface tension. Notably, the needle tip is usually attached to the positive terminal of the electrical field [24]. In these conditions, the increase of the voltage is led to the formation of the Taylor cone in the needle tip [25]. Afterward, increased voltage leads to the creation of the critical value above which the electrostatic forces can overcome the surface tension forces so that it results in ejecting out the fine-jet of the solutions from the tip of the Taylor cone (Figure 1B). In the following, solvents are evaporated at a low boiling point due to contact with the atmosphere and subsequently the charged polymeric strands deposited on the collector. The collectors can play a crucial impact in reaching the various structures of the scaffold. Such that, unidirectionally oriented nanofibers, aligned nanofibers, and nonwoven nanofibers can be designed by square frame collector, rotating collector-drum, and flat plate stationary collector [26], respectively (Figure 1C).

Figure 1.

(A) Schematic of electrospinning process, (B) Taylor cone formation via increasing the voltage, (C) the fibers derived from electrospinning process with different collectors.

Advertisement

3. Electrospinning parameters

The parameters of the electrospinning process play an important role in understanding the nature of this process and conversing polymeric solutions into nanofibers. Indeed, control of these parameters can be led to electrospun fibers with a desired morphology and diameter. Hence, this section has been focused on these parameters and their influence on the properties of produced fibers. The mentioned parameters have been listed in Table 1.

ParametersDetailsDescriptionReferences
SolutionConcentrationThere are four critical concentrations for this parameter:
  1. Very low concentrations that led to polymeric micro/nanoparticles

  2. Little higher concentrations that led to the mixture of bead and fiber.

  3. Suitable concentrations that led to smooth nanofibers.

  4. Very high concentrations that led to helix-shaped micro-ribbons.

Notably, adjusting the concentration can tune the polymeric solution viscosity.
[27, 28, 29, 30]
Molecular weight (Mw)This parameter can reflect the entanglement of polymer chains in the polymeric solution. However, when sufficient intermolecular interactions are supplied via oligomers, the control of Mw is not essential for electrospinning processing.
Generally, in constant concentrations, the reduction of molecular weight is led to the formation of beads, rather than smooth fibers. Likewise, the increase in the polymer molecular weight provides smooth fibers. Notably, the too-high molecular weights can form micro-ribbon and some patterned fibers, even in low concentrations.
[31, 32, 33]
ViscosityThe selection of suitable viscosity plays a critical impact in the determination of the fiber morphologies, so that, the very low viscosity cannot provide the continuous/smooth fibers. Moreover, very high viscosity leads to the harder ejection of the jet from polymeric solutions. The concentration and molecular weight are the effective parameters on the viscosity, such that, adjustment of these two parameters can lead to adjustment solution viscosity.[34, 35, 36, 37]
Surface tensionThis parameter is considered as the solvent composition function in the polymeric solutions and plays a crucial impact in the electrospinning process. Indeed, surface tension can act as a dominant factor for the low viscosity of the solutions and lead to the bead/beaded fiber formation. In these conditions, the controlled constant concentration can reduce the surface tension and beaded fibers convert into smooth fibers. Moreover, given that different solvents possess different surface tensions, the change of the mass ratio of the solvent mix can lead to an adjustment in solution viscosity and surface tension.[30, 38, 39, 40]
Conductivity/surface charge densityThis parameter is usually determined and tuned via the type of solvent (such as formic acid), polymer (such as PCL), and salts (such as KH2PO4, NaCl). Accordingly, the use of ionic salts and organic acids leads to the production of nanofibers with small diameters and solutions with high conductivity, respectively.
Generally, biopolymers are led to more poor fibers compared to synthetic polymers; due to their polyelectrolytic nature that led to an increase in the carried charge by ions, and higher tension by the electric field
[41, 42]
ProcessingVoltageIn this parameter, the voltages higher than the threshold voltage can lead to the ejection of the charged jet from Taylor Cone. Moreover, the applied voltage plays an important role in adjusting the diameter of fibers, so that, based on the studies, high voltage can facilitate the formation of fibers with a large diameter. There are also several reports that show the increased voltages and consequently increase of electrostatic-repulsive forces of the charged jets can provide the fibers with a smaller diameter. Moreover, some studies also indicated that higher voltages can increase the formation of beads[27, 40, 43, 44, 45]
Flow rate of polymeric solutions within the syringeBased on the reports, a lower flow rate can be more effective in the electrospinning process due to the better polarization of polymeric solution. Indeed, in very high flow rates, bead fibers (thick diameters) form rather than the smooth fibers (thin diameters) that can be due to lower stretching force and the short time of drying polymeric solution (prior to contacting the collector)[44, 45]
CollectorThe collectors (mainly Al foil) are usually used as conductive substrates for collecting the charged fiber. In this field, there are many collectors such as wire meshes, parallel or gridded bars, pins, rotating rods or wheels, liquid baths, grids, etc.[46, 47, 48, 49]
Distance between the collector and the syringe tipThis parameter also plays a crucial impact in controlling the diameter of fibers and their morphology [12]. Such that, in the too-short distances, the fibers do not possess enough time for solidifying prior to contacting the collector. Likewise, too-long distances lead to the formation of bead fibers. Indeed, this parameter is known as the physical factor of electrospun fibers and plays an important role in drying solvent[43, 45]
AmbientTemperature and humidityThe temperature and humidity are known as ambient parameters that can be effective on the diameter of fibers and their morphology. Based on the reports, the increase in process temperature can lead to the formation of fibers with a thinner diameter, as well as low humidity can be resulted in increasing solvent evaporation velocity and drying the solvent. Likewise, higher humidity can form the fibers with thick diameters due to the neutralization of charge on the jet and smaller stretching force[45, 50]

Table 1.

The important parameters in the electrospinning process, as well as the morphology of electrospun fibers.

Advertisement

4. Polymers for electrospun fibers

Nowadays, natural and synthetic polymers are widely used in the design of electrospun scaffolds for tissue engineering applications [6, 51, 52, 53, 54, 55, 56]. In this field, synthetic polymers possess high flexibility in the electrospinning process and can provide fibers with better mechanical properties [56, 57]. Although, these polymers are also highly cost-effective than bio/natural polymers, however, a comparison of these two polymers indicates that synthetic polymers lack bioactivity and need more modification to improve biological properties [53, 56, 58]. In contrast, natural polymers possess the properties of inherent bioactive and can be led to an increase in the interactions of scaffold-cell and cell–cell (i.e., adhesion, proliferation, differentiation) [3]. These polymers have a relatively low immune response (in terms of chemical degradation) and can provide a structure similar to native ECM. In recent years, more than 200 natural/synthetic/copolymer/hybrid polymers have been designed and studied to obtain electrospun scaffolds with suitable physicomechanical and biological properties for use in tissue engineering [16, 17, 59, 60, 61, 62, 63, 64, 65]. Some of these polymers and their applications have been listed in Table 2.

The type of polymerNameThe conditions for the electrospinning processApplications in tissue engineeringReferences
NaturalCollagenCollagen (type I), in the presence of EDC and NHS (as chemical crosslinking) solvent: ethanol-PBS 5 mL syringe (21 gauge), voltage: 20 kV, pump rate: 0.5 ml/h, relatively low humidity (20%), the speed of drum rotating: 5 m s−1, and with a distance of 12 cm from the needle and the electrospun fibers
Results:
  • The mean diameter of obtained fibers: 0.42 ± 0.11 μm

  • – The increase of the viscosity by adding crosslinker EDC to the collagen solution

To produce water-insoluble nano-fibers and create a uniaxial tensile behavior similar to native tissue.[66]
The syringe with an 18-gauge, rate of source solution by syringe pump: 0 to 25 ml/h, solvent: HFP, voltage: 15–30 kV, mandrel rotated: ∼500 rpm, the optimal air gap distance: ∼125 mm
Results:
  • 100 nm fibers with the 67 nm banding pattern

  • The control of fiber orientation led to tailoring subtle mechanical properties into the matrix

  • The promotion of cell growth and its penetration into the electrospun collagen matrix

To optimize the concentrations, input voltage, delivery rate, air gap distance, and mandrel motion in electrospinning of collagen (type-I)[67]
Solvent: hexafluoro-2-propanol (HFP), syringe pump, and a 3 mL syringe with a 27½-G needle, voltage: 15 kV, distance between the needle tip and the collector: 15 cm
Results:
  • The improvement of adhesion and proliferation of rabbit conjunctiva fibroblast on aligned collagen scaffolds

Blood vessel tissue engineering[68]
GelatinSolvent: HFP, a syringe with a needle diameter of 500 lm, injection: onto a metal collector, injection rate: 0.06 ml/h, Voltage: 12 kV/8 cm, temperature: 37°C.
Results:
  • The co-solvents can be effective for electrospinning water-soluble natural polymers

Tissue engineering[69]
FibrinogenSolvent: HFP, 5.0-mL syringe (18-gauge needle) into a syringe pump, infection rate: 1.8 mL/h, voltage: 22 kV, distance between the needle tip and grounded target: 10 cm, rotating rate: 500 rpm.
Results:
  • The efficient penetration of rat cardiac fibroblasts an electrospun matrix

  • The excellent candidate for soft tissue applications

To design biomimicking fibrous scaffolds for tissue engineering applications[70]
4 mm diameter syringe with a 23¾-gauge, the syringe pump flow rate: 1.9 ml/h, voltage: 22 kV, distance between the needle tip and collector: 12.5 cm or 20 cm.
Results:
  • High extensibility, low modulus

To apply in tissue engineering and polymer composite and biomimetic.[71]
Hyaluronic acidVoltage: 0–40 kV, solvent: HCL, the aluminum foil as the collecting plate, and distance between the two electrodes: 9.5 cm.
Results:
  • The diameter of fibers was 49–74 nm

Tissue engineering[72]
SilkThe electrospinning process consists of a steel capillary tube (1.5 mm), volume flow rate: constant via the electric potential, the control of the distance between the capillary tip and collection (aluminum foil), and flow rate for creating a stable jet without dripping.
Results:
  • The increase of vascular cell growth

  • The formation of a short cord-like structure derived from HAECs after 4 days.

  • The formation of interconnection network of capillary tubes with lumens, after 7 days.

  • Suitable mechanical properties along with slow degradability

Vascular/blood vessels tissue engineering[73]
SyntheticNano/micro poly(L-lactic acid) fibersSolvent: DMF and DCM, needle: 18-G and inner diameter of 1.2 mm, the distance between the needle tip and the collector: 10 cm, a rotating disk along with a flat aluminum plate, and rotating rate: 1000 rpm, voltage: 12 kV.
Results:
  • Fiber diameter: 300–3500 nm

  • Suitable for the neural stem cells culture

Neural tissue engineering[74]
Poly(lactide-co-glycolide) (PLGA) electrospun matrixSolvent: HFP, and voltage of 25 kV, the needle with18 gage, flow rate: 3.0 mL/h. Distance between the needle tip and circular mandrel: 15 cm, and rotation rate: 500 rpm.
Results:
  • Suitable tissue composition and ideal mechanical properties.

  • Biocompatible with cell interactions and without systemic/local toxic effect.

Blood vessel tissue engineering[75]
Poly(caprolactone) nanofiberSolvent: chloroform/DMF, a syringe pump with a needle (inner diameter: 0.21 mm). Distance between the needle tip and aluminum plate: 15 cm, rotating rate of aluminum: 1000 rpm, voltage: 15 kV.
Results:
  • Improvement of endothelial cell spreading.

  • The promotion of proliferation and control and growth of cell orientation.

Blood vessel tissue engineering[76]
Copolymer/hybridPCL/HA hybrid scaffoldVolume flow rate: 1.5 ml/h with a needle (ID = 0.8 mm), in the range of voltage: 15–20 kV, the distance between the capillary tube and the grounded target: 12–15 cm, and rotating rate: 300 r/min.
Results:
  • The improvement of mechanical properties

  • The improvement of biochemical functions of nanofibers

  • The increase in the motility of skin fibroblast

  • The decrease in the limitations of electrospun scaffolds

To fabricate skin tissue-engineered scaffold[77]
HA–collagen nanofibrous matrixSolvent: formic acid: HFIP (30/70 v/v), 5 ml syringe with a needle tip diameter of 0.96 mm, solution flow rate: 2.54 ml h−1, voltage: 23 kV, the distance between needle tip and collector: 10 cm.
Results:
  • Mean diameter of fibers: 200 nm

  • The structure similar to native ECM

  • As wound dressing for scarless skin regeneration

To fabricate a substrate for skin regeneration[78]
Sodium alginate/poly(vinyl alcohol) nanofibrous containing nano ZnOVoltage: 17 kV, syringe pump with 10 ml syringe and needle of 21 gauges, the target: aluminum foil fixed on a wooden stand, spinning rate: 0.1 ml/h, the distance between the needle tip and aluminum target: 5 cm.
Results:
  • The increase in thermal stability of sodium alginate/PVA fibers, due to the existence of ZnO

To produce antibacterial wound dressings[79]
Alginate-PEO nanofibers containing lavender essential oilSolvent: water and DMF, syringe with a 23-gauge needle, syringe pump flow rate: 0.5 mL/h, voltage: 25 kV, distance between the needle tip and aluminum collector: 20 cm).
Results:
  • Reduction of the production of pro-inflammatory cytokines

  • Wound healing and skin regeneration

To design a hybrid scaffold with antibacterial and anti-inflammatory activities[80]
Composite nanofibers of chitin/silk fibroinSolvent: 1,1,1,3,3,3-hexafluoro-2-propanol (HFIP), distance between the needle tip and target: 7 cm, voltage: 17 kV, the mass flow rate: 4 ml/h, room temperature.
Results:
  • The gradual decrease in the diameter of fibers occurred by increased chitin concentration

  • The improvement of cell attachment, spreading of epidermal keratinocyte and dermal fibroblast

Skin tissue engineering[81]
PLLA-gelatin scaffold with a layer of electrospun PDLASolvent: DCM and DMF
For PDLA solution: plastic syringe with the 18-gauge needle, at the rate of 5 ml h−1, rotating (∼2000 rpm), and the distance: 15 cm, voltage: +12 kV and −5 kV. For gelatin/PLLA scaffold: the working distance: 5 cm (onto the PDLA layer), and voltages: +18 kV and −7 kV
Results:
  • The increase of compressive mechanical properties.

  • Hydroxyapatite incorporation with PLLA/gelatin scaffold provided a significant effect on the mineral depositions on this scaffold.

Bone tissue engineering[82]

Table 2.

The used polymers in the electrospinning process to produce nano/microfibers.

4.1 Natural polymers

4.1.1 Collagen

Collagen is one of the main components of the native ECM with diameters in the range of 10–500 nm that plays an important role in providing mechanical strength of tissue and stimulating cell attachment and its proliferation [67, 83, 84]. Generally, type I collagen is the most common type of this protein in the dermis (70–80%), compared to other types of collagens (i.e., type II and type III) [85]. Given that collagen possesses a low young modulus (0.8 GPa) [83], the processes, such as chemical modification [by covalent of amine/imine linkage], cross-linking [by Glutaraldehyde (GA), NHS, and EDC, genipin as the cross-linking agent], and physical treatment [by UV irradiation, gamma radiation and dehydrothermal treatment (DHT)], can be led to an increase in mechanical properties of electrospun nanofibers based on this biopolymer [84, 86, 87, 88, 89, 90, 91, 92, 93, 94, 95, 96].

4.1.2 Gelatin

Gelatin derived from collagen is also one of the other biopolymers whose surface charge depends on the gelatin processing methods, such as acidic/alkaline processing [97]. The mechanical strength of this biopolymer can also increase via the physical mixing with other polymers or cross-linking process and immersing of gelatin-based scaffolds into glutaraldehyde (25%), carbodiimides, and genipin [83, 98].

4.1.3 Chitosan

Chitosan is another biopolymer that is widely used in biomedical/tissue engineering applications due to its low toxicity, non-immunogenic, biodegradability, and antibacterial properties [51, 98]. This polysaccharide as a cationic biopolymer can interact with structural molecules of the ECM due to having positively charged and be led to the formation of two-component scaffolds with suitable physicomechanical/biological properties when mixed with other anionic biopolymers [99].

4.1.4 Fibrinogen

Fibrinogen is a soluble biopolymer derived from the blood plasma and plays an important role in tissue engineering applications and the development of electrospun substrates/scaffolds [100, 101]. Based on the reports, fibrinogen electrospun fibers are more extensible and elastic compared to other biopolymer-based electrospun fibers [98, 102]. Notably, the mechanical resistance of this biopolymer and its degradation rate can be controlled by crosslinking process or supplementing culture medium [103].

4.1.5 Elastin

Elastin is a biopolymer of highly insoluble with difficult processing, hence the approaches, such as cross-linking and blending with other polymers can be effective in reaching elastin-based scaffolds [104].

4.1.6 Silk fibroin

Silk is a protein-biopolymer derived from Bombyx more (most commonly)/other insects and included two proteins of fibroin (70–80%) and sericin (20–30%) or other insects [105, 106, 107, 108, 109]. This biomacromolecular possesses suitable biocompatibility, biodegradability, and mechanical properties and plays an important role in the biomedical applications and development of engineered scaffolds [110, 111, 112, 113, 114]. Based on the reports, the electrospun nanofibers of this biopolymer can modulate cellular interactions (such as adhesion, spread, the expression level of genes and proteins) [115, 116].

4.1.7 Hyaluronic acid

Hyaluronic acid (HA) is known as a linear polysaccharide and plays an important role in the ECM structure [117, 118]. This glycosaminoglycan (GAG) is a suitable candidate for hydrogel production and tissue regeneration due to its molecular weight (100–8000 kDa) and hygroscopic nature [118, 119, 120]. However, the high viscosity of produced hydrogels leads to limitations in the electrospinning process, hence, the electrospun fibers can form by blending/dissolving this hydrogel with other polymers/ in other solvents [65].

4.1.8 Alginate

Alginate is a copolymer derived from β-D-mannuronic acid and α-L-guluronic acid that is widely used in tissue engineering applications [121, 122]. This biopolymer is known as one of the other polysaccharides that are commercially produced from brown algae or some bacteria, such as Azotobacter chroococcum, Azotobacter vinelandii, and some species of Pseudomonas [121, 123]. The salt of this polymer (alginate sodium) is water-soluble and can also form a highly viscous solution at very low concentrations [124]. Moreover, the electrospinning of alginate sodium solutions carries out in presence of synthetic polymers (PVA, PEO, etc.), owing to the reduction of the repulsive forces among the poly-anionic chains of alginate. Methods, such as cross-linking, are also necessary for resulting scaffold stability in the aqueous environment [65].

4.2 Synthetic polymers

4.2.1 Polylactic acid (PLA)

PLA as thermoplastic polymer fabricates from the polymerization of lactic acid and possesses isomers of poly(L-lactic acid) and poly(D-lactic acid) [125, 126]. This aliphatic polyester possesses biodegradability properties and plays a critical role in the design of tissue-engineered scaffolds [127]. However, PLA is usually copolymerized with other polymers or made as a composite due to its low modulus, especially in bone tissue engineering [128, 129].

4.2.2 Polycaprolactone (PCL)

This polymer is a semicrystalline polyester that is widely used in the applications of tissue engineering due to its biodegradability, nontoxicity, and biocompatibility properties, as well as mechanical strength [130, 131, 132]. This synthetic polymer can improve cellular penetration into the engineered scaffolds due to the presence of cell recognition sites [51, 133, 134]. However, the degraded product of this polymer (the acids of polylactide and glycolide) affects on stability and functions of proteins and bioactive molecules. PCL is also known as the most famous synthetic polymer in the design of bone-engineered scaffolds, due to its low degradation rate and high modulus [83].

4.2.3 Polyglycolic acid (PGA)

PGA is the aliphatic thermoplastic polyester (simple linear) that is usually applied in the applications of bone tissue engineering. This polymer possesses a high young modulus (7 GPa) and can be completely degraded within 4–6 months [133, 135].

4.2.4 Polyethylene glycol (PEG)

PEG is known as one of the most popular synthetic polymers in tissue engineering applications and can lead to a promotion of cellular adhesion and improvement of cell–cell signaling, due to its hydrophilic properties and interactions with the chains of polysaccharides or peptides [51, 136]. Notably, copolymerization of this polymer with other hydrophobic polymers, such as polyglycolic acid, polycaprolactone, and polylactic acid, can lead to an increase in the degradation rate of these polymers and neutralization of their acidic products [137, 138].

4.3 Copolymer/hybrid polymers

Tissue engineering is a strategy for the design of tissue-liked scaffolds/ substrates (in terms of biological and physiological functions). In this field, the combination of natural or synthetic polymers as copolymers or hybrid scaffolds can be played an important role in overcoming the limitations of mono-component systems and improving the interactions of cell/cell and cell/scaffold. Sodium alginate/PVA electrospun mats are a sample from these copolymers that aimed to prepare the antibacterial substrates [79]. Such substrates can be used in wound dressings to reduce wound infection and prevent scars [79, 139]. PCL/HA nanofibrous scaffolds are also one of the other composite scaffolds that can provide substrates with better mechanical and biochemical properties. Based on the reports, such scaffolds lead to an increase in fibroblasts infiltration into the scaffold, cellular proliferation, and consequently tissue regeneration [77]. In this field, the combination of HA and collagen has been reported as an ideal matrix in electrospun dressings that play an important role in reducing scar via proteinase secretion and metalloproteinase inhibition [78]. Alginate-PEO nanofibers containing lavender essential oil are another scaffold that can be used for wound healing and reducing the production of pro-inflammatory cytokines [80]. The PLLA-gelatin scaffold with a layer of electrospun PDLA is one of the other samples in this field [82].

Advertisement

5. Application of electrospun polymers in tissue engineering

The cells play an important role in the formation of organ-dependent extracellular matrix (ECM), and to this end, they need microenvironments to improve their functions. However, the body is unable to repair damaged tissue when tissue damage is severe or extensive. In this field, although, the xenografts, autografts, and allografts approaches can be used, however, the problems of donor sites, antigenicity, and immunogenicity have been limited to the use of these therapeutic methods [51]. In recent years, tissue engineering as a new method has been overcome the mentioned problems via designing engineered substrates with suitable physicomechanical and biological properties [3, 51]. There are many studies that show electrospun substrates can be effective in this field. Such that, the electrospinning process of natural and synthetic polymers can help to address cell requirements, improve its functions, and finally regenerate damaged tissue.

Blood vessels are one of the important organs in the body that need to be repaired quickly in case of damage. Vascular-engineered scaffolds play a key role in this regard so that electrospun matrices based on the natural/synthetic/hybrid polymers can support the adhesion, differentiation, and proliferation of vascular cells and be led to tissue regeneration [140]. Accordingly, the study of Bondar et al. shows that there is ideal intercellular contact between endothelial cells on the nano/micro-scale electrospun silk fiber that led to vascular endothelial cadherin expression [141]. Soffer et al. also designed silk fibroin into a tubular structure (inner-diameter: ∼3 mm, the average wall thickness: 0.15 mm). They reported that the average strength of such scaffolds is much more than collagen scaffolds [142].

Skin is another organ that possesses an important role in the protection of the body against infection and environmental agents. Such that, loss of a large surface of the skin due to burn, wound, etc. can lead to patient death. However, dermis-engineered scaffolds can be an ideal therapeutic option for skin regeneration [6]. In this field, Min et al. stated that silk electrospun substrates with coating collagen I can increase adhesion and spreading of keratinocytes. Moreover, they found that laminin coating can stimulate cellular spreading, but not cellular adhesion [112]. Based on the studies of Pezeshki-Modaress et al., the gelatin/chitosan electrospun scaffolds as the structures containing protein and polysaccharide play a crucial role in the proliferation of dermal fibroblast cells and wound healing. Such scaffolds can maintain their morphology in culture medium and increase the proliferation and attachment of cells [143]. Law et al. also assessed electrospun collagen nanofibers for applications of skin tissue engineering. They stated that collagen nanofibers possess low mechanical properties, hence are usually cross-linked/blended with synthetic polymers [144]. In this field, Jin et al. showed that collagen/poly(l-lactic acid)–co-poly(3-caprolactone) electrospun scaffolds can differentiate mesenchymal stem cells to epidermal and lead to an increase in cell proliferation [145].

One of the other applications of electrospun scaffolds is related to the design of calcified extracellular matrices. These substrates consist of calcium phosphates, carbonated hydroxyapatite, growth factors, and bone marrow stromal cells (MSCs) or osteoblasts and osteoblast-like cells, such as osteoprogenitors and osteocytes [140]. In this regard, Ki et al. designed the 3D silk matrices to produce bone with MC3T3-E1 cells. They found that cell proliferation and spreading increase on the 3D matrices compared to 2D matrices. This can be due to the higher porosity of 3D matrices that provides better cellular adhesion [146]. In another study, Yin et al. showed that electrospun scaffolds can be effective on the regeneration of various tissues, via topography-dependent induction of lineage-specific differentiation [147]. They evaluated the differentiation pathway of MSCs in forming new tissues and observed that these scaffolds can lead to the formation of tendon-like tissue in the Achilles tendon injury models, as well as, chondrogenesis and bone tissue formation via ossification. Delgado-Rangel et al. also developed collagen/poly (vinyl alcohol)/chondroitin sulfate and collagen/poly (vinyl alcohol)/hyaluronic acid 3D electrospun scaffolds to apply tissue engineering [148]. Based on their reports, these scaffolds can increase the biocompatible cross-linker. Moreover, the scaffolds possess the behavior of pH-sensitive swelling and can be used in drug delivery systems.

Flaig et al. also studied the application of electrospun scaffolds in cardiac tissue engineering. They found that electrospun scaffolds based on poly (glycerol sebacate) elastomer and poly (lactic acid) can induce neovascularization without the inflammatory responses and support cardiomyocyte development [149]. Moreover, Vogt et al. stated that poly (ε-caprolactone)/poly (glycerol sebacate)-based electrospun scaffolds possess better mechanical properties compared to native myocardium; hence can be potentially suitable to apply cardiac tissue engineering [150].

Collectively, the studies indicate that electrospun scaffolds possess an inductive role in the regeneration of tissues and the use of hybrid polymers can provide effective insight into the design and development of smart scaffolds for applications of tissue engineering.

Advertisement

6. Challenges and resolutions of the electrospinning process

The techniques of tissue engineering hold promise for developing functional networks similar to native tissue. The designed substrates in this technique can support the formation of 3D tissues by mimicking ECM functions. Among the fabrication techniques of the engineered scaffold, the electrospinning process is known as an outstanding one that can produce a nonwoven structure.

Although this method is considered as the potential technique in the design of tissue-engineered scaffolds, however, it possesses limitations of mechanical strength and cellular infiltration in the application of load-bearing.

Based on the reports, the optimum size of pores for cell infiltration to tissue is in a range of 100–500 μm [151], while, the size of pores in electrospun scaffolds is much lower than the mentioned size. This can lead to inhibition of vascular growth and the creation of the hypoxic region. Moreover, the density of the fibers in electrospun substrates can be one of the reasons for poor cellular infiltration [152].

There are several solutions to overcome these limitations that pore architecture of scaffolds and their surface morphology control are some of the most important solutions. Generally, the diameter of fibers strongly relates to the pore diameter in the electrospun substrates, so that, fibers with smaller diameters lead to smaller pores. Hence, attention to surface topography plays a crucial role in the removal of waste and diffusion of nutrients [153, 154].

Indeed, the manipulation of characteristics of the electrospun scaffolds for enlarging the diameter of pores or reduction of the fibers density, help migration of cells into internal parts of the scaffold. There are other four approaches in this field, including [155]:

  1. Adding biological factors

  2. Electrospraying or layering of cells

  3. Dynamic cellular culture

  4. Combination of electrospinning with other fibers-fabrication techniques.

Advertisement

7. Conclusion

The electrospinning process is known as the powerful, simple, and inexpensive tool to fabricate tissue-engineering substrates that are capable of the formation of ECM-mimicking networks. Although, this technique, in clinical applications; has the limitations, such as low cellular infiltration, high-density of fibers, possible toxicity of solvent/cross-linker, and insufficient mechanical strength. However, some solutions, such as the increased diameter of pores, reduced density of fibers, and electrospinning polymers along with cells can overcome these problems. Combining robust materials or structures also provides the more robust electrospun substrate for the design and production of tissue substitutes with the desired target.

References

  1. 1. Jun I, Han HS, Edwards JR, Jeon H. Electrospun fibrous scaffolds for tissue engineering: Viewpoints on architecture and fabrication. International Journal of Molecular Sciences. 2018;19. DOI: 10.3390/ijms19030745
  2. 2. Chen F-M, Liu X. Advancing biomaterials of human origin for tissue engineering. Progress in Polymer Science. 2016;53:86-168
  3. 3. Izadyari Aghmiuni A, Heidari Keshel S, Sefat F, AkbarzadehKhiyavi A. Fabrication of 3D hybrid scaffold by combination technique of electrospinning-like and freeze-drying to create mechanotransduction signals and mimic extracellular matrix function of skin. Materials Science and Engineering: C. 2021;120:111752
  4. 4. Thomas K, Engler AJ, Meyer GA. Extracellular matrix regulation in the muscle satellite cell niche. Connective Tissue Research. 2015;56:1-8
  5. 5. Wen S-L, Feng S, Tang S-H, Gao J-H, Zhang L, Tong H, et al. Collapsed reticular network and its possible mechanism during the initiation and/or progression of hepatic fibrosis. Scientific Reports. 2016;6:1-11
  6. 6. Izadyari Aghmiuni A, Heidari Keshel S, Sefat F, Akbarzadeh Khiyavi A. Quince seed mucilage-based scaffold as a smart biological substrate to mimic mechanobiological behavior of skin and promote fibroblasts proliferation and h-ASCs differentiation into keratinocytes. International Journal of Biological Macromolecules. 2020;142:668-679
  7. 7. Caralt M, Uzarski JS, Iacob S, Obergfell KP, Berg N, Bijonowski BM, et al. Optimization and critical evaluation of decellularization strategies to develop renal extracellular matrix scaffolds as biological templates for organ engineering and transplantation. American Journal of Transplantation. 2015;15:64-75
  8. 8. Bera B. Literature review on electrospinning process (a fascinating fiber fabrication technique). International Journal of Interdisciplinary Research (IJIR). 2016;2:972-984
  9. 9. Pillay V, Dott C, Choonara YE, Tyagi C, Tomar L, Kumar P, et al. A review of the effect of processing variables on the fabrication of electrospun nanofibers for drug delivery applications. Journal of Nanomaterials. 2013;2013:1-22
  10. 10. Mirjalili M, Zohoori S. Review for application of electrospinning and electrospun nanofibers technology in textile industry. Journal of Nanostructure in Chemistry. 2016;6:207-213
  11. 11. Zafar M, Najeeb S, Khurshid Z, Vazirzadeh M, Zohaib S, Najeeb B, et al. Potential of electrospun nanofibers for biomedical and dental applications. Materials (Basel). 2016;9:73
  12. 12. Correia DM, Ribeiro C, Ferreira JCC, Botelho G, Ribelles JLG, Lanceros-Méndez S, et al. Influence of electrospinning parameters on poly (hydroxybutyrate) electrospun membranes fiber size and distribution. Polymer Engineering and Science. 2014;54:1608-1617
  13. 13. Motamedi AS, Mirzadeh H, Hajiesmaeilbaigi F, Bagheri-Khoulenjani S, Shokrgozar M. Effect of electrospinning parameters on morphological properties of PVDF nanofibrous scaffolds. Progress in Biomaterials. 2017;6:113-123
  14. 14. Vatankhah E, Prabhakaran MP, Semnani D, Razavi S, Morshed M, Ramakrishna S. Electrospun tecophilic/gelatin nanofibers with potential for small diameter blood vessel tissue engineering. Biopolymers. 2014;101:1165-1180
  15. 15. Krishnan KV, Columbus S, Krishnan LK. Alteration of electrospun scaffold properties by silver nanoparticle incorporation: Evaluation for blood vessel tissue engineering. Tissue Engineering Part A. 2015;21:S239-S240
  16. 16. Sundaramurthi D, Krishnan UM, Sethuraman S. Electrospun nanofibers as scaffolds for skin tissue engineering. Polymer Reviews. 2014;54:348-376
  17. 17. Chen S, Liu B, Carlson MA, Gombart AF, Reilly DA, Xie J. Recent advances in electrospun nanofibers for wound healing. Nanomedicine. 2017;12:1335-1352
  18. 18. Norouzi M, Boroujeni SM, Omidvarkordshouli N, Soleimani M. Advances in skin regeneration: application of electrospun scaffolds. Advanced Healthcare Materials. 2015;4:1114-1133
  19. 19. Wang C, Wang M. Electrospun multicomponent and multifunctional nanofibrous bone tissue engineering scaffolds. Journal of Materials Chemistry. 2017;5:1388-1399
  20. 20. Xu T, Miszuk JM, Zhao Y, Sun H, Fong H. Electrospun polycaprolactone 3D nanofibrous scaffold with interconnected and hierarchically structured pores for bone tissue engineering. Advanced Healthcare Materials. 2015;4:2238-2246
  21. 21. Kheradmandi M, Vasheghani-Farahani E, Ghiaseddin A, Ganji F. Skeletal muscle regeneration via engineered tissue culture over electrospun nanofibrous chitosan/PVA scaffold. Journal of Biomedical Materials Research Part A. 2016;104:1720-1727
  22. 22. Elsayed Y, Lekakou C, Labeed F, Tomlins P. Smooth muscle tissue engineering in crosslinked electrospun gelatin scaffolds. Journal of Biomedical Materials Research Part A. 2016;104:313-321
  23. 23. Liu W, Li Z, Zheng L, Zhang X, Liu P, Yang T, et al. Electrospun fibrous silk fibroin/poly (L-lactic acid) scaffold for cartilage tissue engineering. Tissue Engineering and Regenerative Medicine. 2016;13:516-526
  24. 24. Fang J, Wang X, Lin T. Functional applications of electrospun nanofibers. Nanofibers-Production, Properties and Functional Applications. 2011;14:287-302
  25. 25. Awang N, Ismail AF, Jaafar J, Matsuura T, Junoh H, Othman MHD, et al. Functionalization of polymeric materials as a high performance membrane for direct methanol fuel cell: A review. Reactive and Functional Polymers. 2015;86:248-258
  26. 26. Khadka DB, Haynie DT. Protein- and peptide-based electrospun nanofibers in medical biomaterials. Nanomedicine Nanotechnology Biologie et Médecine. 2012;8:1242-1262
  27. 27. Deitzel JK, Harris D, Tan NB. The effect of processing variables on the morphology of electrospun nanofibers and textiles. Polymer (Guildf). 2001;42:261-272
  28. 28. Eda G, Shivkumar S. Bead-to-fiber transition in electrospun polystyrene. Journal of Applied Polymer Science. 2007;106:475-487
  29. 29. Lee KH, Kim HY, Bang HJ, Jung YH, Lee SG. The change of bead morphology formed on electrospun polystyrene fibers. Polymer (Guildf). 2003;44:4029-4034
  30. 30. Yang Q, Li Z, Hong Y, Zhao Y, Qiu S, Wang C, et al. Influence of solvents on the formation of ultrathin uniform poly(vinyl pyrrolidone) nanofibers with electrospinning. Journal of Polymer Science Part B Polymer Physics. 2004;42:3721-3726
  31. 31. McKee MG, Layman JM, Cashion MP, Long TE. Phospholipid nonwoven electrospun membranes. Science. 2006;311(80):353-355
  32. 32. Koski A, Yim K, Shivkumar S. Effect of molecular weight on fibrous PVA produced by electrospinning. Materials Letters. 2004;58:493-497
  33. 33. Zhao YY, Yang QB, Lu X-F, Wang C, Wei Y. Study on correlation of morphology of electrospun products of polyacrylamide with ultrahigh molecular weight. Journal of Polymer Science Part B Polymer Physics. 2005;43:2190-2195
  34. 34. Larrondo L, John Manley RS. Electrostatic fiber spinning from polymer melts. I. Experimental observations on fiber formation and properties. Journal of Polymer Science Part B Polymer Physics. 1981;19:909-920
  35. 35. Sukigara S, Gandhi M, Ayutsede J, Micklus M, Ko F. Regeneration of Bombyx mori silk by electrospinning—part 1: processing parameters and geometric properties. Polymer (Guildf). 2003;44:5721-5727
  36. 36. Zhang Y, Ouyang H, Lim CT, Ramakrishna S, Huang Z-M. Electrospinning of gelatin fibers and gelatin/PCL composite fibrous scaffolds. Journal of Biomedical Materials Research. 2005;72B:156-165
  37. 37. Ki CS, Baek DH, Gang KD, Lee KH, Um IC, Park YH. Characterization of gelatin nanofiber prepared from gelatin–formic acid solution. Polymer (Guildf). 2005;46:5094-5102
  38. 38. Haghi AK, Akbari M. Trends in electrospinning of natural nanofibers. Physica Status Solidi. 2007;204:1830-1834
  39. 39. Pham QP, Sharma U, Mikos AG. Electrospun poly(ε-caprolactone) microfiber and multilayer nanofiber/microfiber scaffolds: Characterization of scaffolds and measurement of cellular infiltration. Biomacromolecules. 2006;7:2796-2805
  40. 40. Zhang C, Yuan X, Wu L, Han Y, Sheng J. Study on morphology of electrospun poly(vinyl alcohol) mats. European Polymer Journal. 2005;41:423-432
  41. 41. Zong X, Kim K, Fang D, Ran S, Hsiao BS, Chu B. Structure and process relationship of electrospun bioabsorbable nanofiber membranes. Polymer (Guildf). 2002;43:4403-4412
  42. 42. Huang C, Chen S, Lai C, Reneker DH, Qiu H, Ye Y, et al. Electrospun polymer nanofibres with small diameters. Nanotechnology. 2006;17:1558-1563
  43. 43. Yuan X, Zhang Y, Dong C, Sheng J. Morphology of ultrafine polysulfone fibers prepared by electrospinning. Polymer International. 2004;53:1704-1710
  44. 44. Buchko CJ, Chen LC, Shen Y, Martin DC. Processing and microstructural characterization of porous biocompatible protein polymer thin films. Polymer (Guildf). 1999;40:7397-7407
  45. 45. Li Z, Wang C. One-Dimensional Nanostructures: Electrospinning Technique and Unique Nanofibers. New York, Dordrecht, London: Springer Berlin Heidelberg; 2013. pp. 15-29. DOI: 10.1007/978-3-642-36427-3_2
  46. 46. Wang X, Um IC, Fang D, Okamoto A, Hsiao BS, Chu B. Formation of water-resistant hyaluronic acid nanofibers by blowing-assisted electro-spinning and non-toxic post treatments. Polymer (Guildf). 2005;46:4853-4867
  47. 47. Li D, Wang Y, Xia Y. Electrospinning nanofibers as uniaxially aligned arrays and layer-by-layer stacked films. Advanced Materials. 2004;16:361-366
  48. 48. Xu C. Aligned biodegradable nanofibrous structure: A potential scaffold for blood vessel engineering. Biomaterials. 2004;25:877-886
  49. 49. Ki CS, Kim JW, Hyun JH, Lee KH, Hattori M, Rah DK, et al. Electrospun three-dimensional silk fibroin nanofibrous scaffold. Journal of Applied Polymer Science. 2007;106:3922-3928
  50. 50. Mit-uppatham C, Nithitanakul M, Supaphol P. Ultrafine electrospun polyamide‐6 fibers: Effect of solution conditions on morphology and average fiber diameter. Macromolecular Chemistry and Physics. 2004;205:2327-2338
  51. 51. Aghmiuni AI, Baei MS, Keshel SH, Khiyavi AA. Design of novel 3D-scaffold as a potential material to induct epidermal-dermal keratinocytes of human-adipose-derived stem cells and promote fibroblast cells proliferation for skin regeneration. Fibers and Polymers. 2020;21:33-44
  52. 52. Biazar E, Keshel SH. The healing effect of stem cells loaded in nanofibrous scaffolds on full thickness skin defects. Journal of Biomedical Nanotechnology. 2013;9:1471-1482
  53. 53. Bassas-Galia M, Follonier S, Pusnik M, Zinn M. Natural polymers. In: Bioresorbable Polymers for Biomedical Applications. Amsterdam, Netherlands: Elsevier; 2017. pp. 31-64. DOI: 10.1016/B978-0-08-100262-9.00002-1
  54. 54. Wang S, Lu L, Wang C, Gao C, Wang X. International Journal of Polymer Science. 2010;2010:1-2
  55. 55. Sell SA, Wolfe PS, Garg K, McCool JM, Rodriguez IA, Bowlin GL. The use of natural polymers in tissue engineering: A focus on electrospun extracellular matrix analogues. Polymers (Basel). 2010;2:522-553
  56. 56. Place ES, George JH, Williams CK, Stevens MM. Synthetic polymer scaffolds for tissue engineering. Chemical Society Reviews. 2009;38:1139-1151
  57. 57. Kitsara M, Agbulut O, Kontziampasis D, Chen Y, Menasché P. Fibers for hearts: A critical review on electrospinning for cardiac tissue engineering. Acta Biomaterialia. 2017;48:20-40
  58. 58. Sengupta P, Prasad BLV. Surface modification of polymeric scaffolds for tissue engineering applications. Regenerative Engineering and Translational Medicine. 2018;4:75-91
  59. 59. Ahire JJ, Robertson D, Neveling DP, Van Reenen AJ, Dicks LMT. Hyaluronic acid-coated poly(d,l-lactide) (PDLLA) nanofibers prepared by electrospinning and coating. RSC Advances. 2016;6:34791-34796
  60. 60. Bosworth LA, Alam N, Wong JK, Downes S. Investigation of 2D and 3D electrospun scaffolds intended for tendon repair. Journal of Materials Science: Materials in Medicine. 2013;24:1605-1614
  61. 61. Kumbar SG, Nukavarapu SP, James R, Nair LS, Laurencin CT. Electrospun poly(lactic acid-co-glycolic acid) scaffolds for skin tissue engineering. Biomaterials. 2008;29:4100-4107
  62. 62. Mehrasa M, Asadollahi MA, Ghaedi K, Salehi H, Arpanaei A. Electrospun aligned PLGA and PLGA/gelatin nanofibers embedded with silica nanoparticles for tissue engineering. International Journal of Biological Macromolecules. 2015;79:687-695
  63. 63. Chen SH, Chen CH, Shalumon KT, Chen JP. Preparation and characterization of antiadhesion barrier film from hyaluronic acid-grafted electrospun poly(caprolactone) nanofibrous membranes for prevention of flexor tendon postoperative peritendinous adhesion. International Journal of Nanomedicine. 2014;9:4079-4092
  64. 64. Nanofibers and their applications as templates for nanofibers preparation metal oxide nanofibers preparation
  65. 65. Mele E. Electrospinning of natural polymers for advanced wound care: Towards responsive and adaptive dressings. Journal of Materials Chemistry B. 2016;4:4801-4812
  66. 66. Meng L, Arnoult O, Smith M, Wnek GE. Electrospinning of in situ crosslinked collagen nanofibers. Journal of Materials Chemistry. 2012;22:19412-19417
  67. 67. Matthews JA, Wnek GE, Simpson DG, Bowlin GL. Electrospinning of collagen nanofibers. Biomacromolecules. 2002;3:232-238
  68. 68. Zhong S, Teo WE, Zhu X, Beuerman RW, Ramakrishna S, Yung LYL. An aligned nanofibrous collagen scaffold by electrospinning and its effects on in vitro fibroblast culture. The Journal of Biomedical Materials Research Part A An Official Journal of Society for Biomaterials Australian Society for Biomaterials; Korean Society for Biomaterials. 2006;79:456-463
  69. 69. Song J-H, Kim H-E, Kim H-W. Production of electrospun gelatin nanofiber by water-based co-solvent approach. Journal of Materials Science: Materials in Medicine. 2008;19:95-102
  70. 70. McManus MC, Boland ED, Simpson DG, Barnes CP, Bowlin GL. Electrospun fibrinogen: Feasibility as a tissue engineering scaffold in a rat cell culture model. The Journal of Biomedical Materials Research Part A An Official Journal of Society for Biomaterials Australian Society for Biomaterials.; Korean Society for Biomaterials. 2007;81:299-309
  71. 71. Carlisle CR, Coulais C, Namboothiry M, Carroll DL, Hantgan RR, Guthold M. The mechanical properties of individual, electrospun fibrinogen fibers. Biomaterials. 2009;30:1205-1213
  72. 72. Um IC, Fang D, Hsiao BS, Okamoto A, Chu B. Electro-spinning and electro-blowing of hyaluronic acid. Biomacromolecules. 2004;5:1428-1436
  73. 73. Zhang X, Baughman CB, Kaplan DL. In vitro evaluation of electrospun silk fibroin scaffolds for vascular cell growth. Biomaterials. 2008;29:2217-2227
  74. 74. Yang F, Murugan R, Wang S, Ramakrishna S. Electrospinning of nano/micro scale poly(L-lactic acid) aligned fibers and their potential in neural tissue engineering. Biomaterials. 2005;26:2603-2610
  75. 75. Stitzel J, Liu J, Lee SJ, Komura M, Berry J, Soker S, et al. Controlled fabrication of a biological vascular substitute. Biomaterials. 2006;27:1088-1094
  76. 76. Ma Z, He W, Yong T, Ramakrishna S. Grafting of gelatin on electrospun poly (caprolactone) nanofibers to improve endothelial cell spreading and proliferation and to control cell orientation. Tissue Engineering. 2005;11:1149-1158
  77. 77. Qian Y, Li L, Jiang C, Xu W, Lv Y, Zhong L, et al. The effect of hyaluronan on the motility of skin dermal fibroblasts in nanofibrous scaffolds. International Journal of Biological Macromolecules. 2015;79:133-143
  78. 78. Hsu F-Y, Hung Y-S, Liou H-M, Shen C-H. Electrospun hyaluronate–collagen nanofibrous matrix and the effects of varying the concentration of hyaluronate on the characteristics of foreskin fibroblast cells. Acta Biomaterialia. 2010;6:2140-2147
  79. 79. Shalumon KT, Anulekha KH, Nair SV, Nair SV, Chennazhi KP, Jayakumar R. Sodium alginate/poly (vinyl alcohol)/nano ZnO composite nanofibers for antibacterial wound dressings. International Journal of Biological Macromolecules. 2011;49:247-254
  80. 80. Hajiali H, Summa M, Russo D, Armirotti A, Brunetti V, Bertorelli R, et al. Alginate–lavender nanofibers with antibacterial and anti-inflammatory activity to effectively promote burn healing. Journal of Materials Chemistry B. 2016;4:1686-1695
  81. 81. Park KE, Jung SY, Lee SJ, Min BM, Park WH. Biomimetic nanofibrous scaffolds: Preparation and characterization of chitin/silk fibroin blend nanofibers. International Journal of Biological Macromolecules. 2006;38:165-173
  82. 82. Andric T, Wright LD, Taylor BL, Freeman JW. Fabrication and characterization of three‐dimensional electrospun scaffolds for bone tissue engineering. Journal of Biomedical Materials Research Part A. 2012;100:2097-2105
  83. 83. Bhattarai DP, Aguilar LE, Park CH, Kim CS. A review on properties of natural and synthetic based electrospun fibrous materials for bone tissue engineering. Membranes. 2018;8:62. DOI: 10.3390/membranes8030062
  84. 84. Zuber M, Zia F, Zia KM, Tabasum S, Salman M, Sultan N. Collagen based polyurethanes—A review of recent advances and perspective. International Journal of Biological Macromolecules. 2015;80:366-374
  85. 85. Shields KJ, Beckman MJ, Bowlin GL, Wayne JS. Mechanical properties and cellular proliferation of electrospun collagen type II. Tissue Engineering. 2004;10:1510-1517
  86. 86. Boland ED, Matthews JA, Pawlowski KJ, Simpson DG, Wnek GE, Bowlin GL. Electrospinning collagen and elastin: preliminary vascular tissue engineering. Frontiers in Bioscience. 2004;9:e32
  87. 87. Barnes CP, Pemble CW IV, Brand DD, Simpson DG, Bowlin GL. Cross-linking electrospun type II collagen tissue engineering scaffolds with carbodiimide in ethano. Tissue Engineering. 2007;13:1593-1605
  88. 88. Sell SA, McClure MJ, Garg K, Wolfe PS, Bowlin GL. Electrospinning of collagen/biopolymers for regenerative medicine and cardiovascular tissue engineering. Advanced Drug Delivery Reviews. 2009;61:1007-1019
  89. 89. Liu S, Wu J, Liu X, Chen D, Bowlin GL, Cao L, et al. Osteochondral regeneration using an oriented nanofiber yarn‐collagen type I/hyaluronate hybrid/TCP biphasic scaffold. Journal of Biomedical Materials Research Part A. 2015;103:581-592
  90. 90. Wang W, Zhang Y, Ye R, Ni Y. Physical crosslinkings of edible collagen casing. International Journal of Biological Macromolecules. 2015;81:920-925
  91. 91. Takitoh T, Bessho M, Hirose M, Ohgushi H, Mori H, Hara M. Gamma-cross-linked nonfibrillar collagen gel as a scaffold for osteogenic differentiation of mesenchymal stem cells. Journal of Bioscience and Bioengineering. 2015;119:217-225
  92. 92. Maslennikova A, Kochueva M, Ignatieva N, Vitkin A, Zakharkina O, Kamensky V, et al. Effects of gamma irradiation on collagen damage and remodeling. International Journal of Radiation Biology. 2015;91:240-247
  93. 93. Suchý T, Šupová M, Sauerová P, Verdánová M, Sucharda Z, Rýglová Š, et al. The effects of different cross-linking conditions on collagen-based nanocomposite scaffolds—An in vitro evaluation using mesenchymal stem cells. Biomedical Materials. 2015;10:65008
  94. 94. Yoshioka SA, Goissis G. Thermal and spectrophotometric studies of new crosslinking method for collagen matrix with glutaraldehyde acetals. Journal of Materials Science: Materials in Medicine. 2008;19:1215-1223
  95. 95. Vrana NE, Builles N, Kocak H, Gulay P, Justin V, Malbouyres M, et al. EDC/NHS cross-linked collagen foams as scaffolds for artificial corneal stroma. Journal of Biomaterials Science, Polymer Edition. 2007;18:1527-1545
  96. 96. Liu T, Wang Z. Collagen crosslinking of porcine sclera using genipin. Acta Ophthalmologica. 2013;91:e253-e257
  97. 97. Young S, Wong M, Tabata Y, Mikos AG. Gelatin as a delivery vehicle for the controlled release of bioactive molecules. Journal of Controlled Release;109, 2005:256-274
  98. 98. Nemati S, Kim S, Shin YM, Shin H. Current progress in application of polymeric nanofibers to tissue engineering. Nano Convergence. 2019. DOI: 10.1186/s40580-019-0209-y
  99. 99. Nilsen-Nygaard J, Strand SP, Vårum KM, Draget KI, Nordgård CT. Chitosan: Gels and interfacial properties. Polymers (Basel). 2015;7:552-579
  100. 100. Wnek GE, Carr ME, Simpson DG, Bowlin GL. Electrospinning of nanofiber fibrinogen structures. Bachelor of Engineering, Virginia Commonwealth University. 2003:1-4
  101. 101. Rothwell SW, Sawyer E, Dorsey J, Flournoy WS, Settle T, Simpson D, et al. Wound healing and the immune response in swine treated with a hemostatic bandage composed of salmon thrombin and fibrinogen. Journal of Materials Science: Materials in Medicine. 2009;20:2155-2166
  102. 102. Baker S, Sigley J, Helms CC, Stitzel J, Berry J, Bonin K, et al. The mechanical properties of dry, electrospun fibrinogen fibers. Materials Science and Engineering: C. 2012;32:215-221
  103. 103. McManus MC, Boland ED, Koo HP, Barnes CP, Pawlowski KJ, Wnek GE, et al. Mechanical properties of electrospun fibrinogen structures. Acta Biomaterialia. 2006;2:19-28
  104. 104. Leach JB, Wolinsky JB, Stone PJ, Wong JY. Crosslinked α-elastin biomaterials: towards a processable elastin mimetic scaffold. Acta Biomaterialia. 2005;1:155-164
  105. 105. Wang Y, Kim H-J, Vunjak-Novakovic G, Kaplan DL. Stem cell-based tissue engineering with silk biomaterials. Biomaterials. 2006;27:6064-6082
  106. 106. Qi C, Xu L, Deng Y, Wang G, Wang Z, Wang L. Sericin hydrogels promote skin wound healing with effective regeneration of hair follicles and sebaceous glands after complete loss of epidermis and dermis. Biomaterials Science. 2018;6:2859-2870
  107. 107. Wray LS, Hu X, Gallego J, Georgakoudi I, Omenetto FG, Schmidt D, et al. Effect of processing on silk‐based biomaterials: Reproducibility and biocompatibility. Journal of Biomedical Materials Research Part B: Applied Biomaterials. 2011;99:89-101
  108. 108. Rockwood DN, Preda RC, Yücel T, Wang X, Lovett ML, Kaplan DL. Materials fabrication from Bombyx mori silk fibroin. Nature Protocols. 2011;6:1612-1631
  109. 109. Melke J, Midha S, Ghosh S, Ito K, Hofmann S. Silk fibroin as biomaterial for bone tissue engineering. Acta Biomaterialia. 2016;31:1-16
  110. 110. Meinel L, Betz O, Fajardo R, Hofmann S, Nazarian A, Cory E, et al. Silk based biomaterials to heal critical sized femur defects. Bone. 2006;39:922-931
  111. 111. Lin S, Chen M, Jiang H, Fan L, Sun B, Yu F, et al. Green electrospun grape seed extract-loaded silk fibroin nanofibrous mats with excellent cytocompatibility and antioxidant effect. Colloids Surfaces B Biointerfaces. 2016;139:156-163
  112. 112. Min B-M, Lee G, Kim SH, Nam YS, Lee TS, Park WH. Biomaterials. 2004;25:1289-1297
  113. 113. Kasoju N, Bora U. Silk fibroin in tissue engineering. Advanced Healthcare Materials. 2012;1:393-412
  114. 114. Sun K, Li H, Li R, Nian Z, Li D, Xu C. Silk fibroin/collagen and silk fibroin/chitosan blended three-dimensional scaffolds for tissue engineering. European Journal of Orthopaedic Surgery and Traumatology. 2015;25:243-249
  115. 115. Fan Z, Shen Y, Zhang F, Zuo B, Lu Q, Wu P, et al. Control of olfactory ensheathing cell behaviors by electrospun silk fibroin fibers. Cell Transplantation. 2013;22:39-50
  116. 116. Di Felice V, Serradifalco C, Rizzuto L, De Luca A, Rappa F, Barone R, et al. Silk fibroin scaffolds enhance cell commitment of adult rat cardiac progenitor cells. Journal of Tissue Engineering and Regenerative Medicine. 2015;9:E51-E64
  117. 117. Kogan G, Šoltés L, Stern R, Gemeiner P. Hyaluronic acid: A natural biopolymer with a broad range of biomedical and industrial applications. Biotechnology Letters. 2007;29:17-25
  118. 118. Sudha PN, Rose MH. Beneficial effects of hyaluronic acid. Advances in Food and Nutrition Research. 2014;72:137-176
  119. 119. Trombino S, Servidio C, Curcio F, Cassano R. Strategies for hyaluronic acid-based hydrogel design in drug delivery. Pharmaceutics. 2019;11:407
  120. 120. Dai Prè E, Conti G, Sbarbati A. Hyaluronic acid (HA) scaffolds and multipotent stromal cells (MSCs) in regenerative medicine. Stem Cell Reviews and Reports. 2016;12:664-681
  121. 121. Khan F, Ahmad SR. Polysaccharides and their derivatives for versatile tissue engineering application. Macromolecular Bioscience. 2013;13:395-421
  122. 122. Gong Y, Han GT, Zhang YM, Zhang JF, Jiang W, Tao XW, et al. Preparation of alginate membrane for tissue engineering. Journal of Polymer Engineering. 2016;36:363-370
  123. 123. Sabra W, Zeng A-P, Deckwer W-D. Bacterial alginate: physiology, product quality and process aspects. Applied Microbiology and Biotechnology. 2001;56:315-325
  124. 124. Bhattarai N, Li Z, Edmondson D, Zhang M. Alginate‐based nanofibrous scaffolds: Structural, mechanical, and biological properties. Advanced Materials. 2006;18:1463-1467
  125. 125. Oksman K, Skrifvars M, Selin J-F. Natural fibres as reinforcement in polylactic acid (PLA) composites. Composites Science and Technology. 2003;63:1317-1324
  126. 126. Pang X, Zhuang X, Tang Z, Chen X. Polylactic acid (PLA): Research, development and industrialization. 2010. pp. 1125-1136
  127. 127. Jem KJ, van der Pol JF, de Vos S. Microbial lactic acid, its polymer poly (lactic acid), and their industrial applications. In: Plastics from Bacteria. Berlin/Heidelberg, Germany: Springer; 2010. pp. 323-346
  128. 128. Athanasiou KA, Niederauer GG, Agrawal CM. Sterilization, toxicity, biocompatibility and clinical applications of polylactic acid/polyglycolic acid copolymers. Biomaterials. 1996;17:93-102
  129. 129. Ruan R, Vasile C. Polylactic acid (PLA) synthesis and modifications: A review related papers. DOI: 10.1007/s11458-009-0092-x
  130. 130. Tiwari AP, Hwang TI, Oh J-M, Maharjan B, Chun S, Kim BS, et al. pH/NIR-responsive polypyrrole-functionalized fibrous localized drug-delivery platform for synergistic cancer therapy. ACS Applied Materials & Interfaces. 2018;10:20256-20270
  131. 131. Sun H, Mei L, Song C, Cui X, Wang P. The in vivo degradation, absorption and excretion of PCL-based implant. Biomaterials. 2006;27:1735-1740
  132. 132. Sung H-J, Meredith C, Johnson C, Galis ZS. The effect of scaffold degradation rate on three-dimensional cell growth and angiogenesis. Biomaterials. 2004;25:5735-5742
  133. 133. Guo B, Ma PX. Synthetic biodegradable functional polymers for tissue engineering: A brief review. Science China. Chemistry. 2014;57:490-500
  134. 134. Shalumon KT, Anulekha KH, Chennazhi KP, Tamura H, Nair SV, Jayakumar R. Fabrication of chitosan/poly(caprolactone) nanofibrous scaffold for bone and skin tissue engineering. International Journal of Biological Macromolecules. 2011;48:571-576
  135. 135. Gunatillake PA, Adhikari R, Gadegaard N. Biodegradable synthetic polymers for tissue engineering. European Cells & Materials. 2003;5:1-16
  136. 136. Chen S-H, Tsao C-T, Chang C-H, Lai Y-T, Wu M-F, Chuang C-N, et al. Assessment of reinforced poly(ethylene glycol) chitosan hydrogels as dressings in a mouse skin wound defect model. Materials Science and Engineering: C. 2013;33:2584-2594
  137. 137. Zhu KJ, Xiangzhou L, Shilin Y. Preparation, characterization, and properties of polylactide (PLA)–poly (ethylene glycol)(PEG) copolymers: A potential drug carrier. The Journal of Applied Polymer Science. 1990;39:1-9
  138. 138. Chai Y, Lin D, Ma Y, Yuan Y, Liu C. RhBMP-2 loaded MBG/PEGylated poly (glycerol sebacate) composite scaffolds for rapid bone regeneration. Journal of Materials Chemistry B. 2017;5:4633-4647
  139. 139. Pegg CE, Jones GH, Athauda TJ, Ozer RR, Chalker JM. Facile preparation of ammonium alginate-derived nanofibers carrying diverse therapeutic cargo. Chemical Communications. 2014;50:156-158
  140. 140. Zhang X, Reagan MR, Kaplan DL. Electrospun silk biomaterial scaffolds for regenerative medicine. Advanced Drug Delivery Reviews. 2009;61:988-1006
  141. 141. Bondar B, Fuchs S, Motta A, Migliaresi C, Kirkpatrick CJ. Functionality of endothelial cells on silk fibroin nets: comparative study of micro-and nanometric fibre size. Biomaterials. 2008;29:561-572
  142. 142. Soffer L, Wang X, Zhang X, Kluge J, Dorfmann L, Kaplan DL, et al. Silk-based electrospun tubular scaffolds for tissue-engineered vascular grafts. Journal of Biomaterials Science, Polymer Edition. 2008;19:653-664
  143. 143. Pezeshki M, Mojgan M, Sarah Z. Tailoring the gelatin/chitosan electrospun scaffold for application in skin tissue engineering: An in vitro study. Progress in Biomaterials. 2018;7:207-218
  144. 144. Xian J, Ling L, Liau L, Saim A, Yang Y. Electrospun collagen nanofibers and their applications in skin tissue engineering. Tissue Engineering and Regenerative Medicine. 2017. DOI: 10.1007/s13770-017-0075-9
  145. 145. Jin G, Prabhakaran MP, Ramakrishna S. Stem cell differentiation to epidermal lineages on electrospun nanofibrous substrates for skin tissue engineering. Acta Biomaterialia. 2011;7:3113-3122
  146. 146. Seok C, Sook KÆ, Park Y, Jeong ÆH, Jung KÆH. Development of 3-D nanofibrous fibroin scaffold with high porosity by electrospinning: implications for bone regeneration. 2008. pp. 405-410
  147. 147. Yin Z, Chen X, Song H, Hu J, Tang Q, Zhu T, et al. Electrospun scaffolds for multiple tissues regeneration in vivo through topography dependent induction of lineage specific differentiation. Biomaterials. 2015;44:173-185. DOI: 10.1016/j.biomaterials.2014.12.027
  148. 148. Delgado-rangel LH, Hernández-vargas J, Becerra-gonzález M. development of collagen/poly (vinyl alcohol)/chondroitin sulfate and collagen/poly(vinyl alcohol)/ha electrospun scaffolds for tissue engineering. 2019;20:2470-2484
  149. 149. Flaig F, Ragot H, Simon A, Revet G, Kitsara M, Kitasato L, et al. Design of functional electrospun scaffolds based on poly (glycerol sebacate) elastomer and poly (lactic acid) for cardiac tissue engineering. ACS Biomaterials Science & Engineering. 2020;6:2388-2400
  150. 150. Vogt L, Rivera LR, Liverani L, Piegat A, El Fray M, Boccaccini AR. Poly(ε-caprolactone)/poly(glycerol sebacate) electrospun scaffolds for cardiac tissue engineering using benign solvents. Materials Science and Engineering: C. 2019;103:109712
  151. 151. Wright LD, Andric T, Freeman JW. Utilizing NaCl to increase the porosity of electrospun materials. Materials Science and Engineering: C. 2011;31:30-36
  152. 152. Blakeney BA, Tambralli A, Anderson JM, Andukuri A, Lim D-J, Dean DR, et al. Cell infiltration and growth in a low density, uncompressed three-dimensional electrospun nanofibrous scaffold. Biomaterials. 2011;32:1583-1590
  153. 153. Rnjak-Kovacina J, Weiss AS. Increasing the pore size of electrospun scaffolds. Tissue Engineering. Part B, Reviews. 2011;17:365-372
  154. 154. Costa PF, Vaquette C, Zhang Q, Reis RL, Ivanovski S, Hutmacher DW. Advanced tissue engineering scaffold design for regeneration of the complex hierarchical periodontal structure. Journal of Clinical Periodontology. 2014;41:283-294
  155. 155. Khorshidi S, Solouk A, Mirzadeh H, Mazinani S, Lagaron JM, Sharifi S, et al. A review of key challenges of electrospun scaffolds for tissue-engineering applications. Journal of Tissue Engineering and Regenerative Medicine. 2016;10:715-738

Written By

Azadeh Izadyari Aghmiuni, Arezoo Ghadi, Elmira Azmoun, Niloufar Kalantari, Iman Mohammadi and Hossein Hemati Kordmahaleh

Submitted: 24 November 2021 Reviewed: 11 January 2022 Published: 08 July 2022