Download as pdf or txt
Download as pdf or txt
You are on page 1of 41

Progress in Materials Science xxx (xxxx) xxxx

Contents lists available at ScienceDirect

Progress in Materials Science


journal homepage: www.elsevier.com/locate/pmatsci

Electrospinning for tissue engineering applications


Maryam Rahmatia, David K. Millsb, Aleksandra M. Urbanskac,
Mohammad Reza Saebd, , Jayarama Reddy Venugopale,f, Seeram Ramakrishnae,

Masoud Mozafarig,

a
Department of Biomaterials, Institute of Clinical Dentistry, University of Oslo, 0317 Oslo, Norway
b
School of Biological Sciences and the Center for Biomedical Engineering and Rehabilitation Science, Louisiana Tech University, Ruston, USA
c
Division of Digestive and Liver Diseases, Columbia University Medical Center, New York, USA
d
Université de Lorraine, CentraleSupélec, LMOPS, F-57000 Metz, France
e
Center for Nanofibers & Nanotechnology, Nanoscience & Nanotechnology Initiative, Department of Mechanical Engineering, Faculty of Engineering,
National University of Singapore, Singapore
f
Faculty of Industrial Sciences & Technology, Universiti Malaysia Pahang, 26300 Gambang, Kuantan, Pahang, Malaysia
g
Department of Tissue Engineering & Regenerative Medicine, Faculty of Advanced Technologies in Medicine, Iran University of Medical Sciences
(IUMS), Tehran, Iran

Abbreviations: ADSCs, adipose-derived stem cells; ALP, alkaline phosphatase; AP-g-GA, pentamer-graft-gelatin; AV, aloe vera; BDDGE, 1,4-buta­
nediol diglycidyl ether; bFGF, basic fibroblast growth factor; BMP-2, bone morphogenetic protein-2; BMSCs, bone marrow stromal cells; β-TCP, beta
tricalcium phosphate; CaP, calcium phosphate; CDM, cartilage-derived matrix; CDPS, cistanche polysaccharide; ChABC, chondroitinase ABC; CMs,
cardiomyocytes; CNM, cardiac nanofibrous meshes; CORMs, carbon monoxide-releasing molecules; CPC, calcium phosphate cement; CTS, chitosan;
CVD, chemical vapor deposition; DMF, dimethylformamide; ECCs, engineered cardiac constructs; ECM, extracellular matrix; FBR, foreign body
response; GAGs, glycosaminoglycans; GBR, guided bone regeneration; GDNF, glial cell-derived neurotrophic factor; GO, graphene oxide; GT, ge­
latin; GP, genipin; HA, hydroxyapatite; HAECs, human aortic endothelial cells; HAM, human amniotic membrane; hASCs, human adipose-derived
stem cells; HCASMCs, human coronary artery smooth muscle cells; HCNFs, hollow carbon nanofibers; HDFs, human dermal fibroblasts; hFobs,
human fetal osteoblasts; hMSCs, human mesenchymal stem cells; HOBs, human osteoblasts; HUVECs, human umbilical vein endothelial cells; LLA-
TMC, L-lactide-co-trimethylene carbonate; MC, methylene chloride; MSM, methylsulfonylmethane; MWNTs, multi-walled carbon nanotubes; NFS,
nanofibrous fibroin scaffold; NHOst, normal human osteoblasts; nHA, HA nanoparticles; NT-3, neurotrophin-3; PA, polyamide; PAA, poly(acrylic
acid); PAN, polyacrylonitrile; PANI, polyaniline; PBLG, poly-benzyl-L-glutamate; PBS, phosphate-buffered saline; PBMSC, peripheral blood mono­
nuclear-stem cell; PCL, polycaprolactone; PCLEEP, caprolactone and ethyl ethylene phosphate; PCU, polycarbonate-urethane; PEO, poly(ethylene
oxide); PEOT/PBT, poly(ethylene oxide terephthalate)/poly(butylene terephthalate); PEDOT, poly(3,4-ethylenedioxythiophene); PELCL, poly
(ethylene glycol)-b-poly(L-lactide-co-ε-caprolactone); P-ESF, pore electrospun silk fibroin; PEU, poly(ester urea); PEUU, poly(ester-urethane) urea;
PG, poly(ε-caprolactone)/gelatin; PHB, polyhydroxybutyrate; PHBV, poly(3-hydroxybutyrate-co-3-hydroxyvalerate); PGS, poly(glycerol sebacate);
PLA, polylactic acid; PLCL, poly(lactic acid-co-caprolactone); PLLA, poly(L-lactic acid); PLGA, poly(lactic-co-glycolic acid); PGA, polyglycolic acid;
POC, poly(1,8-octanediol-co-citrate); PMMA, poly(methyl methacrylate); PSAN, poly(styrene-co-acrylonitrile); PRP, platelet-rich plasma; PTFE,
polytetrafluoroethylene; PU, polyurethane; PVA, polyvinyl alcohol; PVP, polyvinylpyrrolidone; Pμ, PCL microfibers; PμPn, PCL microfibers with PCL
nanofibers; RBMCs, rat bone marrow cells; rhBMP-2, recombinant bone morphogenetic protein-2; SDF-1α, stromal cell derived factor-1α; SEM,
scanning electron microscope; SF, silk fibroin; SIS, small intestine submucosa; SMC, smooth muscle cell; SrBG, strontium-substituted bioactive glass;
SrCO3, strontium carbonate; TCD, tip-to-collector distance; TGF-β, transforming growth factor beta; TPU, thermoplastic polyurethane; UV, ultra­
violet; 2D, two-dimensional; 3D, three-dimensional; 3DF, 3D fiber deposition; nMP, nano-magnesium phosphate; HPG, hyperbranched polyglycerol;
iPSCs, induced pluripotent stem cells; CNT, carbon nanotube; GAS, glucosamine sulfate; ACs, rat articular chondrocytes; CP5, an articular cartilage
progenitor cell line; KGN, kartogenin

Corresponding authors at: Lunenfeld-Tanenbaum Research Institute, Mount Sinai Hospital, University of Toronto, Toronto, Canada (M.
Mozafari).
E-mail addresses: mrsaeb2008@gmail.com (M.R. Saeb), mozafari.masoud@gmail.com, m.mozafari@utoronto.ca (M. Mozafari).

https://doi.org/10.1016/j.pmatsci.2020.100721
Received 22 October 2016; Received in revised form 28 July 2020; Accepted 9 August 2020
0079-6425/ © 2020 Elsevier Ltd. All rights reserved.

Please cite this article as: Maryam Rahmati, et al., Progress in Materials Science, https://doi.org/10.1016/j.pmatsci.2020.100721
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

ARTICLE INFO ABSTRACT

Keywords: Tissue engineering makes use of the principles of medicine, biology and engineering and in­
Electrospinning tegrates them into the design of biological substitutes to restore, maintain and improve the
Nanofibers functions of tissue. To fabricate a functional tissue, the engineered structures have to be able to
Scaffolds mimic the extracellular matrix (ECM), provide the tissue with oxygen and nutrient circulation as
Biomaterials
well as remove metabolic wastes in the period of tissue regeneration. Continued efforts have been
Nanofibrous constructs
Porous structures
made in order to fabricate advanced functional three-dimensional scaffolds for tissue en­
Tissue engineering gineering. Electrospinning has been recognized and served as one of the most useful techniques
Regenerative medicine based on the resemblance between electrospun fibers and the native tissues. Over the past few
decades, a bewildering variety of nanofibrous scaffolds have been developed for various bio­
medical applications, such as tissue regeneration and therapeutic agent delivery. The present
review aims to provide with researchers an in-depth understanding of the promising role and the
practical region of applicability of electrospinning in tissue engineering and regenerative medi­
cine by highlighting the outcomes of the most recent studies performed in this field. We address
the current strategies used for improving the physicochemical interactions between the cells and
the nanofibrous surface. We also discuss the progress and challenges associated with the use of
electrospinning for tissue engineering and regenerative medicine applications.

1. Introduction

In the past few years, biomedical engineers have placed substantial focus on fabricating multidisciplinary platforms to mimic the
structural and physicochemical features of natural tissues [1–3]. Researchers from diverse fields such as medicine, nanotechnology,
biology and engineering are working closely to integrate their findings into a modern technology for appropriate mimicking the
natural tissues [4–6]. Tissue engineering is known as a reliable methodology to combat the tissue injuries by mimicking the phy­
siological microenvironment [7–9]. In addition, recent nanotechnological advances provide an opportunity for further improving the
properties of tissue-engineered scaffolds [10–13]. Using nanoscience approaches, novel substitutes are developing fast, which make
possible to mimic the extracellular matrix (ECM) conditions of the natural tissues, more precisely than those of the macro- or micro-
scale biomaterials [14–16]. The success in making new strategies adjust to the regenerative medicine, e.g. cell-based therapies,
artificial organs and engineered living tissues, roots in proper design of multi-functional biomaterials with optimal physichochemical
properties [17]. Manipulating the physicochemical properties of biomaterial surface in view of the target application is the most
challenging aspect of engineering the injured organs, tissues and/or cells, for they play key roles in supporting the cell survival and
stimulating autologous tissue growth in situ [2,18,19]. A biomaterial should additionally have proper mechanical properties, as
complements to the target cell stiffness, making possible stimulating the neo-tissue formation [2]. There are several fabrication
methods available to obtain biomaterials with optimal physicochemical properties matching with those of the target tissue, such as
self-assembly [20], template synthesis [21], phase-separation [22], melt-blowing [23], and electrospinning [24]. Table 1 provides a
summary of the advantages and disadvantages of the main fabrication methods applied in producing scaffolds.
Among the available techniques, electrospinning is one of the most promising methods frequently used in tissue engineering

Table 1
Comparison of various methods applied in scaffold fabrication in terms of benefits and difficulties associated with their usage [25–27].
Scaffold Fabrication Advantage Disadvantage
Method

Electrospinning Uniform, aligned fiber, strong interconnectivity of porosity, 80–95% Needs high voltage apparatus, solvents used may be
porosity, 100–1100 nm fiber diameter, < 80% cell viability, ECM like toxic, and difficulties in packaging, shipping, and
structure, superior mechanical properties, large surface area, and handling
facile and simple fabrication
Self-assembly 80–90% porosity, 70–90% cell viability, 5–300 nm Using peptides, complex process, not scalable, poor
control over fiber dimension
Phase Separation Simple fabrication, 60–95% porosity Use of potentially toxic solvents, poor control over
architecture, and restricted range of pore sizes
Gas Foaming Solvent-free, no loss of bioactive molecules in the scaffold matrix Needs high pressure, poor interconnectivity of porosity,
the existence of skimming film layers on the scaffold
surface
Solvent Casting Simple fabrication, high mechanical stability Lacks reproducibility, uncontrolled structure
Freeze Drying 30–80% porosity, 50–450 nm, cell viability < 90%, it needs neither Need freeze-dryer, limited to small pore size, irregular
high temperature nor a separate leaching step porosity, long processing time
3D Printing Fabrication desired structure (flexibility in production) Needs 3D printer, uses toxic organic solvents, lacks the
mechanical strength

2
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

applications. This technique has several advantages over the other techniques mentioned above, mainly the possibility of producing
fibers from micro- to nanometer scale with a large surface area [28–30]. Hence, the global interest in this field of research has gained
an ascending trend, such that many studies designed, fabricated and used electrospun nanofibrous scaffolds for different tissues
[31–34]. Inspired by a number of reliable reviews [35–38], the present review aims to revisit the current state of research and to
comprehensively summarizes the key studies on electrospinning in designing tissue-engineered scaffolds. The common techniques
used for evaluating the characteristics of the nanofibers were reviewed; besides, challenges associated with the usage of electrospun
fibers in the field were discussed.

2. General requirements for designing biomaterials

The American National Institute of Health (NIH) defines a biomaterial as “any substance or combination of substances, other than
drugs, synthetic or natural in origin, which can be used for any period of time, which augments or replaces partially or totally any
tissue, organ or function of the body, in order to maintain or improve the quality of life of the individual” [39]. When designing
biomaterial scaffolds as templates to stimulate the neo-tissue formation, a set of requirements should be considered including the
possibility of assessing how the host respond to the biomaterials (known as biocompatibility), physicochemical surface properties,
and the economic aspects related to biomaterials used in the clinical practice. Biocompatibility is defined as “the ability of a material
to perform with an appropriate host response in a specific situation” [40]. The response of the host to the biomaterials should be
evaluated both in vitro and in vivo. The implanted biomaterial should allow the stimulation of the host reactions known as foreign
body response (FBR), which commonly causes restricted in vivo functionality and durability of the biomaterials [41,42]. On the other
hand, this phenomenon essentially requires eliminating the cellular debris and controlling over infection [43]. In addition, resistance
arising from the FBR against biomaterials, e.g. featured by infiltration of macrophages, which may severely destruct the tissue [44].
Hence, recognizing the mechanisms activating or deteriorating the immune system are essential in designing biomaterials. An FBR is
traditionally described as a combination of protein adsorption as well as acute and chronic phases of the inflammation. The me­
chanism is initiated with protein adsorption and desorption (Vroman binding) at the surface of biomaterial after implantation. Based
on the physicochemical properties of biomaterial surface (such as porosity, size, surface charge, and roughness), proteins send signals
to the cells [45,46]. Monocytes subsequently differentiate into type 1 macrophages, which are responsible for the acute inflamma­
tion. After several days, type 1 macrophages differentiate into type 2 macrophages, which are in charge of chronic inflammation [47].
The chronic phase of inflammation is recognized by the presence of mononuclear cells, such as monocytes, plasma cells, and lym­
phocytes [47,48]. Chronic inflammatory responses are commonly characterized by the presence of macrophages, tissue granulation,
fibroblast infiltration, and neovascularization [49]. Granulation of the tissue may be considered as the precursor of fibrous capsule
formation, and is detached from the biomaterial surface by the cellular constituents of FBR [50]. However, this definition remains
invalid for systems in which the host responses belong to the recently developed degradable nanofibers used in tissue regeneration
applications [51–54]. Elaborated nanofibers are formulated to be responsible for a strong affinity to targeted cells making it possible
to stimulate biochemical signaling pathways toward the neo-tissue formation. This ability is entirely dependent on the specific
chemical characteristics of both the nanofibers and the biological environment of targeted tissue [46].
The biochemical perspective, different mechanotransduction, physiological, macromolecular adsorptions and biochemical sig­
naling pathways are designed and practised, which behave differently from tissues to cells [51,52]. Moreover, the biochemical cues of
innate and adaptive immune systems should be dealt with differently, for they respond to the nanofibrous scaffold implantation
individually [55].
The materials engineering point of view, the physicochemical properties of nanofibrous scaffold surface (such as topographical
features, stiffness, functional groups, and interfacial free energy) can strongly affect the biochemical mechanisms. For example, the
scaffold architecture should provide an appropriate cellular environment so as to make it possible to promote the neo-tissue for­
mation, remodeling, vascularization, and integration. The scaffold structure must be either porous or stable allowing for the diffusion
of nutrients and metabolites into the scaffold without the risk of collapse [56]. Designing nanofibrous scaffolds with optimal pore size
based on the target tissue and cells allows for cell migration into the scaffold and minimum ligand density on the scaffold surface [4].
Biodegradability is another key factor in designing smart scaffolds. A scaffold can only be considered as a support structure if the
body can gradually replace it with the ECM components. Waste products originating from scaffold degradation must be non-toxic and
removed without disturbing the surrounding organs [4].
The surface stiffness of nanofibers is also a key player in directing biochemical signaling pathways and cellular behavior such as
cell adhesion, spreading, migration, differentiation [57,58]. A balance between the various mechanical properties and the porosity
allows the scaffold to support sufficient infiltration and vascularization, in addition to providing the correct stability upon im­
plantation [56].
When fabricating a scaffold, considering factors involved in the manufacturing process is also of crucial importance to ensure that
producing the scaffold in large scales is feasible. Such factors include production complexity, cost effectiveness, suitable manu­
facturing processes, production rate, delivery methods, and scaffold storage. In terms of fabrication, the scaffold must be cost ef­
fective, with an easily attainable transition of the production from a small-scale aseptic laboratory procedure to a high-quality batch
production [17,19,59]. As previously mentioned, electrospinning is among the most useful techniques that is used for producing
promising nanofibrous scaffolds with above-mentioned criteria.

3
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

3. The background of electrospinning

Nanofibrous biomaterials take their origin in four main fields, namely, the biological applications, physics underlying Taylor cone
formation, material choices in electrospinning, and the impact of ambient conditions on the fabrication process. Nanofibers received
remarkable attention in biomedical engineering, especially in nanoscience and nanotechnology [60]. This relatively simple and
multipurpose method can make it possible to produce various fibers with beneficial features for regenerative medicine and tissue
engineering uses [61,62]. Formhals [63] reported using electrospinning technique for the first time in the 1930s [63]. However,
scientists and researchers alike paid attention to its applicability for biomedical applications five decades later [64]. This technique
benefits from the ability of applying a high electric field to produce ultra-fine polymeric fibers with micro- to nanometer diameters
[65,66]. A convoluted electro-physical activity between the polymer solution and the electrostatic force is the main mechanism
behind this technique [67]. In electrospinning, a high-voltage electric field is generated between the injection needle and the col­
lecting screen using a power supply and electrodes [68]. After gradually forcing out of the polymer solution, a hemispherical polymer
solution droplet is shaped at the tip of the needle [69]. This polymer droplet lengthens into a conical shape, known as the Taylor
cone, and the surface charge on the polymer droplet increases with time by increasing the voltage [70]. A polymer jet starts to form
immediately after overcoming the surface charge of the polymer droplet. After vaporizing the solvent in the polymer jet, the surface
charge on the jet increases, which destabilizes the polymer jet [70]. The polymer jet is geometrically segregated, initially into two jets
and, eventually, into a large number of jets, to compensate for the instability [71,72]. The electrostatic force, which affects the
constantly splitting polymer droplets causes the nanofiber patterning [70]. In addition, a spinneret with a metallic needle, a syringe
pump, a high-voltage power supply, and a grounded collector are the major constituents of a standard electrospinning system (Fig. 1)
[73,74]. The processing plasticity of the electrospinning facilitates producing various polymeric fibers [75].
Researchers could successfully electrospun natural and synthetic polymers, such as chitosan, collagen, gelatin, polycaprolactone
(PCL) and poly(lactic-co-glycolic) acid (PLGA), as biomimetic and temporal substrates to regulate cellular and molecular activities
[76,77]. Moreover, by combining synthetic polymers (as the backbone material) with natural polymers (on the surface for improving
the cell adhesion) composite fibrous scaffolds with optimal mechanical and compatibility properties could be produced for biome­
dical applications [78,79]. Doshi et al. [80] listed the internal and external parameters governing the structural morphology of the
electrospun nanofibers [80]. Environmental humidity and the temperature (as external parameters) as well as applied voltage,
working distance, conductivity, and viscosity of the polymer solution (as internal parameters) are the major factors [80,81]. In the
following sections, the effects of various parameters on the electrospinning efficacy are discussed in detail.

4. Parameters affecting the electrospinning efficacy

The diameter and the morphology of electrospun nanofibers are dependent on several parameters falling into three main cate­
gories: innate properties of solution, processing, and environmental factors [82,83]. The innate properties of the solution (such as
concentration, viscosity, molecular weight, electrical conductivity, elasticity, as well as polarity, and surface tension of the solvent)
have a remarkable impact on the morphology and the ultimate diameter of the electrospun nanofibers. The concentration of the
solution is a key factor governing the electrospinning process, so that only a small amount of solution is needed to run the elec­
trospinning device [84]. In addition, during electrospinning, ideal solution concentration is required to achieve smooth and uniform
nanofibrous biomaterials. When low-concentration solutions are used undesirable droplets may form, which could be caused by
surface tension effects [85]. However, at high concentrations, because of the high viscosity of the solution, the fiber construction
would be problematic. Additionally, the diameter of the fiber may increase with increasing the polymer concentration [86].
The viscosity of solution can also directly affect the fiber size and its morphology. Some investigators have demonstrated that
suitable viscosity is crucial for electrospinning, so that uniform and smooth fibers cannot be shaped in very low-viscosity solutions,
while a continuous jet makes producing fibers difficult in highly viscous solutions [85,87]. Changing the complex viscosity with ideal
solution concentration helps researchers to identify the best viscosity range suitable for the electrospinning. A polymer solution has

Fig. 1. A schematic representation of the electrospinning technique for designing fibrillar networks in synthetic scaffolds. Reprinted from [35], with
permission from Elsevier.

4
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

four regimes including i) dilute; ii) semidilute I; iii) semidilute II; and iv) concentrated solution. Under the dilute regime, polymer
chains in solution interact rarely with each other and the solution does not show viscoelastic behavior. Under semidilute I regime,
polymer concentration is higher than under the dilute regime, and the polymer chains interact at a critical concentration (C*), but
there is no major entanglement of polymer chains. Under semidilute II regime, polymer chains are entangled and exhibit a viscoe­
lastic behavior. Polymer chains entangle very tightly in concentrated solutions [88]. The entanglement is vital for fiber formation
during the electrospinning. The concentration at which entanglement takes place is called entanglement concentration, Ce. As a
practical rule, the solution concentration should be higher than the Ce [89,90]. The oval blue area in Fig. 2 is the appropriate region
for viscosity selection. In the overlapping concentration region, polymer chains start to affect the viscosity, which can be calculated
using intrinsic viscosity as C* = 1 [60]. For core-shell electrospinning, the value of ηcore/ηshell largely determines the quality of
electrospun fibers. The concentration should be kept high enough to stretch the core by viscous drag [89]. A polymeric droplet
experiencing such high voltage becomes highly electrified under the influence of two major electrostatic forces, namely, electrostatic
repulsion of surface charges and coulombic force exerted by the external field. Under the influence of such forces, the droplet shape
changes from spherical to the conical (Taylor cone) after the voltage reaches a critical value. The minimum voltage needed for
electrospinning can be estimated by Eq. (1):
VC2 = 4(D 2 / Le 2 )(Ln(2Le /R) 3/2)(0.117 R) (1)
where Vc is the critical voltage; D is the distance between the capillary and the collector; Le is the capillary length; R is the capillary
radius; and γ is the surface tension [91].
The molecular weight of polymer can also play a key role in fine-tuning the fiber dimensions, as indicated by a remarkable change
in the rheological behavior of solution during fabrication [92]. Uniform and smooth nanofibrous constructs can be obtained by
choosing polymers with appropriate molecular weights. In low-molecular weight polymer solutions, beads may appear instead of
fibers, while high-molecular weight polymer solutions result in fibers with relatively increased average diameter [92,93].
Furthermore, suitable surface tension, which is a function of solvent nature, can impact the electrospinning process and fiber
fabrication [94]. In a solution with high surface tension, the formation of fibers can be limited because of an unstable jet and
dispersion of droplets [95]. Lower surface tension can facilitate the electrospinning process at lower electric fields [96].
Fiber diameter first increases and then decreases slightly when the electrical conductivity of the solution is increased [85,97,98].
Yang et al. demonstrated that the electrical field distribution has major effects on the fiber diameter. The strength of the electric field
increases with increase of voltage at the nozzle, which causes decreasing the fiber diameter because of the prolonged jet path and an
increased bending frequency. A uniform electric field provides an appropriate electric field distribution, enabling the formation of
thinner fibers as a result of higher bending speed, which stretches the fibers [91].
Processing parameters (such as voltage, distance between the spinneret and the collector, and feeding rate of the polymer so­
lution) belong to another important category in the electrospinning process [82,83]. Electrospinning can generate fibrous scaffolds
only after overcoming the threshold voltage that causes substantial charge differences in solution during the process [99,100]. By
increasing the voltage and, subsequently, the charge value, the formation of droplets and beads in the fibers can be altered [101].
Another important factor with this regard is the flow rate of polymer solution. Decreasing the feed rate increases the time required
for solvent evaporation [102]. In electrospinning, a lower flow rate is usually applied to ensure complete evaporation of solvents from
nanofibrous scaffolds [103]. In addition, the distance between the tip of the syringe and the collector is a key factor that controls the
diameter of the round fibers and their morphology [104]. When the distance is small, the fibers do not have enough time to solidify
before reaching the collector; hence, fibers with larger average diameters will be formed. On the other hand, when the distance is
large, finer fibers may be formed [105]. Thus, choosing inappropriate solution concentration, applied voltage, and tip-to-collector
distance may result in bead formation.
The ambient parameters such as humidity and temperature of the environment when fabricating an electrospun sheet are also
vital, specifically when facing serious difficulties in obtaining uniform fibrous mats. High humidity is negatively correlated with the

Fig. 2. A schematic representation of four regimes suggested for polymer solutions: i) dilute; ii) semidilute I; iii) semidilute II; and iv) concentrated
solution. Regimes' characteristics are specified in Section 4.

5
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

solidification time, so that high humidity impedes the spinning process and prolongs charged jetting [106]. Some studies reported
that solvents can be completely removed by evaporation if the humidity is sufficiently low, while highly humid environments can
impair fibril formation [73].
Temperature is another factor affecting the morphology of the nanofibrous scaffolds. Two types of morphologies are observed
based on temperature differences: beads, which are formed at low temperatures, and condensed and flat fibers, which are formed at
high temperatures [107]. By increasing the temperature, the viscosity of the polymer solution decreases, leading to producing small-
diameter fibers [99]. Fig. 3 provides a phenomenological overview of electrospinning with main components, defined flow patterns,
and mass transfer equation governing the process. Table 2 also summarizes the effects of electrospinning parameters on the resultant
fiber morphology.

5. Electrospun nanofibrous scaffolds for biomedical applications

In the past few decades, many researchers have studied the applicability of different synthetic electrospun biopolymers for
biomedical applications to satisfy clinical needs [20,30,31,37,109]. Members of poly(α-hydroxy acid) family including glycolic acids,
lactic acids, and their copolymers with ε-caprolactone, are a key class of biodegradable polymers with a promising potential in the
biomedical field [110–112]. These biodegradable polymers can be degraded into nontoxic end-products by hydrolysis, and regulated
by some properties, including the morphology and the average molecular weight of these compounds [113]. Hence, by regulating the
electrospinning parameters, any biodegradable and biocompatible polymer can be electrospun for biomedical applications [113]. In
addition, electrospun nanofibers can potentially direct cellular and molecular responses after implantation [114–117]. For example,
Neves et al. [118] fabricated various fiber meshes based on poly(ethylene oxide) (PEO) and PCL, employing differently patterned
collectors with certain dimensions and forms to assess the resulting mesh features for biomedical applications (Fig. 4) [118]. Elec­
trospinning technique has gained considerable attention of researchers and scientists in many areas of biomedical applications,
including regenerative medicine and agent delivery systems, which are described in detail below.

5.1. Electrospun nanofibrous scaffolds for tissue engineering applications

Tissue engineering, also known as regenerative medicine, is an emerging integrative field of study which benefits from the
principles of medicine, biology, and engineering fields to develop biological substitutes that restore, maintain, or improve tissue
functions [119,120]. Tissue engineering requires a scaffold that supports cells, regenerates ECM components, and/or provides a
vector to delivery biochemical factors [4]. The nanoscale structure of the native ECM-containing network of proteins and glycosa­
minoglycans (GAGs) forms a boundary between tissues and a supportive meshwork around the cells to provide cell anchorage [121].

Fig. 3. A phenomenological overview of electrospinning, with the main components, defined flow patterns, and mass transfer equation that govern
the process.

6
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

Table 2
Effects of electrospinning parameters on the morphology of electrospun fibers [108]
Parameter Effect of parameter on fiber morphology

Applied voltage ↑ Fiber diameter ↓ initially, then ↑ (not monotonic)


Flow rate ↑ Fiber diameter ↑ (beaded morphologies occur if the flow rate is too high)
Distance between capillary and collector ↑ Fiber diameter ↓ (beaded morphologies occur if the distance between the capillary and collector is too short)
Polymer concentration (viscosity) ↑ Fiber diameter ↑ (within optimal range)
Solution conductivity ↑ At first fiber diameter ↑ then ↓ (broad diameter distribution)
Solvent volatility ↑

Fig. 4. a) SEM micrographs of meshes obtained using PEO (top) and PCL (bottom); b) PCL deposited on top of a flat copper collector (top) and in the
area around the flat copper plate (bottom); c) PCL fiber meshes deposited on the screw collector: close-up showing the mesh region corresponding to
the thread crest (top) and the region between two consecutive threads of the screw collector (bottom). Reprinted from [118], with permission from
Elsevier.

Hence, many researchers focus on designing scaffolds with similar properties to those of human tissue at the nanoscale level. Several
studies reported that using electrospinning loosely connected three-dimensional (3D) porous nanofibrous constructs with large
surface area can be produced to resemble the native ECM network [122–125]. For example, our research team designed novel
electrospun nanofibrous scaffolds by electrospinning various concentration of Bombyx mori silk fibroin solutions (10, 12, and 14%

Fig. 5. Morphology of normal human osteoblasts (NHOst) cells cultured for seven days on (a) PCL/osteo/gelatin/CaP and (b) PCL/osteo/gelatin/
CaP/PANI scaffolds. Reprinted from [142], with the permission from Elsevier.

7
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

w/v in formic acid), and evaluated their applicability for tissue engineering in vitro and in vivo [126]. The developed constructs allow
good cell adhesion and growth, with no adverse effect on cell viability or cytotoxicity. In addition, the nanofibers could not stimulate
foreign body responses; however, could enhance the total cell number in the implantation area. In the following sections, we discuss
the applications of electrospun nanofibrous scaffolds for engineering different tissues in detail.

5.1.1. Electrospun nanofibrous scaffolds for bone tissue engineering


Bone tissue engineering represents a dynamic strategy for designing scaffolds to deliver therapeutic agents and cells to the
damaged tissue for stimulating the neo-tissue formation [127–129]. Researchers should consider several critical factors when de­
signing bone scaffolds including 1) the porosity size; 2) suitable mechanical strength and tunable biodegradation kinetics; 3) in­
terrelated open porosity for growth factors; 4) sterile environment for cell growth and cell seeding [130–132]; and 5) scaffold
biocompatibility and biodegradability [128,133]. Using electrospinning many of the above-mentioned criteria can be achieved.
Electrospun nanofibrous scaffolds from various natural and synthetic polymers are used for bone engineering such as alginate,
chitosan, collagen, PCL, polyglycolic acid (PGA), PLA, and PLGA [28,128,134–136].
Hydroxyapatite (HA), which has a similar chemical structure to the minerals in the native bone, is one of the most important
biomaterials used for bone engineering applications [137,138]. HA can improve topographical properties of nanofiber surface to
achieve better cell adhesion and growth [139–141]. Rajzer et al. [142] established a new method for inkjet printing to generate

Fig. 6. A,B) SEM micrograph of hMSCs seeded on scaffolds, after 14 d [143]. C, D) Random and aligned electrospun nanofibers [poly(L-lactic acid)-
co-poly-(ε-caprolactone) (PLCL) synthetic biopolymer] influencing the direction of cellular spreading and elongation [144]. E) Core-shell structure
of silicon nanoparticles in interconnected hollow carbon fibers. F) Carbon matrix. Reprinted from [145], with the permission from Elsevier.

8
Table 3
An overview of key electrospun nanofibers for bone tissue engineering applications.
Materials composition Fibers diameter (nm) Studies Cells/ enzymes Voltage (kV) Flow rate Needle–mandrel Needle Major outcomes Ref
(ml.h−1) distance (cm) size
M. Rahmati, et al.

(gauge)

nMP/PCL/HPG/nHA 725 ± 153, 783 ± 125, In vitro MG63 & hMSCs 9–13 500 µl/h−1 11 N/A Increasing swelling, biomineralization, [146]
535 ± 113 and 608 ± 120 nm for and breakages of fibers by adding nMP.
PCL, PCL/nHA, PHAMP, and Increasing cell viability, adhesion and
PGHAMP respectively. spreading by adding nMP.
miRNAs/PCL 538.574 In vitro iPSCs 16 0.4 18 N/A Increasing the expression of osteogenic [147]
markers by adding miRNAs.
PCL/CTS/Mg/HA 419–495 In vitro MG63 14–22 0.3&0.4 14 N/A Improved Cell viability and [148]
proliferation.
Improving bone mineralization through
introducing triazole rings on the
copolymer structure.
SPI/PEO 701.92 ± 315 In vitro rBMSC 23 9 µl/min 13 23 Neo-tissue formation can be detected [149]
for SPI/PEO electrospun nanofibers
with 7:3 concentration.
Improving bone formation in the
presence of rBMSC.
CNT 500 In vitro, In MSCs 13 0.007 ml/min 8 23 Coating nanofibers with CNT stimulates [150]
vivo the expression of angiogenic and bone
formation.
Collagen/ Catecholamines, 900 ± 150 In vitro hFob 17 0.8 13 27 The scaffolds has high mechanical [151]

9
Ca2+ properties with Young’s modulus close
to cancellous bone.
Improved cellular responses.
PLGA/HA1%/MWNTs 1055 In vitro BMSCs 0–45 0.2 13 N/A Depending average diameter of fibers [152]
on HA concentration.
Improved cellular responses.
CTS/HA/GP 335 ± 119 In vitro 7F2 15 1.2 15 N/A Increasing Young’s modulus by [153]
osteoblast-like crosslinking with genipin.
cells, ALP Increasing the osteoinductive
bioactivity.
Suitable for non-weight bearing bone
tissue engineering.
SF 411 ± 98 In vitro MC3T3-E1 12 N/A 10 22 3D NFS with high porosity has more [154]
cell adhesion and proliferation than 2D
NFS.
HA/PLA 1–2 µm In vitro MG63, ALP 25 0.1 15 N/A The surfactant-mediation method is [155]
operational in scattering the HA nano-
powder in the PLA Matrix.
Improved osteoblastic cellular
responses.
(continued on next page)
Progress in Materials Science xxx (xxxx) xxxx
Table 3 (continued)

Materials composition Fibers diameter (nm) Studies Cells/ enzymes Voltage (kV) Flow rate Needle–mandrel Needle Major outcomes Ref
(ml.h−1) distance (cm) size
(gauge)
M. Rahmati, et al.

HA/PLGA 266.6 ± 7.3 In vitro MC3T3-E1, ALP 12 1 15 N/A Equally spreading HA on the fiber [156]
surface at a minor concentration.
Slowly increasing agglomerates by
increasing HA concentration.
Improved biomineralization.
A significant increase of cell viability
until day 7, and then decreasing to day
14.
Considerably high ALP activity in the
scaffolds.
Silk/PEO/nHA/BMP-2 520 ± 55 In vitro hMSCs 12 0.02 ml/min 21.5 N/A Significantly increasing bone formation [143]
through using BMP-2 and/or nHA into
the scaffolds.
PCL/Gel/calcium 1.2 µm In vitro MC3T3-E1 15 0.5 15 24 Improved homogeneous calcium [157]
phosphate phosphate coating in the presence of
gelatin.
Covering the fibers by a thin layer of
mineral deposition after two hours of
incubation.
The structure of the scaffold was a
mixture of dicalcium phosphate

10
dehydrate and apatite.
Improved cell proliferation.
The formation of a multilayered film of
cells on the scaffolds after seven days of
culture.
Improved cytocompatiblity in the
presence of gelatin and calcium
phosphate coatings.
CTS/HA 214 ± 25 In vitro hFOBs 17.5 20 µl/min 34 N/A Remaining the crystalline nature of HA [158]
and surviving the AA-dominant solvent
system.
Remarkable bone formation oriented
through using HA nanoparticles.
Collagen/PLGA/nHA Collagen = 353.8, In vitro N/A 12 1 12 N/A The surface functional groups [159]
Thin PLGA = 97.3, significantly affect the mineral
Thick PLGA = 320 formation.
Bonelike apatite formation is much
copious and constant over collagen
nanofibers than PLGA.
PCL/GT 312 In vitro BMSCs 20 2 18 N/A Inducing the BMSCs recruitment after [160]
In vivo adding SDF-1α to the fibers.

(continued on next page)


Progress in Materials Science xxx (xxxx) xxxx
Table 3 (continued)

Materials composition Fibers diameter (nm) Studies Cells/ enzymes Voltage (kV) Flow rate Needle–mandrel Needle Major outcomes Ref
(ml.h−1) distance (cm) size
(gauge)
M. Rahmati, et al.

AP-g-GA/PLLA 200 In vitro MC3T3-E1 1.5 30–4 µl/min 24 N/A Improved thermal stability, [161]
biodegradation and biocompatibility.
Improved electroactivity and reversible
redox properties.
Improved cell differentiation.
Submicron bioactive glass 50–800 In vitro N/A 7–19 0.5 15 N/A The scaffolds have large surface area, [162]
fibers 70S30C high porosity and fine interconnected
pore network construction.
Improved elastic modulus.
PEOT/PBT/calcium 6.47 ± 1.46 µm In vitro, In hMSCs 12 15,20,25 15 20 The scaffolds could not display any [163]
phosphate vivo considerable change in ALP expression
when compared to uncoated scaffolds.
Stimulating osteogenesis.
PLLA/PBLG/Collagen 200–400 In vitro ADSCs 12 1 12 27 Introducing PBLG/n-HA on the [164]
polymeric nanofibers to adjust and
enhance specific biological functions of
ADSC into osteogenic lineage.
CTS/GP 144–154 In vitro SAOS-2 25 20 µl/min 15 20 Increasing the mechanical properties, [165]
preserve nanofibrous morphology, and
cellular proliferation using genipin.
The mats have adequate mechanical

11
properties, degradation rate and
cytocompatibility for GBR applications.
PHBV/nHA/ Bombyx mori 10–15 μm In vitro HOB 10,15 2,5 15 N/A Increasing fiber diameters and [166]
SF decreasing wettability by increasing
concentration of the co-phases within
the polymeric matrix.
Increasing Young’s modulus for
composites containing 2% wt of
ceramic and 2%wt of SF.
Decreasing Young’s modulus by
increasing the ceramic and proteic
phase’s concentrations to 5% wt.
A relative decreasing in compressive
secant modulus by increasing the
ceramic content.
The scaffolds were bioactive,
supporting bone like apatite crystals
growth after 28 days.
An appropriate topography mimicking
the ECM in the presence of HA.
Capability of the cells to infiltrate into
the scaffolds.
(continued on next page)
Progress in Materials Science xxx (xxxx) xxxx
Table 3 (continued)

Materials composition Fibers diameter (nm) Studies Cells/ enzymes Voltage (kV) Flow rate Needle–mandrel Needle Major outcomes Ref
(ml.h−1) distance (cm) size
(gauge)
M. Rahmati, et al.

SF 200–400 In vitro, MC3T3-E1, ALP 13 N/A 20 22 Considerably increasing proliferation [167]


In vivo and ALP activity of osteoblasts.
Increasing expression levels of
activated adhesion-related proteins in
the P-ESF.
Improved new bone formation.
PLLA/collagen/HA 310 ± 125 In vitro hFOB, ALP 10,12 N/A 15 N/A Improved mineral deposition on [168]
scaffolds.
Providing cell recognition sites and HA
for cell proliferation and
osteoconduction in the presence of an
ECM protein, collagen and HA.
GTl/PCl/calcium 0.3–5 μm In vitro NHOst, ALP 30 1.5 15 N/A Improved mechanical properties of [169]
phosphate scaffolds.
Increasing the nucleation and growth of
apatite crystals on the surface in the
presence of calcium phosphate
nanoparticles.
Improved ALP activity and
mineralization.
PCL/SrBG 46.1 ± 16.6 μm In vitro MC3T3-E1 7 20 μl/h 6 19 Improving osteogenesis by adding [170]

12
SrBG.
Increasing ALP activity in MC3T3-E1
cells by adding SrBG.
PHB/HA 2.0 ± 0.2 μm In vitro preosteoblasts 10–30 2 20 N/A Increasing the fiber diameter by [171]
increasing the applied voltage and
solution concentration.
Positively encapsulating of NPs inside
the ultrafine fibers fabricated.
Improved cell viability and spreading
on the fibrous nanohybrids, and cell
metabolic activity by increasing
incubation time.
PHBV/PRP/nSrCO3 400–800 In vitro hMSCs, ALP 15 0.5 15 N/A Improving the osteogenic [172]
differentiation of hMSCs.
PLGA/PCL 0.36–2.4 μm In vitro BMSCs 13 3 12 N/A Affecting the morphology and [173]
orientation of cells by different
alignments of random PLGA and
aligned PCL fiber regions.
PCL//AV/SF/HA 133 ± 28 In vitro ADSC, ALP 13 1 10 21 ADSCs are appropriate cell therapeutics [174]
for bone regeneration.
The composite scaffold improves bone
regeneration.
(continued on next page)
Progress in Materials Science xxx (xxxx) xxxx
Table 3 (continued)

Materials composition Fibers diameter (nm) Studies Cells/ enzymes Voltage (kV) Flow rate Needle–mandrel Needle Major outcomes Ref
(ml.h−1) distance (cm) size
(gauge)
M. Rahmati, et al.

PCL/PLGA 1.98 ± 0.51 μm In vitro, RBMCs 20 50 µl/min 15 18 Improved cell infiltration and cartilage [175]
In vivo matrix deposition.
Improved new bone formation over the
remodeling of the cartilage template
after 8 weeks of in vivo implantation.
HA/SF 350 ± 67 and 112 ± 89 μm In vitro, BMSCs 40 N/A N/A N/A Decreasing average pore diameters by [176]
In vivo increasing the SF concentration from
5% to 10%.
HA/SF-5% scaffold is a promising
candidate for improving cell adhesion
and biocompatibility of surface.

13
Progress in Materials Science xxx (xxxx) xxxx
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

conductive patterns on electrospun scaffolds for bone tissue engineering. The authors printed conductive polyaniline (PANI) mi­
cropatterns on PCL/osteo/gelatin/CaP scaffolds using an inkjet printer and demonstrated that calcium phosphate nanoparticles and
osteogenon could improve the bioactivity of hybrid scaffolds. Moreover, the morphology and proliferation of normal human os­
teoblasts (NHOst) on PANI-modified surfaces improved compared to non-modified surfaces (Fig. 5) [142].
Li et al. [143] fabricated nanofibrous scaffolds containing silk fibroin, bone morphogenetic protein 2 (BMP-2), and HA nano­
particles (nHA) [143]. After culturing human mesenchymal stem cells (hMSCs) for up to 31 days under static conditions on scaffolds
(silk/PEO/BMP-2, silk/PEO/nHA, and silk/PEO/nHA/BMP-2) and in control solutions (silk/PEO, silk/PEO extracted), the scaffolds
supported hMSCs growth and differentiation towards osteogenesis. The nHA particles were integrated into the scaffolds during
processing and enhanced bone formation. As depicted in the SEM micrographs shown in Fig. 6A and Fig. 6B, after 14 days cell
culture, the progression of cell growth on scaffolds in silk/PEO scaffolds was significantly lower than that of silk/PEO, PEO extracted
scaffolds. This could be because of the presence of PEO, which reduces the porosity of scaffold [143]. It was also reported elsewhere
that the alignment of grown cells largely depends on the randomness of the electrospun fibers. Comparing Fig. 6C and Fig. 6D, it
becomes apparent that highly oriented fibers (Fig. 6D) induce the stem cell growth in the direction of fiber alignment [144]. The fiber
geometry can also be adjusted by fine-tuning the electrospinning process, as illustrated in Fig. 6E and Fig. 6F. Table 3 provides a
comprehensive comparison of different types of electrospun fibers applied in bone tissue engineering.

5.1.2. Electrospun nanofibrous scaffolds for cartilage tissue engineering


Articular cartilage tissue is composed of tangential, transitional, radial, and calcified areas (specific cells and ECM), which reduce
friction and increase wear resistance of tissue [177]. Cartilage defects are among the most common crucial health issues [178,179].
Currently, an electrospinning technique that can be used to mimic the pattern features of the ECM has gained considerable attention
of scientists for applications in cartilage regeneration [180]. Wise et al. [181] developed electrospun oriented nano- and micro-fibrous
PCL scaffolds seeded with hMSCs for articular cartilage regeneration [181]. Their results indicated that an aligned ECM environment
for modulating cartilage arrangement could be achieved using this technique. Xue et al. [182] designed a novel electrospun fibrous
membrane based on GT/PCL, to engineer cartilage with the exact 3D network in a sandwich model. The authors fabricated a rounded
membrane and seeded it with chondrocytes. Macroscopic and histological analyses confirmed the possibility of cartilage regeneration
using these membranes (Fig. 7). Using an ear-shaped titanium alloy mold, the authors also created an ear-shaped cartilage in a
sandwich model. The in vitro and in vivo evaluations exhibited that the cartilage preserved its original shape, sharing up to 91.41%
similarity with the titanium molds [182].
Li et al. [183] designed some PCL nanofibrous scaffolds and then investigated their ability to promote in vitro chondrogenesis of
MSCs for cartilage regeneration. Their results indicated that the nanofibers were uniform and uniformly oriented, with a diameter of

Fig. 7. Engineering of ear-shaped cartilage. A) Tailored electrospun GT/PCL membrane; B) titanium alloy ear-shaped mold; C) gross view of ear-
shaped cell-scaffolds just after stacking (0 h); D) cell-scaffolds 2 weeks after in vitro culturing; E) subcutaneous implantation in nude mice; F–H)
engineered ear-shaped cartilage after 6 weeks of in vivo incubation; and I) similarity testing of the engineered ear and the titanium mold. Reprinted
from [182], with the permission from Elsevier.

14
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

Fig. 8. Immunohistochemical localization of cartilage-specific ECM molecules in day-21 nanofibrous fibroin scaffold (NFS) MSC cultures treated or
untreated with TGF-β1. Positive staining of collagen type II (A), cartilage proteoglycan link protein (B), and aggrecan (C) was observed in TGF-
β1–treated cultures but not in cultures without TGF-β1 exposure (D–F). Bar, 20 μm. Reprinted from [183], with permission from Elsevier.

700 nm. MSCs cultured in the scaffolds differentiated into cells with a chondrocytic phenotype, in the presence of transforming
growth factor beta (TGF-β)-1 (Fig. 8) [183]. Table 4 provides an overview of major nanofibrous scaffolds used for cartilage tissue
engineering applications.

5.1.3. Electrospun nanofibrous scaffolds for vascular tissue engineering


Vascular tissue engineering endeavors to reconstruct blood vessels consisting of endothelial and perivascular cells for clinical
applications [203]. High porosity and aspect ratio of nanofibers enhance nutrient and gaseous exchanges that lead to angiogenesis,
which is the main factor of vascular regeneration [204]. Blood vessels have three layers: the tunica (outer coat) intima, tunica media,
and tunica adventitia. The tunica intima is the innermost lining of a monolayer of non-thrombogenic endothelial cells, which can
induce platelet stimulation and thrombus formation. The second layer is the tunica media, which consists of a great number of
smooth muscle cells. The outermost layer, tunica adventitia, is composed of collagenous ECM and fibroblasts [205,206]. ECM is the
most important component of the vascular system; hence, the tensile stiffness, elasticity, and compressibility of a blood vessel are the
crucial features in designing vascular scaffolds [207]. The advances in nanotechnology ensure the capability of using electrospinning
not only for developing scaffolds but also for creating tubular scaffolds for vascular tissue engineering. Yazdanpanah et al. [208] used
electrospinning to fabricate nanofibrous scaffolds based on poly(L-lactic acid) (PLLA) and gelatin [208]. Tensile tests indicated that
the mechanical strength and estimated burst pressure of graded PLLA/gelatin scaffolds are better than those of layered PLLA/gelatin
and gelatin scaffolds.
Researchers seeded human aortic endothelial cells (HAECs) and human coronary artery smooth muscle cells (HCASMCs) onto
electrospun silk fibroin scaffolds [209]. Fig. 9 shows that HCASMC is a stromal support cell for HAEC, which undergoes alignment/
prolongation, while HAEC is a cell that forms cord-like network. Fig. 9 represents the appropriate alignment and prolongation of
HCASMCs on scaffolds five days after seeding, as well as the formation of enhanced cord-like networks on the surface. Because of
their superior mechanical properties, blood vessels require scaffolds with appropriate mechanical strength. Researchers suggest
developing multi-layer scaffolds for vascular tissue engineering to mimic the tissue while assuring the substrate has mechanical
strength in the range of targeted tissue [210–212]. Table 5 provides an overview of key nanofibrous scaffolds for vascular re­
generation applications.

5.1.4. Electrospun nanofibrous scaffolds for cardiac tissue engineering


Cardiac tissue scaffolds should have high conductivity and elasticity to mimic cardiac functions (Fig. 10). Wang et al. [229]
fabricated conductive nanofibers based on PANI and PLA for cardiac regeneration. The nanofibers exhibited appropriate bio­
compatibility, and enhanced cell-cell interactions and spontaneous beating of primary cardiomyocytes [229]. Further, Dippold et al.
[230] designed elastic nanofibers based on poly(glycerol sebacate) (PGS)-zein for cardiac engineering. Their results indicated that
zein enhance the mechanical properties of PGS, and the fibers exhibit appropriate biocompatibility as well as durability for cardiac
engineering applications [230]. In addition, researchers suggested using elastin to increase elasticity of nanofibers used in cardiac
tissue engineering [231,232].
Zong et al. [233] developed biodegradable non-woven PLGA-based scaffolds for cardiac tissue engineering using electrospinning

15
Table 4
An overview of key electrospun nanofibers for cartilage tissue engineering applications.
Materials composition Fibers diameter (nm) Studies Cells Voltage (kV) Flow rate Needle–mandrel Needle Major outcomes Ref
(ml.h−1) distance (cm) size
M. Rahmati, et al.

(gauge)

GAS/PCL 0.573–1.318 In vitro ACs N/A N/A N/A N/A Coaxial electrospinning can be used to [184]
load GAS in nanofibers.
Improving cell proliferation and growth
by adding GAS.
PCL/GEL/PEG 100–300 In vitro CP5 27 1.5 9 21 Improving cell attachment and [185]
proliferation by adding gelatin.
Increasing pore size and interconnectivity
of fibers by adding gelating.
PGS/PCL/KGN 617–738 In vitro hBMSC 20 50 & 180 μl/ 17 N/A Improving cell proliferation and [186]
min chondrogenic differentiation by adding
KGN into nanofibers.
PCL 0.4–1.4 μm In vitro hBMSCs 9 1 20 N/A The formation of new ECM components [187]
on the scaffolds.
PCL/CDM 0.56–0.58 μm In vitro hASCs 17, 25 1.2 20 21, 25 Stimulating sulfated GAGs synthesis and [188]
COL10A1 gene expression.
Higher cell infiltration and ACAN gene
expression in multilayer scaffolds than
single-layer constructs.
Lower elastic modulus in multilayer
scaffolds than single-layer constructs.

16
Improving homogeneous cell seeding,
and chondrogenesis-related bioactivity in
multilayer scaffolds.
Collagen/PLCL 295 ± 103 In vitro, Chondrocyte 16 1 12 N/A Stimulating more cartilage-like tissue [189]
In vivo with the extension of implantation time in
vivo.
Significantly increasing Young’s modulus
of the scaffolds by PLCL.
3,4,6-O-Bu3GlcNAc–loaded Thick fibers = 200 ± 20, Thin In vitro, Chondrocyte 11, 12 0.1, 0.4 11 22 3,4,6-O-Bu3GlcNAc not only hinders NF- [190]
PLGA fibers = 20 ± 2 μm In vivo B signaling, but also increases
chondrogenic and anti-inflammatory
properties on OA chondrocytes.
Stimulating cartilage tissue production in
3D in vitro hydrogel culture networks.
Pμ, PμPn, PμFn Fibers of Pμ = 10.1 ± 1.1 μm In vitro hMSCs 7,13,18 0.1, 0.4, 8 N/A N/A Preserving scaffold cellularity under [191]
Fibers of serum-free conditions and improving the
PμPn = 9.8 ± 1.4 μm, deposition of GAGs by nanofibers within
194.6 ± 45.9 nm a microfiber mesh.
Fibers of
PμFn = 8.8 ± 1.1 μm,
250 ± 50.2 nm
PEOT/PBT/PEG 268 ± 32 μm for 3DFESP, In vitro Primary bovine 15 0.39 15 N/A Providing structural integrity and [192]
10 ± 2.8 μm for 3DF articular mechanical properties by 3DF scaffold.
chondrocyte The ESP system acts as a “sieving” and
cell entrapment system and proposes
signals at the ECM scale.
Progress in Materials Science xxx (xxxx) xxxx

(continued on next page)


Table 4 (continued)

Materials composition Fibers diameter (nm) Studies Cells Voltage (kV) Flow rate Needle–mandrel Needle Major outcomes Ref
(ml.h−1) distance (cm) size
(gauge)
M. Rahmati, et al.

The scaffolds improved cell entrapment,


an upper amount of ECM, and a
meaningfully greater GAG/DNA ratio
after 28 days.
A direct effect of fiber dimensions on cell
differentiation.
PLGA/HA/Collagen type I 421 ± 208 In vitro hMSCs 14.4 ± 0.4 1 14 N/A Improving the cell viability. [193]
Hypertrophic chondrocytes in the
scaffolds.
Collagen type I & II N/A In vitro hMSCs 13–30 0.3–1.5 N/A N/A Facilitating the differentiation of hMSCs [194]
by creating a biocompatible and
hydrophilic, chondroinductive scaffold
consisting of two types of collagen.
CTS/PVA/CaCO3 71.5–140.7 In vitro ATDC5 17 10 µl/min 15 N/A Increasing the diameter of fibers and [195]
Young’s modulus by increasing the
concentration of CaCO3.
CDPS/PLA N/A In vitro BHK-21 12 1, 4 10 N/A Improved biomechanical and [196]
hemocompatibility properties compared
to natural tissues.
PVA/PCL/BMSC 300–800 In vitro, MSCs 15 N/A N/A 17 Improving cell proliferation and [197]
In vivo chondrogenic differentiation in the

17
scaffolds.
Improving healing of defects by treating
with cell-seeded PVA/PCL scaffolds.
PLGA 0.79–0.92 In vitro L-929 0.56 N/A 25 18 The scaffold is mechanically stable. [198]
Faster degradation of the nanofiber
scaffold than a block-type scaffold.
Degradation of scaffold was Dependent
on the lactic acid/glycolic acid
concentration ratio and may be organized
by mixing ratio of blend PLGA.
increasing chondrocyte proliferation and
ECM formation by applying
Intermittent hydrostatic pressure to cell-
seeded scaffolds.
CTS/PEO N/A In vitro Chondrocyte N/A N/A 12–29 20 The fibers are aligned. [199]
Significantly higher elastic modulus,
Young’s modulus, and good chondrocyte
biocompatibility in the electrospun mats
than those of the cast films.
Slightly higher cell viability in the
electrospun mats than the cast films.
Solid cylindrical and 200–800 In vitro, N/A N/A N/A N/A N/A Fibrous tissue at the articular surface of [200]
cannulated tubular types of In vivo the scaffold two weeks postoperative.
PLGA The formation of cartilage at the articular
surface and bone at the subchondral area
and retained over postoperative week 24.
Progress in Materials Science xxx (xxxx) xxxx

(continued on next page)


Table 4 (continued)

Materials composition Fibers diameter (nm) Studies Cells Voltage (kV) Flow rate Needle–mandrel Needle Major outcomes Ref
(ml.h−1) distance (cm) size
(gauge)
M. Rahmati, et al.

Considerably higher histologic scores in


the treated groups with the scaffolds than
those of control group.
PLGA/MSM 0.65–84 μm In vitro Chondrocyte 40 0.07–0.10 µl/ 20 N/A Improving cell proliferation and ECM [201]
min formation ability of the scaffolds.
Developing ECM formation, the cartilage
related gene expression of collagen type
II, aggrecan, and collagen type I, and
cartilage specific protein expression of
collagen type II in the 10 wt% MSM/
PLGA mats.
PLCL/CTS-QK 194 ± 65 In vitro HUVECs 19 0.4 16 N/A The scaffolds could effectively [202]
encapsulate QK peptide and keep its
secondary structure after release.
Accelerating the proliferation of HUVECs
by the release of QK peptide after nine
days.
GT/PCL 440 ± 63 In vitro, Chondrocyte 10 2 12 18 Using an ear-shaped titanium alloy mold, [182]
In vivo an ear-shaped cartilage can be created in
the sandwich typical.
The ear-shaped cartilage mostly preserves

18
their original form, after two weeks of
culture in vitro and six weeks of
subcutaneous incubation in vivo.
The engineered cartilage has good
elasticity and mechanical strength.
Progress in Materials Science xxx (xxxx) xxxx
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

Fig. 9. Immunocytochemistry staining of HAECs on scaffolds on days 1, 5, 10, and 15. Reprinted from [209], with permission from Elsevier.

[233]. The scaffolds could improve the isotropic and anisotropic growth of cardiomyocytes [233]. Shin et al. [234] designed some
CM-electrospun nanofibers based on biodegradable PCL, by seeding neonatal rat CMs on electrospun meshes. The authors observed
good adherence of cultured CMs on scaffolds, with five even layers of cells without any core ischemia or synchronized beating
(Fig. 11) [234]. Castellano et al. [235] synthesized a number of different scaffolds based on polyhydroxybutyrate (PHB), PCL, silk,
PLA, collagen, and polyamide (PA), to compare their properties as electrospun nanofibrous scaffolds for cardiac regeneration. The
study demonstrated that PHB and PCL nanofibrous scaffolds could enable the maximum adhesion/growth of MSCs, cardiomyocytes,
and cardiac fibroblasts. Moreover, the in vivo studies exhibited that implantation of PCL, silk, PLA, and PA patches at the epicardial
surface of healthy rats could induce a traditional FBR, while collagen and PHB patches were mainly degraded. Although PHB-based
scaffold meaningfully induces angiogenesis, collagen, PCL, and PHB generally reduce tissue remodeling after implantation [235].
Table 6 provides a summary of major nanofibrous scaffolds used for cardiac engineering applications.

5.1.5. Electrospun nanofibrous scaffolds for nerve tissue engineering


Peripheral nerve injury is among the most common diseases in the world that may occur during traffic accidents, resections of
tumors and/or adverse iatrogenic effects of the surgery [245–247]. Over 500,000 cases of peripheral nerve injuries are annually
reported in Europe and the US [248]. An ideal scaffold for nerve tissue engineering should exhibit specific characteristics, such as
appropriate biological and physiochemical properties, excellent biocompatibility, biodegradability, oxygen and nutrient perme­
ability, good mechanical characteristics, and suitable surface features [247,249]. Hence, researchers are paying increasing attention
to electrospinning as a technique for designing synthetic and natural nanofibrous scaffolds suitable for nerve tissue engineering
[250–252]. Our research team designed a novel laminin-incorporating PLCL scaffold for nerve engineering applications using co-axial
electrospinning and blend electrospinning approaches. We observed 78% increase of Schwann cell proliferation on the core-shell-
structured nanofibers after seven days in vitro culture [253]. Because of the neural action potential, conductive scaffolds are suitable
as neural substrates [254–256]. Tian et al. [257] electrospun conductively aligned nanofibers based on PLA and polypyrrole for
neural tissue engineering, in which aligned fibers could support the cell growth, proliferation, and neurite outgrowth than random
fibers. Neurite outgrowth was also enhanced by external electrical stimulation (40 mV) [257]. Fig. 12 provides a schematic re­
presentation of bio-mimicking scaffold for peripheral nerve regeneration. Researchers used polymerized polypyrrole on PCL-PLA
electrospun fibers as a conductive nanofibrous conduit. Upon electrical stimulation, the performance of the conductive conduit is
significantly higher than that of an autografted non-stimulated PPY/PLCL conduit group [258]. Wang et al. [259] studied the effects
of fiber diameter on neural cell growth. The authors prepared three types of fibers with different diameters (1325-, 759-, and 293-
nm). The results indicated that neurite length on the intermediate-size substrate is greater than that of the other two substrates.

19
Table 5
An overview of key electrospun nanofibers for vascular tissue engineering applications.
Materials Fibers diameter (nm) Studies Cells Voltage (kV) Flow rate Needle–mandrel Needle Major outcomes Ref
composition (ml.h−1) distance (cm) size
M. Rahmati, et al.

(gauge)

PU/PEG 394 ± 106 In vitro HUVECs 20 0.6 15 21 The mechanical properties of the scaffolds are [213]
close to those of human and pig arteries.
Improved hemocompatibility by adding PEG.
Improved cell attachment and proliferation on
scaffolds.
LLA-TMC 2–5 μm In vitro SMCs 15 1 20 30 A middle TMC amount has the best [214]
electrospinning behavior.
The scaffolds could tolerate repetitive cyclic
loading, without meaningfully dropping
performance over the period of time tested.
Matching the mechanical properties of the
scaffolds to the stated range of human arteries.
A steady loss of modulus and strength by Bulk
degradation of scaffolds.
Supporting the cell growth and proliferation.
GT/Heparin 814 ± 201 for heparin 1% In vitro HUVECs 18 0.6 20 25 Decreasing in fiber diameter by increasing the [74]
heparin weight ratio.
Improving the tensile strength and elastic
modulus of crosslinked scaffolds.
A sustained release of heparin over 14 days from

20
the scaffolds.
Increasing the cell growth and proliferation by
adding the crosslinker.
Elastin/PU/Collagen 557.17 ± 113.33 In vitro SMCs 15–17 1, 2 10–12 N/A The scaffolds attain viscoelastic properties in the [215]
presence of elastin.
Improving cell proliferation in the presence of
collagen.
Supporting the cell growth and proliferation.
PCL/Collagen type I/ 4.85 ± 0.3 μm In vitro SMCs 20 10 15 18 Providing a mature smooth muscle layer that [216]
SMC expresses robust cell-to-cell junction and
contractile proteins in the presence of pre-
fabricated SMC sheets.
Increasing the growth and proliferation of Cells
by pulsatile perfusion bioreactor conditioning of
the cell sheet-vascular scaffold.
Bombyx mori SF 547 ± 132 and 555 ± 155 In vitro, In SMCs 24 1.1 10 N/A The expected compliance is similar or greater [217]
vivo than that of native rat aorta and
Goretex®prosthesis.
Improving the cell growth and proliferation.
The scaffolds cause a mild host reaction in rat
dorsal tissue.
PET/PU 2.1–2.8 μm In vitro Fibroblast 14–16 10 18, 25 N/A The incorporation of growth of vascular [218]
endothelial and SMCs into the scaffolds for
vascular regeneration should be studied.
TPU/GO 295–397 In vitro 3T3 fibroblast, 18–20 0.5 N/A 18 Cell viability for both types of cells is the utmost [219]
HUVECs at 0.5 wt% GO concentration.
Progress in Materials Science xxx (xxxx) xxxx

(continued on next page)


Table 5 (continued)

Materials Fibers diameter (nm) Studies Cells Voltage (kV) Flow rate Needle–mandrel Needle Major outcomes Ref
composition (ml.h−1) distance (cm) size
(gauge)
M. Rahmati, et al.

Improved cell viability and attachment by


oxygen plasma treatment.
The scaffolds containing 0.5%GO have small
platelet adhesion and activation.
Cycloid PLA 800 μm-1 nm In vitro N/A 8, 20 1 3, 15 N/A The mechanical strength and Young’s module of [220]
the cycloid fibers are upper than those of
conventional tubular random ones.
PCL/PLLA/Collagen PCL = 127 ± 25, In vitro HUVECs 6,10 1, 5 12 27 The 3D scaffolds are more blood compatible [221]
PLLA = 1.2 ± 0.2 μm grafts than ePTFE grafts.
The 3D scaffolds can instantaneously hinder the
first thrombus formation and hasten
endothelialization.
Increasing vascular regeneration by the
nanofibrous structure of scaffolds.
PCL/Collagen N/A In vitro L929, ECs, 20 0.06,0.18 µl/ 15 N/A Substantial enhancement of tensile strength, [222]
SMCs min stitching strength, bursting pressure and
decomposition temperature by crosslinking the
scaffolds with genipin.
The scaffold has good biocompatibility and cell
affinity properties.
GT < 10 μm In vitro hSMCs 22.5 2.5 25 18 Improved cell growth and proliferation. [223]

21
PEU 422 ± 33 In vitro, In A-10 SMCs 10 1 15 23 Improved tissue remodeling in all type A grafts, [224]
vivo and occlusion over the time interval in the type
B grafts.
PCL-a cellular aortic 100–150 μm In vitro, In Endothelial 10 0.8 15 21 Improving the biomechanics of decellularized [225]
vascular vivo vessels.
The scaffolds could prohibit the occurrence of
vasodilation and aneurysm creation after
transplantation.
PLA/PCL combining 0.60 ± 0.08 μm In vitro, In HDFs, HUVEC 18 18 1 21 Improving the collagen remodeling and [226]
with HDFs, vivo biomechanical properties up to day 14.
HUVEC Increasing the cell growth and proliferation.
More infiltration of host cells and collagen
remodeling than those of the HDF-seeded grafts.
Thin constant layers of EC and SMCs could be
formed after 4 weeks.
PCL/ CORMs 1.84 ± 0.47 μm In vitro Rat SMCs 15 2 10 22 The scaffolds can be photoactivated and release [227]
CO.
Improved cell growth and proliferation.
PVA/GT/PCL 432.2 ± 37.9 and In vitro, In HUVECs 12 0.2 12,15 22,27 Quickly in vivo degradation of PVA fibers. [228]
400.1 ± 42.5 vivo Producing electrospun scaffolds with high
porosity.
Improved cell growth and proliferation.
Progress in Materials Science xxx (xxxx) xxxx
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

Fig. 10. A schematic representation of various characteristics of cardiac fibrous scaffolds.

Fig. 11. (a) Hematoxylin and eosin staining. Cells have adhered by the entire mesh. Scale bar = 50 μm. (b) Immunohistochemical staining of F-
actin. Actin fibers traverse the entire thickness of the Cardiac nanofibrous meshes (CNM). Scale bar = 50 μm. (c) Immunohistochemical staining of
cardiac troponin I. Troponin-positive cells are found in the interior of the CNM. Scale bar = 50 μm (d) scanning electron microscopic (SEM)
micrograph of the CNM surface. The mesh is covered with multiple layers of cells. Scale bar = 100 μm. (f) SEM micrograph of a CNM cross-section.
Cells have attached to all fibers in the mesh. Scale bar = 50 μm. Reprinted from [234], with permission from Elsevier.

Further, Schwann cells extend proportionally with neurite extension on small and large fibers; however, this phenomenon delayed on
intermediate fibers [259]. Table 7 presents a summary of major nanofibrous scaffolds used in nerve tissue engineering applications.

5.1.6. Electrospun nanofibrous scaffolds for skin tissue engineering


Skin defects are among the main causes of morbidity and mortality around the world, which account for high socioeconomic costs
[274]. Depending on the duration of healing, these injuries can be acute or chronic [275]. Skin injuries are mainly caused by burns,
diabetes, trauma, surgical procedures, bedsores, aging, and congenital giant nevi [274]. Over the past few decades, researchers have
paid substantial attention to tissue engineering and biomaterials strategies for skin regeneration [109,276–278]. Autograft and
allograft are the two main approaches for skin regeneration [279,280]. Regardless of advances in auto/allografting, these approaches
have some drawbacks including health risks related to surgeries, imperfect donor sites, slow healing rate and scar formation
[281–283]. Biomaterials and interface tissue engineering fields have emerged with the aim of developing promising approaches
through combining biomaterials, engineering strategies and therapeutic agents to overcome these drawbacks [284,285]. The de­
signed skin scaffolds can improve skin healing by protecting the tissue from dehydration and infection, as well as delivering growth
factors and matrix components to the wound sites [286–288]. In addition, by supporting ECM regeneration, scaffolds enable the cell
attachment, proliferation and migration, which eventually lead to the development of new skin [289]. Researchers should consider
some major points in designing skin scaffolds such as angiogenesis, gas exchange, moisture maintenance, and mass transport of tissue.
The biodegradability rate and mechanical properties of skin scaffolds should also match those of the native skin tissue [274]. Several
studies reported that because of the high surface area-to-volume ratio of an electrospun nanofiber, it can be a potential candidate for

22
Table 6
An overview of key electrospun nanofibers used for cardiac tissue engineering applications.
Materials Fibers diameter (nm) Studies cells Voltage (kV) Flow rate Needle–mandrel Needle Major outcomes Ref
composition (ml.h−1) distance (cm) size
M. Rahmati, et al.

(gauge)

PCL 3.4–12.1 μm In vitro Myofibroblast 15–20 8–60 µl/min 15 N/A Improving the cell penetration by [236]
increasing fiber diameter.
Free delivery of cells, with the largest fiber
diameter (12.1 mm).
Homogeneous neo-tissue formation and
adequate matrix deposition.
PCL/GT Random fibers = 239 ± 37, In vitro Rabbit cardiomyocytes 12 1 10 27 Providing anisotropic wetting and [237]
aligned fibers = 269 ± 33 mechanical properties by the scaffolds.
Higher cell attachment and alignment on
aligned PG nanofibrous scaffolds compared
to PCL scaffolds.
PLCL/POC 412–490 In vitro cardiac myoblast 12 1 12 27 Regulating the mechanical and degradation [238]
properties of the scaffolds by changing the
concentration of POC blended with PLCL.
Increasing the cell growth and proliferation
on scaffolds after two and eight days by
increasing the concentration of POC.
PLCL/SF/AV 188 ± 16 In vitro Cardiomyocytes 18 1 12–13 27 The scaffolds are suitable for MI repair. [239]
Improved cardiac cell proliferation on
PLCL/SF/AV nanofibers.

23
The scaffolds have better cardiac expression
proteins myosin and connexin after nine
days cell culture.
Type I Collagen 200–800 In vitro Rat skeletal L6 17 0.6 23 21 The cells seeded onto the scaffolds can [240]
myoblasts preserve their contractile role for 17 days.
Using a new benign solvent, collagen I can
be electrospun without joining other
polymers.
PLGA/PCU PLGA = 280.4 ± 96.5 In vitro H9C2 cardiac myoblasts 20 0.5, 0.8 15, 18 N/A Improved cell growth and proliferation in [241]
PCU = 699.4 ± 201.1 anisotropic textile-templated scaffolds.
All cell-seeded PCU scaffolds had
mechanical properties similar to those of a
human heart.
PGS/zein 300–350 In vitro N/A 15, 20 0.5–2 15 N/A A physicochemically steady structure over a [230]
28-day time period by just a small drop in
pH after 28 days.
(continued on next page)
Progress in Materials Science xxx (xxxx) xxxx
Table 6 (continued)

Materials Fibers diameter (nm) Studies cells Voltage (kV) Flow rate Needle–mandrel Needle Major outcomes Ref
composition (ml.h−1) distance (cm) size
(gauge)
M. Rahmati, et al.

GT 200–600 In vitro Myoblast 17 0.6 23 26 The scaffolds have 19.6 ± 3.6 kPa [242]
modulus, which is similar to natural human
myocardium tissue.
Good cell adhesion and proliferation.
Improved expression of contractile protein
desmin.
PCL 3.3 ± 0.8 μm In vitro, In MSCs N/A N/A N/A N/A The scaffolds provide an appropriate matrix [243]
vivo for MSC cardiac implantation.
The scaffolds could improve cardiac
function and weaken dilatation.
PGS/PCL 1.2 ± 0.2 In vitro C2C12 myoblasts, 15 1.6 15 20 Cell attachment over the first eight hours [244]
neonatal rat after seeding and aligning after 24 h.
cardiomyocytes Scaffolds could direct cell responses.

24
Progress in Materials Science xxx (xxxx) xxxx
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

Fig. 12. A schematic representation of a bio-mimic scaffold for peripheral nerve regeneration application.

skin regeneration applications [109,276–278]. For example, Milan et al. [290] reported that 3D dermal substitutes fabricated using
collagen nanofibers provide a physiological environment for skin cell attachment and expansion [290]. Nanofibers can protect
damaged skin tissue from fluid and protein loss; facilitate elimination of exudates; and prevent invasion by exogenous micro­
organisms [291,292]. Several natural and synthetic polymeric biomaterials can be successfully electrospun for skin regeneration
application. Hadisi et al. [293] demonstrated that a gelatin-oxidized starch nanofiber scaffold loaded with Lawsonia inermis (known as
henna) stimulates healing of second-degree burn wounds while reducing the inflammatory response and macrophage infiltration
[293]. Table 8 summarizes the major nanofibrous scaffolds used in skin tissue engineering applications.

5.2. Electrospun nanofibrous scaffolds for compound delivery

Many studies reported that scaffolds made of polymeric-based electrospun nanofibers can successfully deliver therapeutic agents
for biomedical applications, which is because of their unique functionality, high surface-to-volume ratio, nanoscale morphology, and
high pore connectivity [62,306,307]. The main drawbacks of classical delivery systems are toxicity at the release surface; reduced
drug potency; and costly long-term drug dosage [308]. Controlled and targeted delivery of agents using nanoparticle-based systems
offer more precise solutions to overcome such difficulties by fine-tuning of the drug eluting profile [309–312]. The drug release from
nanoporous 3D scaffolds is dependent on the polymer degradation rate and complex circulation pathway [313]. Various formulations
are suggested to improve the drug release profiles such as combining different polymers and designing polymers with various surface
coatings [314–317]. Electrospun hollow nanofibrous structures produced by coaxial electrospinning methods could be a promising
approach for encapsulating various drugs [318,319]. This method enables high drug loading capacity and enhanced solubilization of
some insoluble drugs [320]. Furthermore, several types of drugs such as anti-cancer drugs [321], antibiotics [310] and poly­
saccharides [322] can be chemically or physically encapsulated in the mass phase of electrospun nanofibers or on their surface.
Kolambkar et al. [323] reported using electrospun PCL nanofibrous tubes loaded with a growth factor and a peptide-modified alginate
hydrogel injected into the tubes to sustain release recombinant bone morphogenetic protein-2 (rhBMP-2) for bone regeneration. The
authors reported that the designed delivery system allows consistent bone bridging of critical bone defects. However, in the absence
of rhBMP-2, the scaffolds are not robust enough to enhance the new bone formation. Fig. 13 shows that by controlling the growth
factor release, the system could improve lamellar bone formation, the tissue integration at interface, and formation of marrow spaces.
In addition, Diab et al. [324] evaluated the efficacy of a delivery system comprising an electrospun PCL nanofibrous mesh tube
and a silk fibroin hydrogel in a rat model for local delivery of BMP-2. Their results indicated that the system enables effective delivery
of BMP-2, which leads to better bone formation than control group (silk fibers) in rats post-surgery. The targeted release of BMP-2
also enhanced the overall biomechanical properties of the scaffold [324]. Hao et al. [325] could successfully develop electrospun
core-shell nano/micro fibers based on PU and cellulose acetate for intravaginal drug delivery [325]. Mendes et al. [326] designed
hybrid electrospun fibers based on chitosan/phospholipids for transdermal drug delivery [326]. The results indicated that fiber
morphology depends on phospholipid content. Additionally, the fibers are stable in phosphate-buffered saline (PBS) because of
chitosan amine groups and phospholipid side chains [326]. Tamayol et al. [327] fabricated elastic PGS/PCL-based nanofibers loaded
with an antibiotic for wound healing applications [327]. They reported that because of the porous structure of the electrospun fibers,
an initial burst release can be detected, which is not suitable for sustained and controlled release (Fig. 14). To address this problem,
various materials and/or methods are used, such as multilayer electrospun fibers, core-shell structures, nanocarriers, and coating
layers [312,319,328]. For example, researchers compared the unitial burst release of a polyester-based scaffold containing dex­
amethasone with and without coating a layer of poly(3,4-ethylenedioxythiophene) (PEDOT). The drug has an initial burst release of
75% in the uncoated group. However, in the coated group, as an electro-responsive system, the initial burst release could be highly
controlled in tissue engineering applications [329].

25
Table 7
An overview of key electrospun nanofibers for nerve tissue engineering applications.
Materials Fibers diameter (nm) Studies Cells Voltage (kV) Flow rate Needle–mandrel Needle size Major outcomes Ref
composition (ml.h−1) distance (cm) (gauge)
M. Rahmati, et al.

PLGA N/A In vitro, In Schwann 12 1 10 N/A The scaffolds do not induce host responses. [260]
vivo Positive nerve regeneration for 5 out of 11 scaffolds, one
month after implantation.
The scaffolds are flexible, penetrable and exhibit no
swelling and tube breakage.
PCL/GT 113–189 In vitro C17.2 12 1 N/A N/A The PCL/gelatin scaffolds have the highest stability at [261]
70:30 proportion.
Improved cell differentiation and proliferation.
The cell direction is parallel to the direction of fibers.
PCL Nanofibers = 251 ± 32, In vitro, In N/A 8 1.5 N/A N/A The scaffolds could improve peripheral nerve [262]
Microfibers = 981 ± 83 vivo regeneration and functional recovery across a 15-mm
serious defect gap in rat models.
P(LLA-CL)/collagen 253 ± 102 In vitro C17.2 12 1 11 27 Fiber orientation, composition, and mechanical [263]
I/collagen III properties have critical roles in improving peripheral
nerve regeneration.
Improving cell proliferation on aligned nanofibers.
CTS 400 In vitro, In Schwann 25 2–8 10 N/A Increasing the tensile strength along the sheet’s axis. [264]
vivo Oriented fibrous sheets could display a Schwann cell
column.
Functional and electrophysiological recovery in the
oriented and the bilayered groups.

26
Sprouting of myelinated axons by axonal maturation in
the isograft, oriented and bilayered groups.
PLLA/PANi 195 ± 30 In vitro Nerve stem 15 1 15 27 The scaffolds presented a conductance of 3 × 10–9 S. [265]
cells Protracted neurite outgrowth compared to the cells
grown on non-stimulated scaffolds.
SF 290 In vitro, In RT4-D6P2T 10 N/A 10 22 Meaningfully increasing ASA in the SFNC group after 1, [266]
vivo Schwann 7 and 10 weeks implantation compared to the control
group (p < 0.05).
Increasing the average ASA in SFNC group after one
week implantation.
Reducing the onset and strictness of autotomy in SFNC
group.
PLCL/collagen 230 ± 31 In vitro MSCs 15 1 12 27 Increasing cell differentiation and proliferation using [267]
neuronal inducing factors such as b-mercaptoethanol,
epidermal growth factor, nerve growth factor and brain
derived growth factor in DMEM/F12 media.
PCL/SIS 0.32–0.47 μm In vitro PC-12 18 2 N/A N/A Increasing the alignment’s grade by increasing electrical [268]
conductivity of the original polymer solution.
Increasing mechanical properties, hydrophilicity and
cell proliferation on mats.
PLCL/Laminin PLCL = 350 ± 112, In vitro Rat Schwann 12, 15 1 11 27 Improving cell proliferation on laminin containing [253]
Laminin = 316 ± 110 core–shell PLCL scaffolds.
PAA 820 In vitro N/A 15 0.8 20 18 The length transition depends on the bathing NaCl [269]
amount and the operation temperature.
PHBV/collagen 386–472 In vitro PC12 15 1 12 27 Improved cell proliferation on aligned PHBV/Collagen [270]
50:50 nanofibers compared to aligned PHBV and PHBV/
Progress in Materials Science xxx (xxxx) xxxx

(continued on next page)


Table 7 (continued)

Materials Fibers diameter (nm) Studies Cells Voltage (kV) Flow rate Needle–mandrel Needle size Major outcomes Ref
composition (ml.h−1) distance (cm) (gauge)
M. Rahmati, et al.

Collagen 75:25 nanofibers.


Aligned nanofibers provide contact guidance to direct
the cells along the direction of fibers.
PCLEEP/GDNF GDNF = 3.96 ± 0.14, In vitro, In GDNF 7.5–8 6, 8 5–6 N/A Randomly dissolving protein throughout the polymer [271]
PCLEEP = 5.08 ± 0.05 μm vivo matrix in an aggregate form, and sustained releasing for
up to two months.
Electrophysiological recovery was detected in 20%,
33%, and 44% of the animals in the EF-C, EF-L, and
EF–L-GDNF groups, respectively.
PCL/GT 0.548 ± 0.140 μm In vitro PC-12 1 0.5 N/A N/A Stimulating the essential biological pathways related to [272]
nerve regeneration including cell adhesion,
proliferation, the neurite outgrowth and differentiation.
PGS/PMMA/GT 225–631 In vitro Rat PC-12 10 1 15 N/A Increasing hydrophilicity and biocompatibility by [273]
adding gelatin.
Improved cell proliferation on the surface.
Elongated cell morphology and inducing nerve stem cell
differentiation.

27
Progress in Materials Science xxx (xxxx) xxxx
Table 8
An overview of key electrospun nanofibers for skin tissue engineering applications.
Materials composition Fibers diameter (nm) Studies Cells Voltage (kV) Flow rate Needle–mandrel Needle size Major outcomes Ref
(ml.h−1) distance (cm) (gauge)
M. Rahmati, et al.

PLGA & PEO containing 280 for PLGA & 760 for PEO In vitro HSF 20 & 25 0.3 & 0.5 15 & 18 N/A Nanofibers have uniform morphology. [294]
rhbFGF& Similar mechanical properties to the human
rhEGF skin.
Improved cell proliferation.
Singular delivery of rhEGF plays a major
role in increasing collagen and elastin
expressions.
PLGA/gelatin gelatin = 175, In vitro HSF 12, 16, 20 0.4, 0.8, 1.2 10 21 Similar mechanical properties to the human [295]
PLGA = 630, skin.
PLGA/EGF = 390 A sustained release of EGF.
Improved blood clotting and platelet
adhesion.
Cell infiltration into the hybrid scaffold and
collagen synthesis.
Gelatin/ BDDGE 276–339 depending on the In vitro hDNF 11 0.4 12 15 Controlling both fiber diameter and [296]
content of BDDGE mechanical properties by BDDGE
concentration.
4% BDDGE has ideal mechanical properties.
No toxicity.
Coll/PCL/ gentamicinn 130 In vitro HDF 12 0,6 12 22 Good controlled release of drug. [297]
No toxicity.

28
Coll/bioactive glass 494 ± 153 In vitro, In HaCaTs, HDF, 16–18 1 10–15 N/A Improving the tensile strength of [298]
vivo endothelial cells. nanofibers.
Nanofibers have antibacterial properties.
Improved HaCaTs proliferation and
migration, secretion of COL-I and VEGF in
HDFs, as well as HUVECs proliferation.
The nanofibers induce skin regeneration
GelMA 1.36–2.18 In vitro, In HDF&HUVEC 20 4 15 N/A Improved cell adhesion, proliferation, [299]
vivo migration in vitro, and formation of 3D
vascular networks in vivo.
Controllable mechanical and degradation
properties.
SF/HAM 250 In vitro MEFs 15–20 0.2–1 ml/min 80 or 180 mm 12–15 Improved mechanical properties of bilayer [300]
scaffolds compared to amniotic membranes.
Additional oxygen plasma treatment
increased the surface wettability.
No toxicity.
PCL N/A In vitro, In N/A 18 1 15 20 Supporting fibroblast functions in vivo. [301]
vivo Increasing the speed of wound healing.
Formation of hydroxyl and carboxyl groups
on the membrane.
PCL/Coll 820 In vitro hEnSCs 24 0.5 20 21 The scaffold is a cost-effective, safe and [302]
angiogenesis stimulator.
High mechanical properties.
No toxicity.
689 ± 48 In vitro L929 18 1 N/A 18 [303]
Progress in Materials Science xxx (xxxx) xxxx

(continued on next page)


Table 8 (continued)

Materials composition Fibers diameter (nm) Studies Cells Voltage (kV) Flow rate Needle–mandrel Needle size Major outcomes Ref
(ml.h−1) distance (cm) (gauge)
M. Rahmati, et al.

PLLA/COL/ aloevera/ The composite scaffold is an appropriate


coated with chitosan structure for sustained release of AV gel.
PCL/HA/ epidermal GFs 272 ± 38 In vitro, In HaCaT, fibroblasts 18 1 15 N/A Increasing the release of EGF by increasing [304]
vivo nanofiber Hydrophilicity through HA.
EGF and HA increase cell infiltration,
collagen and TGF-b1 gene expression as
well as the ration of collagen III to I.
Nanofibers induce epidermis regeneration
in the early phases of wound healing.
β-glucan butyrate/β- N/A In vitro, In 3T3, HaCaTs 30 2 μl/ min 15 18 The scaffold has an advanced structure, a [305]
glucan acetate vivo high degree of aqueous stability and water
holding capacity.
The scaffold has good mechanical
properties and cytocompatibility.
The scaffold accelerates wound healing in a
full thickness mouse model.

29
Progress in Materials Science xxx (xxxx) xxxx
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

Fig. 13. Typical radiographs of reconstituted bones 4 and 12 weeks after implantation. Defects in Groups I (mesh alone) and II (mesh with alginate)
exhibited small amount of bone formation, and did not exhibit bridging, even after 12 weeks implantation. After 4 weeks implantation, defects in
Group III (mesh with alginate containing rhBMP-2) samples were filled with bone tissue, while Group IV (perforated mesh with alginate containing
rhBMP-2) samples showed the most robust mineralization. All samples in Groups III and IV were bridged with densely packed bone after 12 weeks
implantation. Reprinted from [323], with permission from Elsevier.

Fig. 14. A Schematic of controlling drug release using multilayer fibers and core-shell structures.

6. Surface modification of nanofibrous scaffolds

Several surface modification strategies are developed to improve the surface characteristics of electrospun nanofibrous scaffolds
[330–332]. Synthetic polymers require some modification to enhance their hydrophilicity and cellular attachment. There are specific

30
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

surface modification approaches available to change the chemical nature of the surface, improve surface hydrophilicity, and provide
a desirable environment for cellular adhesion as well as proliferation [333–335]. Plasma treatment, wet-chemical methods, surface
graft polymerization, and co-electrospinning of surface functionalization are the most common used approaches, which are discussed
in the following sections.

6.1. Plasma treatment

Plasma treatment is commonly used to immobilize polar groups, such as carboxyl or amine groups, to change the chemical nature
of surface and improve surface wettability [336]. Based on the plasma source different functional groups can be introduced to the
surface, which subsequently enables covalent immobilization of bioactive molecules [337–339]. Plasma treatment with ammonia,
oxygen or air can generate carboxyl groups or amine groups on the surface [340]. In addition, multiple ECM protein constituents,
such as collagen, gelatin, and fibronectin, can be immobilized by plasma treatment, which results in improved cellular adhesion and
proliferation [341–343]. For instance, researchers functionalized the surface of PLGA-based electrospun nanofibrous scaffolds by
plasma, oxygen or ammonia to enhance surface hydrophilicity and evaluated their biological performance [344]. The functionalized
surfaces could improve fibroblast adhesion and proliferation compared to non-functionalized surfaces [344]. Further, Chen et al.
[345] modified electrospun PLLA nanofibers with cationized gelatin to enhance their compatibility with chondrocytes, and under­
lined the potential of such modified nanofibers as candidates for cartilage regeneration. They treated scaffolds with oxygen plasma to
introduce surface carboxylic groups. Then, using a coupling agent (water-soluble carbodiimide), they covalently grafted cationized
gelatin onto the fiber surface. The scaffolds improve the viability, proliferation and differentiation of rabbit articular chondrocyte
compared to untreated PLLA nano-fibrous membranes. In vivo studies indicated the formation of ectopic cartilage tissue 28 days after
implantation of the scaffolds (Fig. 15) [345].

6.2. Wet-chemical method

Surface-modified nanofibrous scaffolds are treated under acidic or basic conditions to generate reactive functional groups and to
improve surface wettability [346]. This method is based on random chemical breaking of ester linkages in the polymer backbone on
the surface, resulting in generating surface hydroxyl and carboxylic groups from degraded water-insoluble polymer fragments [347].
One research group demonstrated the benefits of wet-chemical surface modification for bone tissue engineering applications by
fabricating a novel electrospun PCL nanofibrous scaffold that was hydrolyzed with NaOH [344]. The researchers observed that the
modified fibers provide a better microenvironment for cell growth and proliferation than unmodified surfaces [344]. A variety of
molecules are used as surface-bioactive agents such as ECM proteins and ECM-derived peptides. One of the most common molecules
is collagen, which can be immobilized onto a nanofibrous scaffold using this surface modification method [348]. The surface
modification by immobilized collagen significantly improves the neural stem cell attachment and growth in neural tissue engineering
[349].

6.3. Surface graft polymerization

Grafting hydrophilic components on the surface of hydrophobic polymers results in improving the molecular and cellular re­
sponses to the nanofiber surface [350,351]. Surface graft polymerization is used not only to increase surface hydrophilicity but also to
introduce multifunctional groups for covalent immobilization of bioactive molecules on the surface [352]. Surface graft

Fig. 15. Histological analysis of (a–d) control modified scaffolds and (e–h) chondrocyte-seeded modified scaffolds 4 weeks after implantation.
Sections were stained with hematoxylin and eosin (a, e), Alcian blue (b, f), or safranin O (c, g); and immunohistochemically, for collagen type II (d,
f). N, nanofiber; IC, inflammatory cells; ST, surrounding tissue. Scale bar = 100 μm. Reprinted from [345], with the permission from Elsevier.

31
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

polymerization begins with plasma and ultraviolet (UV)-irradiation treatment to produce free radicals [353]. In a typical reaction,
when acrylic acid is successfully grafted onto a PLCL scaffold by UV-irradiation, the modified layer creates a surface that is com­
patible with hMSCs [354].

6.4. Co-electrospinning of surface-active agents and polymers

During the initial surface modification, it is possible to directly graft nanoparticles and/or functional polymers onto the nanofiber
surface. For instance, PLLA solution could be combined with HA nanoparticles by co-electrospinning, leading to the fabrication of
nanofibrous scaffolds containing HA nanoparticles on the surface [355]. Compared with pure PLLA nanofibrous scaffolds, these
composite nanofibrous scaffolds exhibit an extended degradation rate because of the ionic interactions between PLLA ester groups
and calcium ions of HA [356]. The aim of this method is to combine a range of electrospinnable functional polymers and polymers
conjugated with nanoparticles to yield bio-functionalized nanofibrous surfaces [357].

7. Innovative approaches to electrospinning

Electrospinning can be used to create nanofibrous scaffolds of various shapes, such as core-shell, hollow, porous, and honeycomb
[20,145,358–360]. The generated scaffolds are promising because of their structural similarities to the natural ECM. Nguyen et al.
[361] designed some 3D nanofiber-hydrogel systems for therapeutic agent delivery to direct axon regeneration for spinal cord injury
treatment [361]. The authors reported that the resulting nanofibers mimic the topographical properties of natural ECM more closely
than the hydrogels and micron-sized structures [361]. Because of the high surface area and porous structure, cells have a high
tendency to proliferate and differentiate on a nanofiber surface. In addition, the high surface area of such a nanofiber enables higher
surface adsorption of proteins and also promotes higher binding site formation. Using electrospinning, researchers are attempting to
develop nanofibrous scaffolds with different shapes and properties such as core-shell, highly porous, and multilayer nanofibers
[20,145,358–360,362]. Core-shell structures are used to make nanofibrous scaffolds [222,310]. This strategy allows the inclusion of
bioactive agents, such as bacteria, enzymes, drugs, viruses, and proteins, within the core, and controls their release kinetics
[363–365]. Coaxial core-shell electrospinning set-up is among the most common set-ups used for making such structures
[196,222,366]. This technique involves two separate solutions (generally, a hydrophobic polymer and a solution of water-soluble
bioactive factors) that can be co-electrospun without the need for direct mixing using two separate syringes [367].
Hollow nanofibrous scaffolds that are fabricated by direct coaxial electrospinning and chemical vapor deposition (CVD) could be
appropriate for special applications such as hydrogen storage and nanofluidic systems [359,368,369]. During coaxial electrospinning,
nanofibers are created by the same technique with which core-shell structures are produced. However, the core material is dissolved
in an appropriate solvent at the end of electrospinning process [370]. In addition, coaxial electrospinning is useful for creating highly
porous hollow nanofiber scaffolds for applications in the fields of optoelectronics, catalysis, nanofluidics, biosensor systems, drug
delivery, and tissue engineering [371,372].
Highly porous nanofibers are more useful for a variety of applications than the hollow nanofibers [373]. They can be fabricated
with specific topology by choosing either special solvents or polymer composites. These structures can also be electrospun using a
common solvent for immiscible polymers that is later dissolved yielding perfectly porous materials [374]. This method is grounded
on phase separation, which is dependent on the evaporation rate of the solution [375]. When generating 3D fibrous structures using
common electrospinning techniques, the longer the spinning time, the thicker the fabricated mat. This can yield 3D nanofibrous
structures with variable thicknesses [376].
The 3D fibrous structures can be designed by combining a nanofiber layer with another fibrous micro- or nanolayer (known as
multilayering-electrospinning) [31,328,377,378]. Some critical scaffold parameters can be controlled during multilayered electro­
spinning process including blending, fiber diameter, layer porosity, and the number of electrospun fiber layers, which are all essential
for directing cellular responses [379].
Because of some drawbacks of traditional methods (low drug-loading efficiency, poor scalability, and inefficient incorporation of
hydrophilic drugs, etc.), many studies suggest combining electrospinning with electrospraying [380,381]. Electrospraying is an
advanced version of the electrospinning process, which can potentially be useful for the creation of nanofibrous mats and nano­
particles [382]. Using this method, nanofibers can be synthesized by applying a solution of polymer and drug in an adequately
conductive solvent. Because of the fact that electrospinning and electrospraying methods share similar principles in regulating their
parameters, they could be combined to design highly advanced nanofibers. Electrospinning-electrospraying is a novel method that
can be used to fabricate multilayered scaffolds, by applying layer-by-layer deposition of different mats made of various polymers
[383,384].
In addition, by combining 3D printing with electrospinning, better micro- to nano-structured scaffolds/devices and flexible
electronics/flex circuits can be synthesized for controlled delivery of therapeutic agents as well as tissue engineering applications
[385–387].

8. Concluding remarks

Using advanced biomaterials, together with more promising engineering strategies opens the routes of innovation in the field of
tissue engineering and regenerative medicine. Electrospinning is one of the most adaptable and promising techniques for producing a
range of fibrous mats [82]. It can be used in different approaches to combine the material features with various morphological

32
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

characteristics for advanced tissue engineering uses. Nanofibrous scaffolds produced by electrospinning are biomaterials quite sur­
prisingly similar to the ECM microstructure. Several studies demonstrated that these advanced nanofibrous scaffolds improve cell
attachment, differentiation, and proliferation. Considering these features, electrospun nanofibrous scaffolds are capable of providing
a wide range of functions needed for tissue engineering and biomedical applications in the near future [388]. Some novel techniques
are tested to generate electrospun 3D fibrous structures, including multi-layering electrospinning [389], self-assembly [390], using a
3D collecting template [380], and folding or accumulating 2D fiber films [391,392]. Among these techniques, consecutive elec­
trospinning or multi-layering electrospinning emerged as the simplest and the most useful technique. It allows electrospun fiber
layers to be fabricated with a defined thickness and accurate 3D constructs, on a microscale range. By controlling some key para­
meters of the electrospinning process, such as the polymer solution concentration, viscosity, ambient humidity, and electrostatic
field, a prompt growth of 3D nanofibrous structures could be achieved [393]. Nevertheless, further in vitro and in vivo studies are
necessary to characterize previously fabricated micro- and nano-scale fibers in detail. In addition, as outlined in the technical tables
of this review paper, some nanofibrous scaffolds are studied extensively in vitro, and we believe that it is time to focus on their
interactions with the native tissues in vivo. Instead of generally considering designed nanofibrous scaffolds for biomedical applica­
tions, scientists are about to introduce specific applications for these scaffolds by precisely adjusting system features to match the
targeted cells and/or tissues. To enhance the properties of the produced electrospun nanofibrous scaffolds, a combination of natural
and synthetic polymers with surface modification might constitute another challenge. It is clear that these advanced 3D scaffolds will
be key players in the future of tissue engineering and regenerative medicine.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Acknowledgement

Maryam Rahmati was supported by a project “Promoting patient safety by a novel combination of imaging technologies for
biodegradable magnesium implants, MgSafe” funded by European Training Network within the framework of Horizon 2020 Marie
Skłodowska-Curie Action (MSCA) grant number No 811226 (www.mgsafe.eu). The authors would like to thank Dr. Payam Zarrintaj
from the School of Chemical Engineering, Oklahoma State University, USA, for kind assistance with modification of the manuscript
and discussion on the compatibility mechanisms and the functions of polymers on tissues.

References

[1] Kemppainen JM. Mechanically stable solid freeform fabricated scaffolds with permeability optimized for cartilage tissue engineering. ProQuest 2008.
[2] Rahmati M, Silva EA, Reseland JE, Catherine AH, Haugen HJ. Biological responses to physicochemical properties of biomaterial surface. Chem Soc Rev 2020.
[3] Pina S, Ribeiro VP, Marques CF, Maia FR, Silva TH, Reis RL, et al. Scaffolding strategies for tissue engineering and regenerative medicine applications. Materials
(Basel) 2019;12:1824.
[4] O’brien FJ. Biomaterials & scaffolds for tissue engineering. Mater Today 2011;14:88–95.
[5] Khang G. Handbook of intelligent scaffold for tissue engineering and regenerative medicine. CRC Press; 2012.
[6] Fathi-Achachelouei M, Knopf-Marques H, da Silva CER, Barthes J, Bat E, Tezcaner A, et al. Use of nanoparticles in tissue engineering and regenerative medicine.
Front Bioeng Biotechnol 2019;7:113.
[7] Ramakrishna S, Mayer J, Wintermantel E, Leong KW. Biomedical applications of polymer-composite materials: a review. Compos Sci Technol
2001;61:1189–224.
[8] Geckil H, Xu F, Zhang XH, Moon S, Demirci U. Engineering hydrogels as extracellular matrix mimics. Nanomedicine 2010;5:469–84.
[9] Rahmati M, Pennisi CP, Mobasheri A, Mozafari M. Bioengineered scaffolds for stem cell applications in tissue engineering and regenerative medicine. Adv Exp
Med Biol 2018;1107:73–89.
[10] Zhang LG, Khademhosseini A, Webster T. Tissue and organ regeneration: advances in micro-and nanotechnology. CRC Press; 2016.
[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] Datta S, Menon G. Nanofibers from polyhydroxyalkanoates and their applications in tissue engineering. Biotechnological applications of polyhydroxyalk­
anoates. Springer; 2019. p. 409–20.
[13] Joshi V, Srivastava CM, Gupta AP, Vats M. Electrospun nano-architectures for tissue engineering and regenerative medicine. Nanoscience in Medicine Vol 1:
Springer 2020:213–48.
[14] von der Mark K, Park J, Bauer S, Schmuki P. Nanoscale engineering of biomimetic surfaces: cues from the extracellular matrix. Cell Tissue Res 2010;339:131.
[15] Santo VE, Gomes ME, Mano JF, Reis RL. From nano-to macro-scale: nanotechnology approaches for spatially controlled delivery of bioactive factors for bone
and cartilage engineering. Nanomedicine 2012;7:1045–66.
[16] Wu RX, Ma C, Liang YX, Chen FM, Liu XH. ECM-mimicking nanofibrous matrix coaxes macrophages toward an anti-inflammatory phenotype: Cellular behaviors
and transcriptome analysis. Appl Mater Today 2020;18:100508.
[17] Ayoub AS, Lucia LA. Introduction to renewable biomaterials: first principles and concepts. John Wiley & Sons; 2017.
[18] Sekuła M, Zuba-Surma EK. Biomaterials and stem cells: promising tools in tissue engineering and biomedical applications. Biomaterials in regenerative
medicine. InTech 2018.
[19] Ducheyne P. Comprehensive biomaterials. Elsevier; 2015.
[20] Yao TY, Chen HL, Samal P, Giselbrecht S, Baker MB, Moroni L. Self-assembly of electrospun nanofibers into gradient honeycomb structures. Mater Des
2019;168:107614.
[21] Rasouli R, Barhoum A, Bechelany M, Dufresne A. Nanofibers for biomedical and healthcare applications. Macromol Biosci 2019;19:e1800256.
[22] Kang J, Hwang JY, Huh M, Yun SI. Porous poly(3-hydroxybutyrate) scaffolds prepared by non-solvent-induced phase separation for tissue engineering.
Macromol Res 2020;1.
[23] Jin K, Eyer S, Dean W, Kitto D, Bates FS, Ellison CJ. Bimodal nanofiber and microfiber nonwovens by melt-blowing immiscible ternary polymer blends. Ind Eng
Chem Res 2019;59:5238–46.

33
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

[24] Heidari M, Bahrami SH, Ranjbar-Mohammadi M, Milan PB. Smart electrospun nanofibers containing PCL/gelatin/graphene oxide for application in nerve tissue
engineering. Mater Sci Eng C Mater Biol Appl 2019;103:109768.
[25] Lu T, Li Y, Chen T. Techniques for fabrication and construction of three-dimensional scaffolds for tissue engineering. Int J Nanomed 2013;8:337–50.
[26] Gervaso F, Sannino A, Peretti GM. The biomaterialist's task: scaffold biomaterials and fabrication technologies. Joints 2013;1:130–7.
[27] Garg T, Singh O, Arora S, Murthy R. Scaffold: a novel carrier for cell and drug delivery. Crit Rev Ther Drug Carrier Syst 2012;29:1–63.
[28] Lin W, Chen M, Qu T, Li J, Man Y. Three-dimensional electrospun nanofibrous scaffolds for bone tissue engineering. J Biomed Mater Res B Appl Biomater
2020;108:1311–21.
[29] Gao X, Han S, Zhang R, Liu G, Wu J. Progress in electrospun composite nanofibers: composition, performance and applications for tissue engineering. J Mater
Chem B 2019;7:7075–89.
[30] Movahedi M, Asefnejad A, Rafienia M, Khorasani MT. Potential of novel electrospun core-shell structured polyurethane/starch (hyaluronic acid) nanofibers for
skin tissue engineering: In vitro and in vivo evaluation. Int J Biol Macromol 2020;146:627–37.
[31] Rad ZP, Mokhtari J, Abbasi M. Calendula officinalis extract/PCL/Zein/Gum arabic nanofibrous bio-composite scaffolds via suspension, two-nozzle and mul­
tilayer electrospinning for skin tissue engineering. Int J Biol Macromol 2019;135:530–43.
[32] Guo TW, Yang X, Deng J, Zhu LC, Wang BC, Hao SL. Keratin nanoparticles-coating electrospun PVA nanofibers for potential neural tissue applications. J Mater
Sci-Mater Med 2019;30:9.
[33] Tondnevis F, Keshvari H, Mohandesi JA. Fabrication, characterization, and in vitro evaluation of electrospun polyurethane-gelatin-carbon nanotube scaffolds
for cardiovascular tissue engineering applications. J Biomed Mater Res B Appl Biomater 2020;108:2276–93.
[34] Bochicchio B, Barbaro K, De Bonis A, Rau JV, Pepe A. Electrospun poly(d, l-lactide)/gelatin/glass-ceramics tricomponent nanofibrous scaffold for bone tissue
engineering. J Biomed Mater Res A 2020;108:1064–76.
[35] Wang X, Ding B, Li B. Biomimetic electrospun nanofibrous structures for tissue engineering. Mater Today (Kidlington) 2013;16:229–41.
[36] Zhong S, Zhang Y, Lim CT. Fabrication of large pores in electrospun nanofibrous scaffolds for cellular infiltration: a review. Tissue Eng Part B Rev
2012;18:77–87.
[37] Keirouz A, Chung M, Kwon J, Fortunato G, Radacsi N. 2D and 3D electrospinning technologies for the fabrication of nanofibrous scaffolds for skin tissue
engineering: A review. Wiley Interdiscip Rev Nanomed Nanobiotechnol 2020;12:e1626.
[38] Wani S, Sofi HS, Majeed S, Sheikh FA. Recent trends in chitosan nanofibers: from tissue-engineering to environmental importance: a review. Mater Sci Res India
2017;14:89–99.
[39] Bergmann CP, Stumpf A. Dental ceramics. Biomaterials 2013.
[40] Williams D. The Williams dictionary of biomaterials Liverpool, UK: Liverpool. University Press; 1999. 42 p.
[41] Morais JM, Papadimitrakopoulos F, Burgess DJ. Biomaterials/tissue interactions: possible solutions to overcome foreign body response. AAPS J
2010;12:188–96.
[42] Trindade R, Albrektsson T, Tengvall P, Wennerberg A. Foreign body reaction to biomaterials: on mechanisms for buildup and breakdown of osseointegration.
Clin Implant Dentistry Related Res 2016;18:192–203.
[43] Gethin G. Understanding the inflammatory process in wound healing. Brit J Commun Nurs 2012.
[44] Butterfield TA, Best TM, Merrick MA. The dual roles of neutrophils and macrophages in inflammation: a critical balance between tissue damage and repair. J
Athletic Train 2006;41:457.
[45] Rahmati M, Mozafari MJABS, Engineering. Selective Contribution of Bioactive Glasses to Molecular and Cellular Pathways; 2019.
[46] Rahmati M, Mozafari M. Protein adsorption on polymers. Mater Today Commun 2018;17:527–40.
[47] Anderson JM, Rodriguez A, Chang DT. Foreign body reaction to biomaterials. Seminars Immunol: Elsevier 2008:86–100.
[48] Anderson JM. Inflammatory response to implants. ASAIO Trans 1988;34:101–7.
[49] Xia Z, Triffitt JT. A review on macrophage responses to biomaterials. Biomed Mater 2006;1:R1.
[50] Sheikh Z, Brooks PJ, Barzilay O, Fine N, Glogauer M. Macrophages, foreign body giant cells and their response to implantable biomaterials. Materials
2015;8:5671–701.
[51] Williams DF. There is no such thing as a biocompatible material. Biomaterials 2014;35:10009–14.
[52] Williams DF. Biocompatibility in clinical practice: predictable and unpredictable outcomes. Prog Biomed Eng 2019;1:013001.
[53] Williams DF. Biocompatibility pathways: biomaterials-induced sterile inflammation, mechanotransduction, and principles of biocompatibility control. ACS
Biomater. Sci. Eng 2016;3:2–35.
[54] Williams DF. On the mechanisms of biocompatibility. Biomaterials 2008;29:2941–53.
[55] Anderson JM. Future challenges in the in vitro and in vivo evaluation of biomaterial biocompatibility. Regen Biomater 2016;3:73–7.
[56] Chan BP, Leong KW. Scaffolding in tissue engineering: general approaches and tissue-specific considerations. Eur Spine J 2008;17(Suppl 4):467–79.
[57] Jahanmard F, Eslaminejad MB, Amani-Tehran M, Zarei F, Rezaei N, Croes M, et al. Incorporation of F-MWCNTs into electrospun nanofibers regulates osteo­
genesis through stiffness and nanotopography. Mater Sci Eng C-Mater Biol Appl 2020;106:110163.
[58] Pauly HM, Kelly DJ, Popat KC, Trujillo NA, Dunne NJ, McCarthy HO, et al. Mechanical properties and cellular response of novel electrospun nanofibers for
ligament tissue engineering: Effects of orientation and geometry. J Mech Behav Biomed Mater 2016;61:258–70.
[59] Hollinger JO. An introduction to biomaterials. CRC Press; 2011.
[60] Kalantari K, Afifi AM, Jahangirian H, Webster TJ. Biomedical applications of chitosan electrospun nanofibers as a green polymer - Review. Carbohydr Polym
2019;207:588–600.
[61] Xue J, Wu T, Dai Y, Xia Y. Electrospinning and electrospun nanofibers: methods, materials, and applications. Chem Rev 2019;119:5298–415.
[62] Bhattarai RS, Bachu RD, Boddu SHS, Bhaduri S. Biomedical applications of electrospun nanofibers: drug and nanoparticle delivery. Pharmaceutics 2019;11:5.
[63] Formhals A. Method and apparatus for spinning (US, 2349950); 1944.
[64] He JH, Wan YQ, Yu JY. Scaling law in electrospinning: relationship between electric current and solution flow rate. Polymer 2005;46:2799–801.
[65] Chronakis IS. Micro-/nano-fibers by electrospinning technology: processing, properties and applications. Micro-Manufacturing Engineering and Technology,
William Andrew-Elsevier 2010:264-86.
[66] Ghorbani F, Nojehdehyan H, Zamanian A, Gholipourmalekabadi M, Mozafari M. Synthesis, physico-chemical characteristics and cellular behavior of poly
(lactic-co-glycolic acid)/gelatin nanofibrous scaffolds for engineering soft connective tissues.
[67] Faridi MR, Naderi N, Koochaki JA. Needleless electrospinning apparatus. Google Patents 2016.
[68] Andrady AL, Ensor DS, Walker TA, Prabhu P. Nanofiber mats and production methods thereof. Google Patents 2009.
[69] Liu W, Thomopoulos S, Xia Y. Electrospun nanofibers for regenerative medicine. Adv Healthc Mater 2012;1:10–25.
[70] Garg K, Bowlin GL. Electrospinning jets and nanofibrous structures. Biomicrofluidics 2011;5:13403.
[71] Ding Y, Xu W, Xu T, Zhu Z, Fong H. Theories and principles behind electrospinning. Adv Nanofibrous Mater Manuf Technol Electrospinning 2019;22.
[72] Liu Y, Wang C. Advanced nanofibrous materials manufacture technology based on electrospinning. CRC Press; 2019.
[73] Huang ZM, Zhang YZ, Kotaki M, Ramakrishna S. A review on polymer nanofibers by electrospinning and their applications in nanocomposites. Compos Sci
Technol 2003;63:2223–53.
[74] Wang HY, Feng YK, Fang ZC, Xiao RF, Yuan WJ, Khan M. Fabrication and characterization of electrospun gelatin-heparin nanofibers as vascular tissue
engineering. Macromol Res 2013;21:860–9.
[75] Sasipriya K, Suriyaprabha R, Prabu P, Rajendran V. Synthesis and characterisation of polymeric nanofibers poly (vinyl alcohol) and poly (vinyl alcohol)/silica
using indigenous electrospinning set up. Mater Res-Ibero-Am J Mater 2013;16:824–30.
[76] Fu W, Liu Z, Feng B, Hu R, He X, Wang H, et al. Electrospun gelatin/PCL and collagen/PLCL scaffolds for vascular tissue engineering. Int J Nanomed
2014;9:2335–44.
[77] Baptista A, Ferreira I, Borges J. Electrospun fibers in composite materials for medical applications. J Compos Biodegrad Polym 2013;1:56–65.

34
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

[78] Cesur S, Oktar FN, Ekren N, Kilic O, Alkaya DB, Seyhan SA, et al. Preparation and characterization of electrospun polylactic acid/sodium alginate/orange oyster
shell composite nanofiber for biomedical application. J Aust Ceram Soc 2019;1–11.
[79] Nagam Hanumantharao S, Rao S. Multi-functional electrospun nanofibers from polymer blends for scaffold tissue engineering. Fibers 2019;7:66.
[80] Doshi J, Srinivasan G, Reneker D. A novel electrospinning process. Polym News 1995;20:206–13.
[81] Li WJ, Tuan RS. Fabrication and application of nanofibrous scaffolds in tissue engineering. Curr Protoc Cell Biol 2009;Chapter 25:Unit 25 2.
[82] Sun B, Long YZ, Zhang HD, Li MM, Duvail JL, Jiang XY, et al. Advances in three-dimensional nanofibrous macrostructures via electrospinning. Prog Polym Sci
2014;39:862–90.
[83] Reneker DH, Chun I. Nanometre diameter fibres of polymer, produced by electrospinning. Nanotechnology 1996;7:216–23.
[84] Reneker DH, Yarin AL. Electrospinning jets and polymer nanofibers. Polymer 2008;49:2387–425.
[85] Li Z, Wang C. Effects of working parameters on electrospinning. One-dimensional nanostructures. Springer; 2013. p. 15–28.
[86] Yang QB, Li ZY, Hong YL, Zhao YY, Qiu SL, Wang C, et al. Influence of solvents on the formation of ultrathin uniform poly(vinyl pyrrolidone) nanofibers with
electrospinning. J Polym Sci Part B-Polym Phys 2004;42:3721–6.
[87] Deitzel JM, Kleinmeyer J, Harris D, Tan NCB. The effect of processing variables on the morphology of electrospun nanofibers and textiles. Polymer
2001;42:261–72.
[88] Sperling LH. Introduction to physical polymer science. John Wiley & Sons; 2005.
[89] Tiwari SK, Venkatraman SS. Importance of viscosity parameters in electrospinning: Of monolithic and core-shell fibers. Mater Sci Eng C-Mater Biol Appl
2012;32:1037–42.
[90] Gupta P, Elkins C, Long TE, Wilkes GL. Electrospinning of linear homopolymers of poly(methyl methacrylate): exploring relationships between fiber formation,
viscosity, molecular weight and concentration in a good solvent. Polymer 2005;46:4799–810.
[91] Yang Y, Jia ZD, Liu JA, Li Q, Hou L, Wang LM, et al. Effect of electric field distribution uniformity on electrospinning. J Appl Phys 2008;103:104307.
[92] Koski A, Yim K, Shivkumar S. Effect of molecular weight on fibrous PVA produced by electrospinning. Mater Lett 2004;58:493–7.
[93] Wen XT, Fan HS, Tan YF, Cao H, Li H, Cai B, et al. Preparation of electrospun PLA nanofiber scaffold and the evaluation in vitro. Key Engineering Materials:
Trans Tech Publ 2005:139–42.
[94] Agarwal S, Wendorff JH, Greiner A. Use of electrospinning technique for biomedical applications. Polymer 2008;49:5603–21.
[95] Rezaei A, Nasirpour A, Fathi M. Application of cellulosic nanofibers in food science using electrospinning and its potential risk. Compr Rev Food Sci Food Saf
2015;14:269–84.
[96] Haghi AK, Akbari M. Trends in electrospinning of natural nanofibers. Phys Status Solidi A-Appl Mater Sci 2007;204:1830–4.
[97] Mit-uppatham C, Nithitanakul M, Supaphol P. Ultrafine electrospun polyamide-6 fibers: effect of solution conditions on morphology and average fiber diameter.
Macromol Chem Phys 2004;205:2327–38.
[98] Angammana CJ, Jayaram SH. Analysis of the effects of solution conductivity on electrospinning process and fiber morphology. IEEE Trans Ind Appl
2011;47:1109–17.
[99] Rodoplu D, Mutlu M. Effects of electrospinning setup and process parameters on nanofiber morphology intended for the modification of quartz crystal mi­
crobalance surfaces. J Eng Fibers Fabr 2012;7:118–23.
[100] Pham QP, Sharma U, Mikos AG. Electrospun poly(epsilon-caprolactone) microfiber and multilayer nanofiber/microfiber scaffolds: characterization of scaffolds
and measurement of cellular infiltration. Biomacromolecules 2006;7:2796–805.
[101] Beachley V, Wen X. Effect of electrospinning parameters on the nanofiber diameter and length. Mater Sci Eng C Mater Biol Appl 2009;29:663–8.
[102] Greiner A, Wendorff JH. Electrospinning: a fascinating method for the preparation of ultrathin fibers. Angew Chem Int Ed Engl 2007;46:5670–703.
[103] Fridrikh SV, Yu JH, Brenner MP, Rutledge GC. Controlling the fiber diameter during electrospinning. Phys Rev Lett 2003;90:144502.
[104] 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 2003;44:5721–7.
[105] Yuan XY, Zhang YY, Dong CH, Sheng J. Morphology of ultrafine polysulfone fibers prepared by electrospinning. Polym Int 2004;53:1704–10.
[106] Casper CL, Stephens JS, Tassi NG, Chase DB, Rabolt JF. Controlling surface morphology of electrospun polystyrene fibers: Effect of humidity and molecular
weight in the electrospinning process. Macromolecules 2004;37:573–8.
[107] Khademhosseini A, Vacanti JP, Langer R. Progress in tissue engineering. Sci Am 2009;300:64–71.
[108] Sill TJ, von Recum HA. Electrospinning: applications in drug delivery and tissue engineering. Biomaterials 2008;29:1989–2006.
[109] Koosha M, Raoufi M, Moravvej H. One-pot reactive electrospinning of chitosan/PVA hydrogel nanofibers reinforced by halloysite nanotubes with enhanced
fibroblast cell attachment for skin tissue regeneration. Colloids Surf B Biointerfaces 2019;179:270–9.
[110] Gunatillake P, Mayadunne R, Adhikari R. Recent developments in biodegradable synthetic polymers. Biotechnol Annu Rev 2006;12:301–47.
[111] Nejad MR, Yousefzadeh M, Solouk A. Electrospun PET/PCL small diameter nanofibrous conduit for biomedical application. Mater Sci Eng C-Mater Biol Appl
2020;110:110692.
[112] Huang J, Zhou XB, Shen YB, Li HB, Zhou GD, Zhang WJ, et al. Asiaticoside loading into polylactic-co-glycolic acid electrospun nanofibers attenuates host
inflammatory response and promotes M2 macrophage polarization. J Biomed Mater Res Part A 2020;108:69–80.
[113] Piskin E, Bölgen N, Egri S, Isoglu IA. Electrospun matrices made of poly (α-hydroxy acids) for medical use; 2007.
[114] Wang YQ, Yao DY, Li LH, Qian ZY, He W, Ding R, et al. Effect of electrospun silk fibroin-silk sericin films on macrophage polarization and vascularization. ACS
Biomater Sci Eng 2020;6:3502–12.
[115] Wang Y, Cao ZJ, Cheng RY, Qin ML, Zhang DD, Deng LF, et al. Immunomodulated electrospun fibrous scaffolds via bFGF camouflage for pelvic regeneration.
Appl Mater Today 2019;15:570–81.
[116] Mukherjee S, Darzi S, Paul K, Cousins FL, Werkmeister JA, Gargett CE. Electrospun nanofiber meshes with endometrial MSCs modulate foreign body response
by increased angiogenesis, matrix synthesis, and anti-inflammatory gene expression in mice: implication in pelvic floor. Front Pharmacol 2020;11:353.
[117] Garcia-Salinas S, Evangelopoulos M, Gamez-Herrera E, Arruebo M, Irusta S, Taraballi F, et al. Electrospun anti-inflammatory patch loaded with essential oils for
wound healing. Int J Pharm 2020;577:119067.
[118] Neves NM, Campos R, Pedro A, Cunha J, Macedo F, Reis RL. Patterning of polymer nanofiber meshes by electrospinning for biomedical applications. Int J
Nanomed 2007;2:433–48.
[119] Wobma H, Vunjak-Novakovic G. Tissue engineering and regenerative medicine 2015: a year in review. Tissue Eng 2015.
[120] Van Pham P. Tissue engineering and regenerative medicine. Springer; 2019.
[121] Higuchi A, Ling QD, Hsu ST, Umezawa A. Biomimetic cell culture proteins as extracellular matrices for stem cell differentiation. Chem Rev 2012;112:4507–40.
[122] Qian YF, Zhang Z, Zheng LJ, Song RY, Zhao YP. Fabrication and characterization of electrospun polycaprolactone blended with chitosan-gelatin complex
nanofibrous mats. J Nanomater 2014;2014:1.
[123] Sobreiro-Almeida R, Fonseca DR, Neves NM. Extracellular matrix electrospun membranes for mimicking natural renal filtration barriers. Mater Sci Eng C Mater
Biol Appl 2019;103:109866.
[124] Watcharajittanont N, Putson C, Pripatnanont P, Meesane J. Electrospun polyurethane fibrous membranes of mimicked extracellular matrix for periodontal
ligament: Molecular behavior, mechanical properties, morphology, and osseointegration. J Biomater Appl 2020;34:753–62.
[125] Unal S, Arslan S, Yilmaz BK, Oktar FN, Ficai D, Ficai A, et al. Polycaprolactone/gelatin/hyaluronic acid electrospun scaffolds to mimic glioblastoma extra­
cellular matrix. Materials (Basel) 2020;13:2661.
[126] Gholipourmalekabadi M, Mozafari M, Bandehpour M, Salehi M, Sameni M, Caicedo HH, et al. Optimization of nanofibrous silk fibroin scaffold as a delivery
system for bone marrow adherent cells: in vitro and in vivo studies. Biotechnol Appl Biochem 2015;62:785–94.
[127] Martins A, Chung S, Pedro AJ, Sousa RA, Marques AP, Reis RL, et al. Hierarchical starch-based fibrous scaffold for bone tissue engineering applications. J Tissue
Eng Regen Med 2009;3:37–42.
[128] Qu HW, Fu HY, Han ZY, Sun Y. Biomaterials for bone tissue engineering scaffolds: a review. RSC Adv 2019;9:26252–62.

35
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

[129] Haleem A, Javaid M, Khan RH, Suman R. 3D printing applications in bone tissue engineering. J Clin Orthop Trauma 2020;11:S118–24.
[130] Bose S, Roy M, Bandyopadhyay A. Recent advances in bone tissue engineering scaffolds. Trends Biotechnol 2012;30:546–54.
[131] Rouwkema J, Rivron NC, van Blitterswijk CA. Vascularization in tissue engineering. Trends Biotechnol 2008;26:434–41.
[132] Fielding GA, Bandyopadhyay A, Bose S. Effects of silica and zinc oxide doping on mechanical and biological properties of 3D printed tricalcium phosphate tissue
engineering scaffolds. Dent Mater 2012;28:113–22.
[133] Chocholata P, Kulda V, Babuska V. Fabrication of scaffolds for bone-tissue regeneration. Materials (Basel) 2019;12:568.
[134] Ranganathan S, Balagangadharan K, Selvamurugan N. Chitosan and gelatin-based electrospun fibers for bone tissue engineering. Int J Biol Macromol
2019;133:354–64.
[135] Toloue EB, Karbasi S, Salehi H, Rafienia M. Potential of an electrospun composite scaffold of poly (3-hydroxybutyrate)-chitosan/alumina nanowires in bone
tissue engineering applications. Mater Sci Eng C Mater Biol Appl 2019;99:1075–91.
[136] Dalgic AD, Atila D, Karatas A, Tezcaner A, Keskin D. Diatom shell incorporated PHBV/PCL-pullulan co-electrospun scaffold for bone tissue engineering. Mater
Sci Eng C Mater Biol Appl 2019;100:735–46.
[137] Nezafati N, Faridi-Majidi R, Pazouki M, Hesaraki S. Synthesis and characterization of a novel freeze-dried silanated chitosan bone tissue engineering scaffold
reinforced with electrospun hydroxyapatite nanofiber. Polym Int 2019;68:1420–9.
[138] Januariyasa IK, Ana ID, Yusuf Y. Nanofibrous poly(vinyl alcohol)/chitosan contained carbonated hydroxyapatite nanoparticles scaffold for bone tissue en­
gineering. Mater Sci Eng C-Mater Biol Appl 2020;107:110347.
[139] Fang R, Zhang EW, Xu L, Wei SC. Electrospun PCL/PLA/HA based nanofibers as scaffold for osteoblast-like cells. J Nanosci Nanotechnol 2010;10:7747–51.
[140] Chen P, Liu L, Pan J, Mei J, Li C, Zheng Y. Biomimetic composite scaffold of hydroxyapatite/gelatin-chitosan core-shell nanofibers for bone tissue engineering.
Mater Sci Eng C Mater Biol Appl 2019;97:325–35.
[141] Al-Wafi R, Mansour SF, Ahmed MK. Mechanical, microstructural properties and cell adhesion of Sr/Se-hydroxyapatite/graphene/polycaprolactone nanofibers.
J Thermoplast Compos Mater 2020;0892705720912781.
[142] Rajzer I, Rom M, Menaszek E, Pasierb P. Conductive PANI patterns on electrospun PCL/gelatin scaffolds modified with bioactive particles for bone tissue
engineering. Mater Lett 2015;138:60–3.
[143] Li C, Vepari C, Jin HJ, Kim HJ, Kaplan DL. Electrospun silk-BMP-2 scaffolds for bone tissue engineering. Biomaterials 2006;27:3115–24.
[144] Shin SH, Purevdorj O, Castano O, Planell JA, Kim HW. A short review: Recent advances in electrospinning for bone tissue regeneration. J Tissue Eng 2012;3.
2041731412443530.
[145] Wu JX, Qin XY, Miao C, He YB, Liang GM, Zhou D, et al. A honeycomb-cobweb inspired hierarchical core-shell structure design for electrospun silicon/carbon
fibers as lithium-ion battery anodes. Carbon 2016;98:582–91.
[146] Perumal G, Sivakumar PM, Nandkumar AM, Doble M. Synthesis of magnesium phosphate nanoflakes and its PCL composite electrospun nanofiber scaffolds for
bone tissue regeneration. Mater Sci Eng C Mater Biol Appl 2020;109:110527.
[147] Tahmasebi A, Enderami SE, Saburi E, Islami M, Yaslianifard S, Mahabadi JA, et al. Micro-RNA-incorporated electrospun nanofibers improve osteogenic
differentiation of human-induced pluripotent stem cells. J Biomed Mater Res A 2020;108:377–86.
[148] Sedghi R, Shaabani A, Sayyari N. Electrospun triazole-based chitosan nanofibers as a novel scaffolds for bone tissue repair and regeneration. Carbohydr Polym
2020;230:115707.
[149] Lee HJ, Abueva CD, Padalhin AR, Lee BT. Soya protein isolate-polyethylene oxide electrospun nanofiber membrane with bone marrow-derived mesenchymal
stem cell for enhanced bone regeneration. J Biomater Appl 2020;34:1142–9.
[150] Patel KD, Kim TH, Mandakhbayar N, Singh RK, Jang JH, Lee JH, et al. Coating biopolymer nanofibers with carbon nanotubes accelerates tissue healing and
bone regeneration through orchestrated cell- and tissue-regulatory responses. Acta Biomater 2020;108:97–110.
[151] Dhand C, Barathi VA, Ong ST, Venkatesh M, Harini S, Dwivedi N, et al. Latent oxidative polymerization of catecholamines as potential cross-linkers for
biocompatible and multifunctional biopolymer scaffolds. ACS Appl Mater Interfaces 2016;8:32266–81.
[152] Zhang HL, Chen ZQ. Fabrication and characterization of electrospun PLGA/MWNTs/hydroxyapatite biocomposite scaffolds for bone tissue engineering. J
Bioactive Compatible Polym 2010;25:241–59.
[153] Frohbergh ME, Katsman A, Botta GP, Lazarovici P, Schauer CL, Wegst UG, et al. Electrospun hydroxyapatite-containing chitosan nanofibers crosslinked with
genipin for bone tissue engineering. Biomaterials 2012;33:9167–78.
[154] Ki CS, Park SY, Kim HJ, Jung HM, Woo KM, Lee JW, et al. Development of 3-D nanofibrous fibroin scaffold with high porosity by electrospinning: implications
for bone regeneration. Biotechnol Lett 2008;30:405–10.
[155] Kim HW, Lee HH, Knowles JC. Electrospinning biomedical nanocomposite fibers of hydroxyapatite/poly(lactic acid) for bone regeneration. J Biomed Mater Res
A 2006;79:643–9.
[156] Lao LH, Wang YJ, Zhu Y, Zhang YY, Gao CY. Poly(lactide-co-glycolide)/hydroxyapatite nanofibrous scaffolds fabricated by electrospinning for bone tissue
engineering. J Mater Sci-Mater Med 2011;22:1873–84.
[157] Li X, Xie J, Yuan X, Xia Y. Coating electrospun poly (ε-caprolactone) fibers with gelatin and calcium phosphate and their use as biomimetic scaffolds for bone
tissue engineering. Langmuir 2008;24:14145–50.
[158] Zhang Y, Venugopal JR, El-Turki A, Ramakrishna S, Su B, Lim CT. Electrospun biomimetic nanocomposite nanofibers of hydroxyapatite/chitosan for bone
tissue engineering. Biomaterials 2008;29:4314–22.
[159] Liao S, Murugan R, Chan CK, Ramakrishna S. Processing nanoengineered scaffolds through electrospinning and mineralization, suitable for biomimetic bone
tissue engineering. J Mech Behav Biomed Mater 2008;1:252–60.
[160] Ji W, Yang F, Ma J, Bouma MJ, Boerman OC, Chen Z, et al. Incorporation of stromal cell-derived factor-1alpha in PCL/gelatin electrospun membranes for
guided bone regeneration. Biomaterials 2013;34:735–45.
[161] Liu Y, Cui H, Zhuang X, Wei Y, Chen X. Electrospinning of aniline pentamer-graft-gelatin/PLLA nanofibers for bone tissue engineering. Acta Biomater
2014;10:5074–80.
[162] Lu H, Zhang T, Wang XP, Fang QF. Electrospun submicron bioactive glass fibers for bone tissue scaffold. J Mater Sci Mater Med 2009;20:793–8.
[163] Nandakumar A, Yang L, Habibovic P, van Blitterswijk C. Calcium phosphate coated electrospun fiber matrices as scaffolds for bone tissue engineering. Langmuir
2009;26:7380–7.
[164] Ravichandran R, Venugopal JR, Sundarrajan S, Mukherjee S, Ramakrishna S. Precipitation of nanohydroxyapatite on PLLA/PBLG/Collagen nanofibrous
structures for the differentiation of adipose derived stem cells to osteogenic lineage. Biomaterials 2012;33:846–55.
[165] Norowski Jr. PA, Fujiwara T, Clem WC, Adatrow PC, Eckstein EC, Haggard WO, et al. Novel naturally crosslinked electrospun nanofibrous chitosan mats for
guided bone regeneration membranes: material characterization and cytocompatibility. J Tissue Eng Regen Med 2015;9:577–83.
[166] Pascu EI, Stokes J, McGuinness GB. Electrospun composites of PHBV, silk fibroin and nano-hydroxyapatite for bone tissue engineering. Mater Sci Eng C Mater
Biol Appl 2013;33:4905–16.
[167] Park SY, Ki CS, Park YH, Jung HM, Woo KM, Kim HJ. Electrospun silk fibroin scaffolds with macropores for bone regeneration: an in vitro and in vivo study.
Tissue Eng Part A 2010;16:1271–9.
[168] Prabhakaran MP, Venugopal J, Ramakrishna S. Electrospun nanostructured scaffolds for bone tissue engineering. Acta Biomater 2009;5:2884–93.
[169] Rajzer I, Menaszek E, Kwiatkowski R, Planell JA, Castano O. Electrospun gelatin/poly(epsilon-caprolactone) fibrous scaffold modified with calcium phosphate
for bone tissue engineering. Mater Sci Eng C Mater Biol Appl 2014;44:183–90.
[170] Ren J, Blackwood KA, Doustgani A, Poh PP, Steck R, Stevens MM, et al. Melt-electrospun polycaprolactone strontium-substituted bioactive glass scaffolds for
bone regeneration. J Biomed Mater Res A 2014;102:3140–53.
[171] Sadat-Shojai M. Electrospun Polyhydroxybutyrate/Hydroxyapatite Nanohybrids: Microstructure and Bone Cell Response. J Mater Sci Technol
2016;32:1013–20.
[172] Sajesh K, Kiran K, Nair SV, Jayakumar R. Sequential layer-by layer electrospinning of nano SrCO 3/PRP loaded PHBV fibrous scaffold for bone tissue

36
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

engineering. Compos Part B: Eng 2016.


[173] Samavedi S, Vaidya P, Gaddam P, Whittington AR, Goldstein AS. Electrospun meshes possessing region-wise differences in fiber orientation, diameter,
chemistry and mechanical properties for engineering bone-ligament-bone tissues. Biotechnol Bioeng 2014;111:2549–59.
[174] Shanmugavel S, Reddy VJ, Ramakrishna S, Lakshmi BS, Dev VG. Precipitation of hydroxyapatite on electrospun polycaprolactone/aloe vera/silk fibroin
nanofibrous scaffolds for bone tissue engineering. J Biomater Appl 2014;29:46–58.
[175] Yang W, Yang F, Wang Y, Both SK, Jansen JA. In vivo bone generation via the endochondral pathway on three-dimensional electrospun fibers. Acta Biomater
2013;9:4505–12.
[176] Gholipourmalekabadi M, Mozafari M, Gholipourmalekabadi M, Nazm Bojnordi M, Hashemi-Soteh MB, Salimi M, et al. In vitro and in vivo evaluations of three-
dimensional hydroxyapatite/silk fibroin nanocomposite scaffolds. Biotechnol Appl Biochem 2015;62:441–50.
[177] Sefat F, Raja TI, Zafar MS, Khurshid Z, Najeeb S, Zohaib S, et al. Nanoengineered biomaterials for cartilage repair. Nanoeng Biomater Regenerative Med:
Elsevier 2019:39–71.
[178] Rahmati M, Nalesso G, Mobasheri A, Mozafari M. Aging and osteoarthritis: Central role of the extracellular matrix. Ageing Res Rev 2017;40:20–30.
[179] Rahmati M, Mobasheri A, Mozafari M. Inflammatory mediators in osteoarthritis: A critical review of the state-of-the-art, current prospects, and future chal­
lenges. Bone 2016;85:81–90.
[180] Holmes B, Castro NJ, Zhang LG, Zussman E. Electrospun fibrous scaffolds for bone and cartilage tissue generation: recent progress and future developments.
Tissue Eng Part B Rev 2012;18:478–86.
[181] Wise JK, Yarin AL, Megaridis CM, Cho M. Chondrogenic differentiation of human mesenchymal stem cells on oriented nanofibrous scaffolds: engineering the
superficial zone of articular cartilage. Tissue Eng Part A 2008;15:913–21.
[182] Xue JX, Feng B, Zheng R, Lu Y, Zhou GD, Liu W, et al. Engineering ear-shaped cartilage using electrospun fibrous membranes of gelatin/polycaprolactone.
Biomaterials 2013;34:2624–31.
[183] Li WJ, Tuli R, Okafor C, Derfoul A, Danielson KG, Hall DJ, et al. A three-dimensional nanofibrous scaffold for cartilage tissue engineering using human
mesenchymal stem cells. Biomaterials 2005;26:599–609.
[184] Chen W, Wang C, Gao YY, Wu Y, Wu GM, Shi XW, et al. Incorporating chitin derived glucosamine sulfate into nanofibers via coaxial electrospinning for
cartilage regeneration. Carbohydr Polym 2020;229:115544.
[185] Semitela A, Girao AF, Fernandes C, Ramalho G, Bdikin I, Completo A, et al. Electrospinning of bioactive polycaprolactone-gelatin nanofibres with increased
pore size for cartilage tissue engineering applications. J Biomater Appl 2020;885328220940194.
[186] Silva JC, Udangawa RN, Chen J, Mancinelli CD, Garrudo FFF, Mikael PE, et al. Kartogenin-loaded coaxial PGS/PCL aligned nanofibers for cartilage tissue
engineering. Mater Sci Eng C Mater Biol Appl 2020;107:110291.
[187] Alves da Silva M, Martins A, Costa-Pinto A, Costa P, Faria S, Gomes M, et al. Cartilage tissue engineering using electrospun PCL nanofiber meshes and MSCs.
Biomacromolecules 2010;11:3228–36.
[188] Garrigues NW, Little D, Sanchez-Adams J, Ruch DS, Guilak F. Electrospun cartilage-derived matrix scaffolds for cartilage tissue engineering. J Biomed Mater
Res A 2014;102:3998–4008.
[189] He X, Fu W, Feng B, Wang H, Liu Z, Yin M, et al. Electrospun collagen–poly (l-lactic acid-co-ε-caprolactone) membranes for cartilage tissue engineering.
Regenerative Med 2013;8:425–36.
[190] Kim C, Shores L, Guo Q, Aly A, Jeon OH, Kim DH, et al. Electrospun microfiber scaffolds with anti-inflammatory tributanoylated N-acetyl-d-glucosamine
promote cartilage regeneration. Tissue Eng Part A 2016;22:689–97.
[191] Levorson EJ, Raman Sreerekha P, Chennazhi KP, Kasper FK, Nair SV, Mikos AG. Fabrication and characterization of multiscale electrospun scaffolds for
cartilage regeneration. Biomed Mater 2013;8:014103.
[192] Moroni L, Schotel R, Hamann D, de Wijn JR, van Blitterswijk CA. 3D fiber-deposited electrospun integrated scaffolds enhance cartilage tissue formation. Adv
Funct Mater 2008;18:53–60.
[193] Mouthuy PA, El-Sherbini Y, Cui Z, Ye H. Layering PLGA-based electrospun membranes and cell sheets for engineering cartilage-bone transition. J Tissue Eng
Regen Med 2016;10:E263–74.
[194] Reboredo JW, Weigel T, Steinert A, Rackwitz L, Rudert M, Walles H. Investigation of migration and differentiation of human mesenchymal stem cells on five-
layered collagenous electrospun scaffold mimicking native cartilage structure. Adv Healthc Mater 2016;5:2191–8.
[195] Sambudi NS, Sathyamurthy M, Lee GM, Park SB. Electrospun chitosan/poly (vinyl alcohol) reinforced with CaCO 3 nanoparticles with enhanced mechanical
properties and biocompatibility for cartilage tissue engineering. Compos Sci Technol 2015;106:76–84.
[196] Weijie Z, Zhuo C, Sujuan M, Yonggang W, Fei Z, Keyi W, et al. Cistanche polysaccharide (CDPS)/polylactic acid (PLA) scaffolds based coaxial electrospinning
for vascular tissue engineering. Int J Polym Mater Polym Biomater 2016;65:38–46.
[197] Shafiee A, Soleimani M, Chamheidari GA, Seyedjafari E, Dodel M, Atashi A, et al. Electrospun nanofiber-based regeneration of cartilage enhanced by me­
senchymal stem cells. J Biomed Mater Res A 2011;99:467–78.
[198] Shin HJ, Lee CH, Cho IH, Kim YJ, Lee YJ, Kim IA, et al. Electrospun PLGA nanofiber scaffolds for articular cartilage reconstruction: mechanical stability,
degradation and cellular responses under mechanical stimulation in vitro. J Biomater Sci Polym Ed 2006;17:103–19.
[199] Subramanian A, Vu D, Larsen GF, Lin HY. Preparation and evaluation of the electrospun chitosan/PEO fibers for potential applications in cartilage tissue
engineering. J Biomater Sci Polym Ed 2005;16:861–73.
[200] Toyokawa N, Fujioka H, Kokubu T, Nagura I, Inui A, Sakata R, et al. Electrospun synthetic polymer scaffold for cartilage repair without cultured cells in an
animal model. Arthroscopy – J Arthroscopic Related Surg 2010;26:375–83.
[201] Wang ZL, Wang Y, Zhang PB, Chen XS. Methylsulfonylmethane-loaded electrospun poly(lactide-co-glycolide) mats for cartilage tissue engineering. RSC Adv
2015;5:96725–32.
[202] Yang Y, Yang Q, Zhou F, Zhao Y, Jia X, Yuan X, et al. Electrospun PELCL membranes loaded with QK peptide for enhancement of vascular endothelial cell
growth. J Mater Sci Mater Med 2016;27:106.
[203] Wanjare M, Kusuma S, Gerecht S. Perivascular cells in blood vessel regeneration. Biotechnol J 2013;8:434–47.
[204] Fraisl P, Mazzone M, Schmidt T, Carmeliet P. Regulation of angiogenesis by oxygen and metabolism. Dev Cell 2009;16:167–79.
[205] Brey EM. Vascularization: regenerative medicine and tissue engineering. CRC Press; 2014.
[206] Rhodin JA. Architecture of the vessel wall. Comprehensive Physiol 2011.
[207] Thottappillil N, Nair PD. Scaffolds in vascular regeneration: current status. Vascular Health Risk Manage 2014;11:79–91.
[208] Yazdanpanah A, Tahmasbi M, Amoabediny G, Nourmohammadi J, Moztarzadeh F, Mozafari M. Fabrication and characterization of electrospun poly-L-lactide/
gelatin graded tubular scaffolds: Toward a new design for performance enhancement in vascular tissue engineering. Prog Nat Sci-Mater Int 2015;25:405–13.
[209] Zhang X, Baughman CB, Kaplan DL. In vitro evaluation of electrospun silk fibroin scaffolds for vascular cell growth. Biomaterials 2008;29:2217–27.
[210] Hasan A, Memic A, Annabi N, Hossain M, Paul A, Dokmeci MR, et al. Electrospun scaffolds for tissue engineering of vascular grafts. Acta Biomater
2014;10:11–25.
[211] Mi HY, Jiang Y, Jing X, Enriquez E, Li H, Li Q, et al. Fabrication of triple-layered vascular grafts composed of silk fibers, polyacrylamide hydrogel, and
polyurethane nanofibers with biomimetic mechanical properties. Mater Sci Eng C Mater Biol Appl 2019;98:241–9.
[212] Hu Q, Su C, Zeng Z, Zhang H, Feng R, Feng J, et al. Fabrication of multilayer tubular scaffolds with aligned nanofibers to guide the growth of endothelial cells. J
Biomater Appl 2020. 885328220935090.
[213] Wang HY, Feng YK, Fang ZC, Yuan WJ, Khan M. Co-electrospun blends of PU and PEG as potential biocompatible scaffolds for small-diameter vascular tissue
engineering. Mater Sci Eng C – Mater Biol Appl 2012;32:2306–15.
[214] Dargaville BL, Vaquette C, Rasoul F, Cooper-White JJ, Campbell JH, Whittaker AK. Electrospinning and crosslinking of low-molecular-weight poly(tri­
methylene carbonate-co-L-lactide) as an elastomeric scaffold for vascular engineering. Acta Biomater 2013;9:6885–97.
[215] Wong CS, Liu X, Xu ZG, Lin T, Wang XG. Elastin and collagen enhances electrospun aligned polyurethane as scaffolds for vascular graft. J Mater Sci-Mater Med

37
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

2013;24:1865–74.
[216] Ahn H, Ju YM, Takahashi H, Williams DF, Yoo JJ, Lee SJ, et al. Engineered small diameter vascular grafts by combining cell sheet engineering and electro­
spinning technology. Acta Biomater 2015;16:14–22.
[217] Catto V, Fare S, Cattaneo I, Figliuzzi M, Alessandrino A, Freddi G, et al. Small diameter electrospun silk fibroin vascular grafts: Mechanical properties, in vitro
biodegradability, and in vivo biocompatibility. Mater Sci Eng C Mater Biol Appl 2015;54:101–11.
[218] Hasan A, Deeb G, Atwi K, Soliman S. Electrospun PET-PU scaffolds for vascular tissue engineering. Advances in biomedical engineering (ICABME), 2015
international conference on: IEEE. 2015. p. 217–21.
[219] Jing X, Mi HY, Salick MR, Cordie TM, Peng XF, Turng LS. Electrospinning thermoplastic polyurethane/graphene oxide scaffolds for small diameter vascular
graft applications. Mater Sci Eng C Mater Biol Appl 2015;49:40–50.
[220] Yazdanpanah A, Amoabediny G, Shariatpanahi P, Nourmohammadi J, Tahmasbi M, Mozafari M. Synthesis and characterization of polylactic acid tubular
scaffolds with improved mechanical properties for vascular tissue engineering. Trends Biomater Artif Organs 2014;28:99.
[221] Javanmard SH, Anari J, Kharazi AZ, Vatankhah E. In vitro hemocompatibility and cytocompatibility of a three-layered vascular scaffold fabricated by se­
quential electrospinning of PCL, collagen, and PLLA nanofibers. J Biomater Appl 2016. 0885328216652068.
[222] Duan N, Geng X, Ye L, Zhang A, Feng Z, Guo L, et al. A vascular tissue engineering scaffold with core-shell structured nano-fibers formed by coaxial elec­
trospinning and its biocompatibility evaluation. Biomed Mater 2016;11:035007.
[223] Elsayed Y, Lekakou C, Labeed F, Tomlins P. Fabrication and characterisation of biomimetic, electrospun gelatin fibre scaffolds for tunica media-equivalent,
tissue engineered vascular grafts. Mater Sci Eng C Mater Biol Appl 2016;61:473–83.
[224] Gao YH, Yi T, Shinoka T, Lee YU, Reneker DH, Breuer CK, et al. Pilot mouse study of 1 mm inner diameter (ID) vascular graft using electrospun poly(ester urea)
nanofibers. Adv Healthcare Mater 2016;5:2427–36.
[225] Gong W, Lei D, Li S, Huang P, Qi Q, Sun Y, et al. Hybrid small-diameter vascular grafts: Anti-expansion effect of electrospun poly epsilon-caprolactone on
heparin-coated decellularized matrices. Biomaterials 2016;76:359–70.
[226] Lee B, Shafiq M, Jung Y, Park JC, Kim S. Characterization and preparation of bio-tubular scaffolds for fabricating artificial vascular grafts by combining
electrospinning and a co-culture system. Macromol Res 2016;24:131–42.
[227] Michael E, Abeyrathna N, Patel AV, Liao Y, Bashur CA. Incorporation of photo-carbon monoxide releasing materials into electrospun scaffolds for vascular
tissue engineering. Biomed Mater 2016;11:025009.
[228] Tan Z, Wang H, Gao X, Liu T, Tan Y. Composite vascular grafts with high cell infiltration by co-electrospinning. Mater Sci Eng C Mater Biol Appl
2016;67:369–77.
[229] Wang L, Wu Y, Hu T, Guo B, Ma PX. Electrospun conductive nanofibrous scaffolds for engineering cardiac tissue and 3D bioactuators. Acta Biomater
2017;59:68–81.
[230] Dippold D, Tallawi M, Tansaz S, Roether JA, Boccaccini AR. Novel electrospun poly(glycerol sebacate)-zein fiber mats as candidate materials for cardiac tissue
engineering. Eur Polym J 2016;75:504–13.
[231] Liu Y, Xu Y, Wang Z, Wen D, Zhang W, Schmull S, et al. Electrospun nanofibrous sheets of collagen/elastin/polycaprolactone improve cardiac repair after
myocardial infarction. Am J Transl Res 2016;8:1678–94.
[232] Gonzalez de Torre I, Alonso M, Rodriguez-Cabello JC. Elastin-based materials: promising candidates for cardiac tissue regeneration. Front Bioeng Biotechnol
2020;8:657.
[233] Zong X, Bien H, Chung CY, Yin L, Fang D, Hsiao BS, et al. Electrospun fine-textured scaffolds for heart tissue constructs. Biomaterials 2005;26:5330–8.
[234] Shin M, Ishii O, Sueda T, Vacanti JP. Contractile cardiac grafts using a novel nanofibrous mesh. Biomaterials 2004;25:3717–23.
[235] Castellano D, Blanes M, Marco B, Cerrada I, Ruiz-Sauri A, Pelacho B, et al. A comparison of electrospun polymers reveals poly(3-hydroxybutyrate) fiber as a
superior scaffold for cardiac repair. Stem Cells Dev 2014;23:1479–90.
[236] Balguid A, Mol A, van Marion MH, Bank RA, Bouten CV, Baaijens FP. Tailoring fiber diameter in electrospun poly (ɛ-caprolactone) scaffolds for optimal cellular
infiltration in cardiovascular tissue engineering. Tissue Eng Part A 2008;15:437–44.
[237] Kai D, Prabhakaran MP, Jin G, Ramakrishna S. Guided orientation of cardiomyocytes on electrospun aligned nanofibers for cardiac tissue engineering. J Biomed
Mater Res B Appl Biomater 2011;98:379–86.
[238] Prabhakaran MP, Nair AS, Kai D, Ramakrishna S. Electrospun composite scaffolds containing poly(octanediol-co-citrate) for cardiac tissue engineering.
Biopolymers 2012;97:529–38.
[239] Bhaarathy V, Venugopal J, Gandhimathi C, Ponpandian N, Mangalaraj D, Ramakrishna S. Biologically improved nanofibrous scaffolds for cardiac tissue
engineering. Mater Sci Eng C Mater Biol Appl 2014;44:268–77.
[240] Punnoose AM, Elamparithi A, Kuruvilla S. Electrospun type 1 collagen matrices using a novel benign solvent for cardiac tissue engineering. J Cell Physiol 2015.
[241] Ayaz HGS, Perets A, Ayaz H, Gilroy KD, Govindaraj M, Brookstein D, et al. Textile-templated electrospun anisotropic scaffolds for regenerative cardiac tissue
engineering. Biomaterials 2014;35:8540–52.
[242] Elamparithi A, Punnoose AM, Kuruvilla S. Gelatin electrospun nanofibrous matrices for cardiac tissue engineering applications. Int J Polym Mater Polym
Biomater 2016.
[243] Guex AG, Frobert A, Valentin J, Fortunato G, Hegemann D, Cook S, et al. Plasma-functionalized electrospun matrix for biograft development and cardiac
function stabilization. Acta Biomater 2014;10:2996–3006.
[244] Tallawi M, Dippold D, Rai R, D'Atri D, Roether JA, Schubert DW, et al. Novel PGS/PCL electrospun fiber mats with patterned topographical features for cardiac
patch applications. Mater Sci Eng C Mater Biol Appl 2016;69:569–76.
[245] Anesti K, Caine P. Peripheral nerve injuries. Plast Aesth Res 2015;6:309–10.
[246] Wang ML, Rivlin M, Graham JG, Beredjiklian PK. Peripheral nerve injury, scarring, and recovery. Connect Tissue Res 2019;60:3–9.
[247] Zhang PX, Han N, Kou YH, Zhu QT, Liu XL, Quan DP, et al. Tissue engineering for the repair of peripheral nerve injury. Neural Regen Res 2019;14:51–8.
[248] Ichihara S, Inada Y, Nakamura T. Artificial nerve tubes and their application for repair of peripheral nerve injury: an update of current concepts. Injury-Int J
Care Injured 2008;39:S29–39.
[249] Yi S, Xu L, Gu X. Scaffolds for peripheral nerve repair and reconstruction. Exp Neurol 2019;319:112761.
[250] Jaswal R, Shrestha S, Shrestha BK, Kumar D, Park CH, Kim CS. Nanographene enfolded AuNPs sophisticatedly synchronized polycaprolactone based elec­
trospun nanofibre scaffold for peripheral nerve regeneration. Mater Sci Eng: C 2020(111213).
[251] Yen C-M, Shen C-C, Yang Y-C, Liu B-S, Lee H-T, Sheu M-L, et al. Novel electrospun poly (ε-caprolactone)/type I collagen nanofiber conduits for repair of
peripheral nerve injury. Neural Regener Res 2019;14:1617.
[252] Entekhabi E, Haghbin Nazarpak M, Shafieian M, Mohammadi H, Firouzi M, Hassannejad Z. Fabrication and in vitro evaluation of 3D composite scaffold based
on collagen/hyaluronic acid sponge and electrospun PCL nanofibers for peripheral nerve regeneration. J Biomed Mater Res Part A 2020.
[253] Kijeńska E, Prabhakaran MP, Swieszkowski W, Kurzydlowski KJ, Ramakrishna S. Interaction of Schwann cells with laminin encapsulated PLCL core–shell
nanofibers for nerve tissue engineering. Eur Polym J 2014;50:30–8.
[254] Moroder P, Runge MB, Wang HA, Ruesink T, Lu LC, Spinner RJ, et al. Material properties and electrical stimulation regimens of polycaprolactone fumarate-
polypyrrole scaffolds as potential conductive nerve conduits. Acta Biomater 2011;7:944–53.
[255] Xu H, Holzwarth JM, Yan Y, Xu P, Zheng H, Yin Y, et al. Conductive PPY/PDLLA conduit for peripheral nerve regeneration. Biomaterials 2014;35:225–35.
[256] Atoufi Z, Zarrintaj P, Motlagh GH, Amiri A, Bagher Z, Kamrava SK. A novel bio electro active alginate-aniline tetramer/ agarose scaffold for tissue engineering:
synthesis, characterization, drug release and cell culture study. J Biomater Sci Polym Ed 2017;28:1617–38.
[257] Tian L, Prabhakaran MP, Hu J, Chen M, Besenbacher F, Ramakrishna S. Synergistic effect of topography, surface chemistry and conductivity of the electrospun
nanofibrous scaffold on cellular response of PC12 cells. Colloids Surf B Biointerfaces 2016;145:420–9.
[258] Song J, Sun B, Liu S, Chen W, Zhang Y, Wang C, et al. Polymerizing pyrrole coated poly (l-lactic acid-co-epsilon-caprolactone) (PLCL) conductive nanofibrous
conduit combined with electric stimulation for long-range peripheral nerve regeneration. Front Mol Neurosci 2016;9:117.

38
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

[259] Wang HB, Mullins ME, Cregg JM, McCarthy CW, Gilbert RJ. Varying the diameter of aligned electrospun fibers alters neurite outgrowth and Schwann cell
migration. Acta Biomater 2010;6:2970–8.
[260] Bini TB, Gao SJ, Tan TC, Wang S, Lim A, Hai LB, et al. Electrospun poly(L-lactide-co-glycolide) biodegradable polymer nanofibre tubes for peripheral nerve
regeneration. Nanotechnology 2004;15:1459–64.
[261] Ghasemi-Mobarakeh L, Prabhakaran MP, Morshed M, Nasr-Esfahani M-H, Ramakrishna S. Electrospun poly (ɛ-caprolactone)/gelatin nanofibrous scaffolds for
nerve tissue engineering. Biomaterials 2008;29:4532–9.
[262] Jiang X, Mi R, Hoke A, Chew SY. Nanofibrous nerve conduit-enhanced peripheral nerve regeneration. J Tissue Eng Regen Med 2014;8:377–85.
[263] Kijeńska E, Prabhakaran MP, Swieszkowski W, Kurzydlowski KJ, Ramakrishna S. Electrospun bio-composite P (LLA-CL)/collagen I/collagen III scaffolds for
nerve tissue engineering. J Biomed Mater Res B Appl Biomater 2012;100:1093–102.
[264] Wang W, Itoh S, Konno K, Kikkawa T, Ichinose S, Sakai K, et al. Effects of Schwann cell alignment along the oriented electrospun chitosan nanofibers on nerve
regeneration. J Biomed Mater Res A 2009;91:994–1005.
[265] Prabhakaran MP, Ghasemi-Mobarakeh L, Jin G, Ramakrishna S. Electrospun conducting polymer nanofibers and electrical stimulation of nerve stem cells. J
Biosci Bioeng 2011;112:501–7.
[266] Park SY, Ki CS, Park YH, Lee KG, Kang SW, Kweon HY, et al. Functional recovery guided by an electrospun silk fibroin conduit after sciatic nerve injury in rats. J
Tissue Eng Regen Med 2015;9:66–76.
[267] Prabhakaran MP, Venugopal JR, Ramakrishna S. Mesenchymal stem cell differentiation to neuronal cells on electrospun nanofibrous substrates for nerve tissue
engineering. Biomaterials 2009;30:4996–5003.
[268] Hong S, Kim G. Electrospun micro/nanofibrous conduits composed of poly(epsilon-caprolactone) and small intestine submucosa powder for nerve tissue
regeneration. J Biomed Mater Res B Appl Biomater 2010;94:421–8.
[269] Meng LH, Klinkajon W, K-hasuwan PR, Harkin S, Supaphol P, Wnek GE. Electrospun crosslinked poly(acrylic acid) fiber constructs: towards a synthetic model
of the cortical layer of nerve. Polym Int 2015;64:42–8.
[270] Prabhakaran MP, Vatankhah E, Ramakrishna S. Electrospun aligned PHBV/collagen nanofibers as substrates for nerve tissue engineering. Biotechnol Bioeng
2013;110:2775–84.
[271] Chew SY, Mi R, Hoke A, Leong KW. Aligned protein-polymer composite fibers enhance nerve regeneration: a potential tissue-engineering platform. Adv Funct
Mater 2007;17:1288–96.
[272] Alvarez-Perez MA, Guarino V, Cirillo V, Ambrosio L. Influence of gelatin cues in PCL electrospun membranes on nerve outgrowth. Biomacromolecules
2010;11:2238–46.
[273] Hu J, Kai D, Ye H, Tian L, Ding X, Ramakrishna S, et al. Electrospinning of poly(glycerol sebacate)-based nanofibers for nerve tissue engineering. Mater Sci Eng
C Mater Biol Appl 2017;70:1089–94.
[274] Rahmati M, Blaker JJ, Lyngstadaas SP, Mano JF, Haugen HJ. Designing multigradient biomaterials for skin regeneration. Mater Today Adv 2020;5:100051.
[275] Mayet N, Choonara YE, Kumar P, Tomar LK, Tyagi C, Du Toit LC, et al. A comprehensive review of advanced biopolymeric wound healing systems. J Pharm Sci
2014;103:2211–30.
[276] Adeli-Sardou M, Yaghoobi MM, Torkzadeh-Mahani M, Dodel M. Controlled release of lawsone from polycaprolactone/gelatin electrospun nano fibers for skin
tissue regeneration. Int J Biol Macromol 2019;124:478–91.
[277] Fathi HA, Abdelkader A, AbdelKarim MS, Abdelaziz AA, El-Mokhtar MA, Allam A, et al. Electrospun vancomycin-loaded nanofibers for management of
methicillin-resistant Staphylococcus aureus-induced skin infections. Int J Pharm 2020;586:119620.
[278] Mahanty B, Maity K, Sarkar S, Mandal D. Human skin interactive self-powered piezoelectric e-skin based on PVDF/MWCNT electrospun nanofibers for non-
invasive health care monitoring. Mater Today-Proc 2020;21:1964–8.
[279] Paggiaro AO, Bastianelli R, Carvalho VF, Isaac C, Gemperli R. Is allograft skin, the gold-standard for burn skin substitute? A systematic literature review and
meta-analysis. J Plastic, Reconstruct Aesthetic Surg 2019;72:1245–53.
[280] Filippova OV, Afonochev KA, Nikitin MS, Govorov AV. Long-term results of plasty of granulating wounds of the distal extremities with mesh and solid skin
autografts of burn wounds in children. Pediatric Traumatol, Orthopaedics Reconstruct Surg 2019;7:35–44.
[281] Atiyeh BS, Costagliola M. Cultured epithelial autograft (CEA) in burn treatment: Three decades later. Burns 2007;33:405–13.
[282] Clark RAF, Ghosh K, Tonnesen MG. Tissue Engineering for Cutaneous Wounds. J, Invest Dermatol 2007;127:1018–29.
[283] MacNeil S. Progress and opportunities for tissue-engineered skin. Nature 2007;445:874.
[284] Przekora A. A concise review on tissue engineered artificial skin grafts for chronic wound treatment: can we reconstruct functional skin tissue in vitro? Cells
2020;9:1622.
[285] Kim HS, Sun X, Lee JH, Kim HW, Fu X, Leong KW. Advanced drug delivery systems and artificial skin grafts for skin wound healing. Adv Drug Deliv Rev
2019;146:209–39.
[286] Debels H, Hamdi M, Abberton K, Morrison W. Dermal matrices and bioengineered skin substitutes: a critical review of current options. Plast Reconstruct Surg
Global Open 2015;3:e284.
[287] Lee K, Silva EA, Mooney DJ. Growth factor delivery-based tissue engineering: general approaches and a review of recent developments. J R Soc Interface
2011;8:153–70.
[288] Mitchell AC, Briquez PS, Hubbell JA, Cochran JR. Engineering growth factors for regenerative medicine applications. Acta Biomater 2016;30:1–12.
[289] Mogosanu GD, Grumezescu AM. Natural and synthetic polymers for wounds and burns dressing. Int J Pharm 2014;463:127–36.
[290] Milan PB, Lotfibakhshaiesh N, Joghataie M, Ai J, Pazouki A, Kaplan D, et al. Accelerated wound healing in a diabetic rat model using decellularized dermal
matrix and human umbilical cord perivascular cells. Acta Biomater 2016;45:234–46.
[291] Khil MS, Cha DI, Kim HY, Kim IS, Bhattarai N. Electrospun nanofibrous polyurethane membrane as wound dressing. J Biomed Mater Res Part B-Appl Biomater
2003;67b:675–9.
[292] Zhang Y, Lim CT, Ramakrishna S, Huang ZM. Recent development of polymer nanofibers for biomedical and biotechnological applications. J Mater Sci Mater
Med. 2005;16:933–46.
[293] Hadisi Z, Nourmohammadi J, Nassiri SM. The antibacterial and anti-inflammatory investigation of Lawsonia Inermis-gelatin-starch nano-fibrous dressing in
burn wound. Int J Biol Macromol 2017.
[294] Mirdailami O, Soleimani M, Dinarvand R, Khoshayand MR, Norouzi M, Hajarizadeh A, et al. Controlled release of rh EGF and rhb FGF from electrospun
scaffolds for skin regeneration. J Biomed Mater Res Part A 2015;103:3374–85.
[295] Norouzi M, Shabani I, Ahvaz HH, Soleimani M. PLGA/gelatin hybrid nanofibrous scaffolds encapsulating EGF for skin regeneration. J Biomed Mater Res A.
2015;103:2225–35.
[296] Dias JR, Baptista-Silva S, de Oliveira CMT, Sousa A, Oliveira AL, Bartolo PJ, et al. In situ crosslinked electrospun gelatin nanofibers for skin regeneration. Eur
Polym J 2017;95:161–73.
[297] Abdul Khodir WKW, Abdul Razak AH, Ng MH, Guarino V, Susanti D. Encapsulation and characterization of gentamicin sulfate in the collagen added electrospun
nanofibers for skin regeneration. J Funct Biomater. 2018;9:36.
[298] Zhou T, Sui B, Mo X, Sun J. Multifunctional and biomimetic fish collagen/bioactive glass nanofibers: fabrication, antibacterial activity and inducing skin
regeneration in vitro and in vivo. Int J Nanomed 2017;12:3495–507.
[299] Sun XM, Lang Q, Zhang HB, Cheng LY, Zhang Y, Pan GQ, et al. Electrospun Photocrosslinkable Hydrogel Fibrous Scaffolds for Rapid In Vivo Vascularized Skin
Flap Regeneration. Adv Funct Mater 2017;27:1604617.
[300] Arasteh S, Kazemnejad S, Khanjani S, Heidari-Vala H, Akhondi MM, Mobini S. Fabrication and characterization of nano-fibrous bilayer composite for skin
regeneration application. Methods 2016;99:3–12.
[301] Augustine R, Dominic EA, Reju I, Kaimal B, Kalarikkal N, Thomas S. Electrospun poly(epsilon-caprolactone)-based skin substitutes: In vivo evaluation of wound
healing and the mechanism of cell proliferation. J Biomed Mater Res B Appl Biomater. 2015;103:1445–54.

39
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

[302] Sharif S, Ai J, Azami M, Verdi J, Atlasi MA, Shirian S, et al. Collagen-coated nano-electrospun PCL seeded with human endometrial stem cells for skin tissue
engineering applications. J Biomed Mater Res B Appl Biomater. 2018;106:1578–86.
[303] Salehi M, Farzamfar S, Bastami F, Tajerian R. Fabrication and characterization of electrospun plla/collagen nanofibrous scaffold coated with chitosan to sustain
release of aloe vera gel for skin tissue engineering. Biomed Eng-Appl Basis Commun 2016;28:1650035.
[304] Wang Z, Qian Y, Li L, Pan L, Njunge LW, Dong L, et al. Evaluation of emulsion electrospun polycaprolactone/hyaluronan/epidermal growth factor nanofibrous
scaffolds for wound healing. J Biomater Appl. 2016;30:686–98.
[305] Wu C, Chen T, Xin Y, Zhang Z, Ren Z, Lei J, et al. Nanofibrous asymmetric membranes self-organized from chemically heterogeneous electrospun mats for skin
tissue engineering. Biomed Mater. 2016;11:035019.
[306] Merlettini A. Micro-nanostructured polymeric materials with specific functionalities for advanced biomedical applications: alma; 2019.
[307] Sabra S, Ragab DM, Agwa MM, Rohani S. Recent advances in electrospun nanofibers for some biomedical applications. Eur J Pharm Sci 2020;144:105224.
[308] Biondi M, Ungaro F, Quaglia F, Netti PA. Controlled drug delivery in tissue engineering. Adv Drug Deliv Rev. 2008;60:229–42.
[309] Mourino V, Boccaccini AR. Bone tissue engineering therapeutics: controlled drug delivery in three-dimensional scaffolds. J R Soc Interface. 2010;7:209–27.
[310] Wen S, Hu Y, Zhang Y, Huang S, Zuo Y, Min Y. Dual-functional core-shell electrospun mats with precisely controlled release of anti-inflammatory and anti-
bacterial agents. Mater Sci Eng C Mater Biol Appl. 2019;100:514–22.
[311] Amiri S, Rahimi A. Poly (ε-caprolactone) electrospun nanofibers containing cinnamon essential oil nanocapsules: a promising technique for controlled release
and high solubility. J Ind Text 2019;48:1527–44.
[312] Cheng X, Cheng G, Xing X, Yin CC, Cheng Y, Zhou X, et al. Controlled release of adenosine from core-shell nanofibers to promote bone regeneration through
STAT3 signaling pathway. J Control Release 2020;319:234–45.
[313] Zilberman M, Elsner JJ. Antibiotic-eluting medical devices for various applications. J Control Release 2008;130:202–15.
[314] Xu X, Yang L, Xu X, Wang X, Chen X, Liang Q, et al. Ultrafine medicated fibers electrospun from W/O emulsions. J Control Release. 2005;108:33–42.
[315] Hafshejani TM, Zamanian A, Venugopal JR, Rezvani Z, Sefat F, Saeb MR, et al. Antibacterial glass-ionomer cement restorative materials: A critical review on the
current status of extended release formulations. J Control Release. 2017;262:317–28.
[316] Quiles-Carrillo L, Montanes N, Lagaron JM, Balart R, Torres-Giner S. Bioactive multilayer polylactide films with controlled release capacity of gallic acid
accomplished by incorporating electrospun nanostructured coatings and interlayers. Appl Sci-Basel. 2019;9:533.
[317] Farboudi A, Nouri A, Shirinzad S, Sojoudi P, Davaran S, Akrami M, et al. Synthesis of magnetic gold coated poly (epsilon-caprolactonediol) based polyurethane/
poly(N-isopropylacrylamide)-grafted-chitosan core-shell nanofibers for controlled release of paclitaxel and 5-FU. Int J Biol Macromol. 2020;150:1130–40.
[318] Jiang H, Hu Y, Li Y, Zhao P, Zhu K, Chen W. A facile technique to prepare biodegradable coaxial electrospun nanofibers for controlled release of bioactive
agents. J Control Release. 2005;108:237–43.
[319] Eskitoros-Togay ŞM, Bulbul YE, Dilsiz N. Controlled release of doxycycline within core/shell poly (ε-caprolactone)/poly (ethylene oxide) fibers via coaxial
electrospinning. J Appl Polym Sci 2020;49273.
[320] Ma PX, Zhang R. Synthetic nano-scale fibrous extracellular matrix. J Biomed Mater Res. 1999;46:60–72.
[321] Wang J, Windbergs M. Controlled dual drug release by coaxial electrospun fibers - Impact of the core fluid on drug encapsulation and release. Int J Pharm
2019;556:363–71.
[322] Zhu LF, Chen XL, Ahmad Z, Li JS, Chang MW. Engineering of Ganoderma lucidum polysaccharide loaded polyvinyl alcohol nanofibers for biopharmaceutical
delivery. J Drug Delivery Sci Technol 2019;50:208–16.
[323] Kolambkar YM, Dupont KM, Boerckel JD, Huebsch N, Mooney DJ, Hutmacher DW, et al. An alginate-based hybrid system for growth factor delivery in the
functional repair of large bone defects. Biomaterials 2011;32:65–74.
[324] Diab T, Pritchard EM, Uhrig BA, Boerckel JD, Kaplan DL, Guldberg RE. A silk hydrogel-based delivery system of bone morphogenetic protein for the treatment
of large bone defects. J Mech Behav Biomed Mater. 2012;11:123–31.
[325] Hua D, Liu Z, Wang F, Gao B, Chen F, Zhang Q, et al. pH responsive polyurethane (core) and cellulose acetate phthalate (shell) electrospun fibers for
intravaginal drug delivery. Carbohydr Polym. 2016;151:1240–4.
[326] Mendes AC, Gorzelanny C, Halter N, Schneider SW, Chronakis IS. Hybrid electrospun chitosan-phospholipids nanofibers for transdermal drug delivery. Int J
Pharm. 2016;510:48–56.
[327] Tamayol A, Hassani Najafabadi A, Mostafalu P, Yetisen AK, Commotto M, Aldhahri M, et al. Biodegradable elastic nanofibrous platforms with integrated
flexible heaters for on-demand drug delivery. Sci Rep. 2017;7:9220.
[328] Wu C, Zhang H, Hu Q, Ramalingam M. Designing biomimetic triple-layered nanofibrous vascular grafts via combinatorial electrospinning approach. J Nanosci
Nanotechnol. 2020;20:6396–405.
[329] Zhao Y, Tavares AC, Gauthier MA. Nano-engineered electro-responsive drug delivery systems. J Mater Chem B. 2016;4:3019–30.
[330] Rim NG, Shin CS, Shin H. Current approaches to electrospun nanofibers for tissue engineering. Biomed Mater. 2013;8:014102.
[331] Wu X-Q, Wu X, Wang T-Y, Zhao L, Truong YB, Ng D, et al. Omniphobic surface modification of electrospun nanofiber membrane via vapor deposition for
enhanced anti-wetting property in membrane distillation. J Membr Sci 2020:118075.
[332] Cui Z, Lin J, Zhan C, Wu J, Shen S, Si J, et al. Biomimetic composite scaffolds based on surface modification of polydopamine on ultrasonication induced
cellulose nanofibrils (CNF) adsorbing onto electrospun thermoplastic polyurethane (TPU) nanofibers. J Biomater Sci Polym Ed. 2020;31:561–77.
[333] Palo M, Ronkonharju S, Tiirik K, Viidik L, Sandler N, Kogermann K. Bi-layered polymer carriers with surface modification by electrospinning for potential
wound care applications. Pharmaceutics. 2019;11:678.
[334] Ardila DC, Liou JJ, Maestas D, Slepian MJ, Badowski M, Wagner WR, et al. Surface modification of electrospun scaffolds for endothelialization of tissue-
engineered vascular grafts using human cord blood-derived endothelial cells. J Clin Med 2019;8:185.
[335] Zhao YL, Jie X, Shi XY, Ko F. Capturing cancer cells using hyaluronic acid-immobilized electrospun random or aligned PLA nanofibers. Colloids Surf A-
Physicochem Eng Aspects 2019;583:123978.
[336] Grace JM, Gerenser LJ. Plasma treatment of polymers. J Dispersion Sci Technol 2003;24:305–41.
[337] Binkley DM, Lee BEJ, Saem S, Moran-Mirabal J, Grandfield K. Fabrication of polycaprolactone electrospun nanofibers doped with silver nanoparticles formed
by air plasma treatment. Nanotechnology. 2019;30:215101.
[338] Dufay M, Jimenez M, Degoutin S. Effect of cold plasma treatment on electrospun nanofibers properties: a review. ACS Appl Bio Mater 2020.
[339] Fasano V, Laurita R, Moffa M, Gualandi C, Colombo V, Gherardi M, et al. Enhanced electrospinning of active organic fibers by plasma treatment on conjugated
polymer solutions. ACS Appl Mater Interfaces. 2020;12:26320–9.
[340] Hilal N, Khayet M, Wright CJ. Membrane modification: technology and applications. CRC Press; 2012.
[341] Li D, Wang Q, Huang F, Wei Q. Electrospun nanofibers for enzyme immobilization. Electrospinning: Nanofabrication Appl: Elsevier 2019:765–81.
[342] Sharifi F, Atyabi SM, Irani S, Bakhshi H. Bone morphogenic protein-2 immobilization by cold atmospheric plasma to enhance the osteoinductivity of car­
boxymethyl chitosan-based nanofibers. Carbohydr Polym. 2020;231:115681.
[343] Meghdadi M, Pezeshki-Modaress M, Irani S, Atyabi SM, Zandi M. Chondroitin sulfate immobilized PCL nanofibers enhance chondrogenic differentiation of
mesenchymal stem cells. Int J Biol Macromol. 2019;136:616–24.
[344] Park H, Lee KY, Lee SJ, Park KE, Park WH. Plasma-treated poly(lactic-co-glycolic acid) nanofibers for tissue engineering. Macromol Res 2007;15:238–43.
[345] Chen JP, Su CH. Surface modification of electrospun PLLA nanofibers by plasma treatment and cationized gelatin immobilization for cartilage tissue en­
gineering. Acta Biomater. 2011;7:234–43.
[346] Bosworth LA, Hu WX, Shi YN, Cartmell SH. Enhancing biocompatibility without compromising material properties: an optimised NaOH treatment for elec­
trospun polycaprolactone fibres. J Nanomater 2019;2019.
[347] Sun H, Onneby S. Facile polyester surface functionalization via hydrolysis and cell-recognizing peptide attachment. Polym Int 2006;55:1336–40.
[348] Brown JH, Das P, DiVito MD, Ivancic D, Tan LP, Wertheim JA. Nanofibrous PLGA electrospun scaffolds modified with type I collagen influence hepatocyte
function and support viability in vitro. Acta Biomater. 2018;73:217–27.

40
M. Rahmati, et al. Progress in Materials Science xxx (xxxx) xxxx

[349] Hackett JM, Dang TT, Tsai EC, Cao XD. Electrospun biocomposite polycaprolactone/collagen tubes as scaffolds for neural stem cell differentiation. Materials
2010;3:3714–28.
[350] Rianjanu A, Julian T, Hidayat SN, Yulianto N, Majid N, Syamsu I, et al. Quartz crystal microbalance humidity sensors integrated with hydrophilic poly­
ethyleneimine-grafted polyacrylonitrile nanofibers. Sens Actuat B-Chem 2020;319:128286.
[351] Shapourzadeh A, Atyabi SM, Irani S, Bakhshi H. Osteoinductivity of polycaprolactone nanofibers grafted functionalized with carboxymethyl chitosan: Synergic
effect of beta-carotene and electromagnetic field. Int J Biol Macromol. 2020;150:152–60.
[352] Chen Q, Zheng J, Wen L, Yang C, Zhang L. A multi-functional-group modified cellulose for enhanced heavy metal cadmium adsorption: Performance and
quantum chemical mechanism. Chemosphere 2019;224:509–18.
[353] Liu ZM, Xu ZK, Wang JQ, Wu J, Fu JJ. Surface modification of polypropylene microfiltration membranes by graft polymerization of N-vinyl-2-pyrrolidone. Eur
Polym J 2004;40:2077–87.
[354] Shin YM, Lee YB, Kim SJ, Kang JK, Park JC, Jang W, et al. Mussel-inspired immobilization of vascular endothelial growth factor (VEGF) for enhanced
endothelialization of vascular grafts. Biomacromolecules 2012;13:2020–8.
[355] Luong ND, Moon IS, Lee DS, Lee YK, Nam JD. Surface modification of poly(L-lactide) electrospun fibers with nanocrystal hydroxyapatite for engineered scaffold
applications. Mater Sci Eng C-Biomimetic Supramol Syst 2008;28:1242–9.
[356] Dai JM, Yang SL, Jin JH, Li G. Electrospinning of PLA/pearl powder nanofibrous scaffold for bone tissue engineering. RSC Adv 2016;6:106798–805.
[357] Salgado AJ, Coutinho OP, Reis RL. Bone tissue engineering: state of the art and future trends. Macromol Biosci. 2004;4:743–65.
[358] Fang Y, Zhu X, Wang N, Zhang X, Yang D, Nie J, et al. Biodegradable core-shell electrospun nanofibers based on PLA and γ-PGA for wound healing. Eur Polym J
2019;116:30–7.
[359] Duan GA, Greiner A. Air-blowing-assisted coaxial electrospinning toward high productivity of core/sheath and hollow fibers. Macromol Mater Eng
2019;304:1800669.
[360] Huang C, Thomas NL. Fabrication of porous fibers via electrospinning: strategies and applications. Polym Rev 2019;1–53.
[361] Nguyen LH, Gao M, Lin J, Wu W, Wang J, Chew SY. Three-dimensional aligned nanofibers-hydrogel scaffold for controlled non-viral drug/gene delivery to
direct axon regeneration in spinal cord injury treatment. Sci Rep. 2017;7:42212.
[362] Srivastava Y, Rhodes C, Marquez M, Thorsen T. Electrospinning hollow and core/sheath nanofibers using hydrodynamic fluid focusing. Microfluid Nanofluid
2008;5:455–8.
[363] Li F, Song Y, Zhao Y. Core-shell nanofibers: nano channel and capsule by coaxial electrospinning. INTECH Open Access Publisher; 2010.
[364] Zupančič Š. Core-shell nanofibers as drug-delivery systems. Acta Pharm 2019;69:131–53.
[365] Abdullah MF, Nuge T, Andriyana A, Ang BC, Muhamad F. Core-shell fibers: design, roles, and controllable release strategies in tissue engineering and drug
delivery. Polymers (Basel). 2019;11:2008.
[366] Kiatyongchai T, Wongsasulak S, Yoovidhya T. Coaxial electrospinning and release characteristics of cellulose acetate-gelatin blend encapsulating a model drug.
J Appl Polym Sci 2014;131.
[367] Konno M, Kishi Y, Tanaka M, Kawakami H. Core/shell-like structured ultrafine branched nanofibers created by electrospinning. Polym J 2014;46:792–9.
[368] Li D, Xia Y. Electrospinning of nanofibers: reinventing the wheel? Adv Mater 2004;16:1151–70.
[369] Alobaidy AA, Sherhan BY, Barood AD, Alsalhy QF. Effect of bore fluid flow rate on formation and properties of hollow fibers. Appl Water Sci 2017;7:4387–98.
[370] Zhan S, Chen D, Jiao X, Tao C. Long TiO2 hollow fibers with mesoporous walls: sol-gel combined electrospun fabrication and photocatalytic properties. J Phys
Chem B 2006;110:11199–204.
[371] Cao Y, Lu H, Hong Q, Xu B, Wang J, Deng Y, et al. Synthesis of Ag/Co@ CoO NPs anchored within N-doped hierarchical porous hollow carbon nanofibers as a
superior free-standing cathode for LiO2 batteries. Carbon 2019;144:280–8.
[372] Kumar L, Singh S, Horechyy A, Formanek P, Hübner R, Albrecht V, et al. Hollow Au@ TiO 2 porous electrospun nanofibers for catalytic applications. RSC Adv
2020;10:6592–602.
[373] McCann JT, Marquez M, Xia YN. Highly porous fibers by electrospinning into a cryogenic liquid. J Am Chem Soc 2006;128:1436–7.
[374] Malda J, Rouwkema J, Martens DE, Le Comte EP, Kooy FK, Tramper J, et al. Oxygen gradients in tissue-engineered PEGT/PBT cartilaginous constructs:
measurement and modeling. Biotechnol Bioeng 2004;86:9–18.
[375] Katsogiannis KAG, Vladisavljevic GT, Georgiadou S. Porous electrospun polycaprolactone (PCL) fibres by phase separation. Eur Polym J 2015;69:284–95.
[376] Loh QL, Choong C. Three-dimensional scaffolds for tissue engineering applications: role of porosity and pore size. Tissue Eng Part B Rev. 2013;19:485–502.
[377] Lee BS, Yang HS, Jung H, Jeon SY, Jung C, Kim SW, et al. Novel multi-layered 1-D nanostructure exhibiting the theoretical capacity of silicon for a super-
enhanced lithium-ion battery. Nanoscale 2014;6:5989–98.
[378] Birhanu G, Tanha S, Akbari Javar H, Seyedjafari E, Zandi-Karimi A, Kiani Dehkordi B. Dexamethasone loaded multi-layer poly-l-lactic acid/pluronic P123
composite electrospun nanofiber scaffolds for bone tissue engineering and drug delivery. Pharm Dev Technol. 2019;24:338–47.
[379] Yi F, LaVan DA. Poly(glycerol sebacate) nanofiber scaffolds by core/shell electrospinning. Macromol Biosci. 2008;8:803–6.
[380] Huang J, Koutsos V, Radacsi N. Low-cost FDM 3D-printed modular electrospray/electrospinning setup for biomedical applications. 3D Print Med. 2020;6:8.
[381] Soares RMD, Siqueira NM, Prabhakaram MP, Ramakrishna S. Electrospinning and electrospray of bio-based and natural polymers for biomaterials develop­
ment. Mater Sci Eng C Mater Biol Appl. 2018;92:969–82.
[382] Steipel RT, Gallovic MD, Batty CJ, Bachelder EM, Ainslie KM. Electrospray for generation of drug delivery and vaccine particles applied in vitro and in vivo.
Mater Sci Eng C Mater Biol Appl. 2019;105:110070.
[383] Unal S, Arslan S, Gokce T, Atasoy BM, Karademir B, Oktar FN, et al. Design and characterization of polycaprolactone-gelatin-graphene oxide scaffolds for drug
influence on glioblastoma cells. Eur Polym J 2019;115:157–65.
[384] Li JL, Pan K, Tian HF, Yin LJ. The potential of electrospinning/electrospraying technology in the rational design of hydrogel structures. Macromol Mater Eng
2020. 2000285.
[385] Chen WM, Xu Y, Liu YQ, Wang ZX, Li YQ, Jiang GN, et al. Three-dimensional printed electrospun fiber-based scaffold for cartilage regeneration. Mater Des
2019;179:107886.
[386] Toure ABR, Mele E, Christie JK. Multi-layer scaffolds of poly(caprolactone), poly(glycerol sebacate) and bioactive glasses manufactured by combined 3D
printing and electrospinning. Nanomate (Basel). 2020;10:626.
[387] Huang B, Aslan E, Jiang Z, Daskalakis E, Jiao M, Aldalbahi A, et al. Engineered dual-scale Poly (ε-caprolactone) scaffolds using 3D printing and rotational
electrospinning for bone tissue regeneration. Addit Manuf 2020;101452.
[388] Konwarh R, Karak N, Misra M. Electrospun cellulose acetate nanofibers: the present status and gamut of biotechnological applications. Biotechnol Adv.
2013;31:421–37.
[389] Li X, Huang L, Li L, Tang Y, Liu Q, Xie H, et al. Biomimetic dual-oriented/bilayered electrospun scaffold for vascular tissue engineering. J Biomater Sci Polym
Ed. 2020;31:439–55.
[390] Xu T, Ding YC, Liang ZP, Sun HL, Zheng F, Zhu ZT, et al. Three-dimensional monolithic porous structures assembled from fragmented electrospun nanofiber
mats/membranes: Methods, properties, and applications. Prog Mater Sci 2020;112:100656.
[391] Putzu M. Biopolymeric Scaffold obtained by Electrospinning for Intima Vessel regeneration; 2015.
[392] Chen S, John JV, McCarthy A, Xie J. New forms of electrospun nanofiber materials for biomedical applications. J Mater Chem B. 2020;8:3733–46.
[393] Sionkowska A, Planecka A. Preparation and characterization of silk fibroin/chitosan composite sponges for tissue engineering. J Mol Liq 2013;178:5–14.

41

You might also like