Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

Biotechnology Advances 36 (2018) 68–91

Contents lists available at ScienceDirect

Biotechnology Advances
journal homepage: www.elsevier.com/locate/biotechadv

Research review paper

Silk fibroin/hydroxyapatite composites for bone tissue engineering T


a,⁎ b c a
Mehdi Farokhi , Fatemeh Mottaghitalab , Saeed Samani , Mohammad Ali Shokrgozar ,
Subhas C. Kundud, Rui L. Reisd, Yousef Fatahie, David L. Kaplanf
a
National Cell Bank of Iran, Pasteur Institute of Iran, Tehran, Iran
b
Nanotechnology Research Centre, Faculty of Pharmacy, Tehran University of Medical Sciences, Tehran, Iran
c
Department of Tissue Engineering & Applied Cell Sciences, School of Advanced Technologies in Medicine, Tehran University of Medical Sciences, Tehran, Iran
d
3Bs Research Group, Headquarters of the European Institute of Excellence on Tissue Engineering and Regenerative Medicine, University of Minho, AvePark, 4805-017
Barco, Guimaraes, Portugal
e
Department of pharmaceutical nanotechnology, Faculty of Pharmacy, Tehran University of Medical Sciences, Tehran, Iran
f
Department of Biomedical Engineering, Tufts University, 4 Colby St, Medford, MA 02155, USA

A R T I C L E I N F O A B S T R A C T

Keywords: Silk fibroin (SF) is a natural fibrous polymer with strong potential for many biomedical applications. SF has
Silk fibroin attracted interest in the field of bone tissue engineering due to its extraordinary characteristics in terms of
Hydroxyapatite elasticity, flexibility, biocompatibility and biodegradability. However, low osteogenic capacity has limited ap-
Scaffold plications for SF in the orthopedic arena unless suitably functionalized. Hydroxyapatite (HAp) is a well-estab-
Bone tissue engineering
lished bioceramic with biocompatibility and appropriate for constructing orthopedic and dental substitutes.
However, HAp ceramics tend to be brittle which can restrict applications in the repair of load-bearing tissues
such as bones. Therefore, blending SF and HAp combines the useful properties of both materials as bone con-
structs for tissue engineering, the subject of this review.

1. Introduction biocompatibility, biodegradability, tunable structural properties and


ease of fabrication. Degradable polymers are important for these types
Bone has an intrinsic capacity for regeneration; however, strategies of repairs to avoid barriers to full regeneration of bone, and are cate-
to improve the repair capacity of bone are still needed in many cir- gorized to natural polymers such as alginate, chitosan, collagen, silk
cumstances. Bone regeneration depends on the type of defect. For ex- and synthetic polymers including polydioxanone (PDS), poly-
ample, about 10% of fractures are non-union that do not heal sponta- caprolactone (PCL), polyglycolic acid (PGA), polylactic acid (PLA), and
neously and lead to inappropriate bone tissue regeneration and delayed poly (lactic-co-glycolic acid) (PLGA) (Lee and Yuk, 2007; Rezwan et al.,
unions. The probability of normal bone healing in these fractures is 2006). Recently, more attention has been paid to natural polymers in
influenced by age, metabolic condition, and severity of the trauma. comparison to synthetic polymers for bone tissue engineering applica-
Autologous bone or autografts are still considered the clinical “gold tions due to their unique biocompatibility, versatility, and biodegrad-
standard” and the most effective method for bone regeneration. This ability. These types of polymers provide well-patterned structures
approach promotes bone formation by direct bone bonding (osteo- consisting of ligands that can bind to cell surface receptors or provide
conduction) and induces local stem cells to differentiate into bone cells accessible enzymatic degradation sites (Swetha et al., 2010). Silk fi-
(osteoinduction) without any immune responses (Zhang et al., 2014). broin (SF) as a natural protein polymer has potent characteristics for
About 2.2 million bone grafts are performed annually (Giannoudis tissue engineering applications, as a biocompatible, biodegradable, and
et al., 2005) and are met some limitations due to the limited supply, low immunogenic polymer (Farokhi et al., 2014a; Farokhi et al., 2016a;
morbidity of the donor site and are associated with > 50% failure in Mottaghitalab et al., 2013; Omenetto and Kaplan, 2010; Sugihara et al.,
specific regions (Bajaj et al., 2003; Clavero and Lundgren, 2003). 2000). Moreover, SF possesses extraordinary properties for stimulating
Therefore, many attempts have been made to develop suitable bone bone repair; for example, the fibrous structure of SF is mostly similar to
constructs consisting of natural or synthetic biomaterials in order to collagen type I (Col I). The amorphous links between the β-sheets in the
increase bone regeneration capacity and avoid the above limitations. structure of SF act as sites for deposition of HAp nanocrystals because it
Some polymers are usually good candidates for this purpose due to their can mimic the anionic structure of noncollagenous proteins (NCPs)


Corresponding author.
E-mail address: M_farokhi@pasteur.ac.ir (M. Farokhi).

http://dx.doi.org/10.1016/j.biotechadv.2017.10.001
Received 12 June 2017; Received in revised form 12 September 2017; Accepted 4 October 2017
Available online 07 October 2017
0734-9750/ © 2017 Elsevier Inc. All rights reserved.
M. Farokhi et al. Biotechnology Advances 36 (2018) 68–91

Fig. 1. Applications of (SF) as a biomaterial based on Scopus database. (A) SF-related articles (B) SF used for tissue engineering (C) SF used for bone tissue engineering (D) Silk/
hydroxyapatite (SF-HAp) composites. (E) to (H) remarkable studies related to parts (A) to (D).

(Marelli et al., 2012). However, there are also some reports believed metallic implants (Bauer et al., 1991). There are two common struc-
that β-sheets crystal can alone act as a nucleation site for deposition of tures of HAp including monoclinic and hexagonal constructs. In humans
HAp nanaocrystals (Vetsch et al., 2015). Furthermore, this polymer is and animal bones, HAp is mainly in the hexagonal structure and thus
introduced as a potential bone construct due to appropriate mechanical synthetic HAp is the predominant form of CaP used in biomedical fields
behaviors and structural integrity (Mandal et al., 2012; Sofia et al., (Figueiredo et al., 2010). However, HAp is weakly dispersed in aqueous
2001). The ultimate tensile strength of SF extracted from Bombyx mori media which leads to aggregation. Thus, it is necessary to functionalize
(B. mori) is 300–740 MPa (Cunniff et al., 1994; Shao and Vollrath, the surface of HAp (Kim et al., 2014c). In addition, poor fracture
2002) and it also has great breaking strain and high toughness more toughness of HAp (Azadi et al., 2016; Li et al., 2011; Wang et al.,
than synthetic fibers such as Kevlar (Du et al., 2011; Gosline et al., 2004b), migration of nanoparticles from implanted sites (Azadi et al.,
1999). Upon searching the Scopus database for studies related to tissue 2016; Wang et al., 2005) and challenging processability (Azadi et al.,
engineering and bone tissue engineering applications of SF from 2005 2016) restrict the clinical applications of this biomaterial. In order to
to Sep 2017, the number of peer-reviewed original articles per year and overcome these limitations, many efforts have been directed to develop
the first ten scientific studies that attract the most attention and the hybrid structures containing HAp and biopolymers or organic mole-
highest numbers of published articles were assembled (Fig. 1). cules (Li et al., 2008; Li et al., 2011; Wang et al., 2004b; Wang et al.,
Various studies have also recently focused on the applications of 2005). Several natural polymers, including collagen, gelatin, chitosan,
natural polymers as well as SF in bone repair. Although some of these and silk have been used in combination with HAp to enhance bone
polymers have intrinsic capacities for the bone regeneration, this is regeneration (He et al., 2012c). Many structures based on SF/HAp
insufficient to repair large bone defects. For this reason, incorporating composites alone or in combination with bioactive molecules and stem
particles into the polymeric matrix can be advantageous to improve the cells for bone tissue regeneration are schematically represented in
mechanical behavior and tune the topographical features of the scaffold Fig. 2.
in order to mimic the structure of natural bone (Abadi et al., 2010; In this review, we focused on recent studies that used SF/HAp as
Aboudzadeh et al., 2010; Shokrgozar et al., 2010; Swetha et al., 2010). bone constituents. This review starts with a brief description about
HAp with similar properties to bone tissue and suitable biocompat- bioceramics and CaPs, peruses structural properties of HAp and SF, and
ibility has been introduced as an appropriate bioceramic phase which continues with descriptions of the signaling responses. Finally, the
can be incorporated in different polymeric phases (Ginebra et al., combination of SF/HAp for orthopedic applications will be discussed in
2006a; Ginebra et al., 2006b). This is osteoconductive (Cai et al., 2009), detail.
non-toxic (He et al., 2012a), low in immunogenicity with the ability to
directly bind to hard tissues (Azadi et al., 2016; Chen et al., 2015b; Li
2. Bioceramics for bone tissue engineering applications
et al., 2011). HAp is the most stable calcium phosphates (CaPs) under
neutral or basic pHs in the biological fluids and humid conditions (Aoki,
Ceramics are solid materials consisting of non-metallic, inorganic
1991). Thus, this ceramic is usually formed in the fluids containing
substances in two crystalline and non-crystalline (amorphous) forms
phosphates in the bodies of vertebrates that has the pH between 7.2 and
(Hench and Jones, 2005). Ceramic and glasses have different applica-
7.4. The chemical structure of HAp is mostly similar to inorganic
tions in various fields of human health and in industry, such as diag-
compounds existed in the bone matrix. High-affinity of HAp to bone
nostic instruments, thermometers, eyeglasses, tissue culture flasks and
tissue makes it a good bone replacement in comparison to allografts and
fiber optics for endoscopy (Dorozhkin, 2010). During past decades,

69
M. Farokhi et al. Biotechnology Advances 36 (2018) 68–91

Fig. 2. Silk/hydroxyapatite (SF-HAp) composites for bone tissue engineering.

Fig. 3. Different types and applications of bioceramics.

bioceramics have been used for repairing bone defects (Eshtiagh- 2016a, 2016b, 2016c), and integration into living tissue similar to the
Hosseini et al., 2007) and hard tissues (Hench and Jones, 2005). Bio- healthy bone (Dorozhkin, 2010). These inorganic materials include
ceramics are included in the ceramic materials category that usually diverse types from bioresorbable tricalcium phosphate (TCP) to bioac-
have non-crystalline structures (Baino and Vitale-Brovarone, 2014). tive and biostable HAp (Fathi et al., 2008). Most CaP bioceramics are
The main classes of bioceramics and their subset groups and some ap- based on HAp, beta-tricalcium phosphate (β-TCP), alpha-tricalcium
plications are presented in Fig. 3. phosphate (α-TCP) or biphasic calcium phosphate (BCP) containing a
Nontoxic and biologically inactive ceramics such as Al2O3 and ZrO2 mixture of α-TCP + HAp or β-TCP + HAp (Dorozhkin, 2010). The
as bioinert bioceramics (Park, 2009) are usually used for joint pros- major physical and structural properties of CaPs and inorganic com-
theses because of their mechanical properties. Bioactive glasses and ponents of bone and tooth are listed in Table 1.
glass-ceramics directly bond to hard tissues and support new bone
formation (Baino and Vitale-Brovarone, 2014). In the biomaterial in- 2.1. Hydroxyapatite
dustry, CaPs are used predominantly as bone replacements due to their
biocompatibility, low density, chemical stability (Fathi et al., 2008), The term ‘apatite’, given by Wener in 1788, refers to inorganic
chemical and crystallographic similarity to bone mineral (Ding et al., compounds with similar structure but different compositions (Kokubo,

70
M. Farokhi et al. Biotechnology Advances 36 (2018) 68–91

Table 1
Major properties of calcium phosphates and inorganic phases of adult human calcified tissues (Dorozhkin, 2012; Ducheyne et al., 2015; Park, 2009; Ratner et al., 2004; Wnek et al., 2008).

Chemical name Composition Lattice parameters Density Za Lattice structure Transformation Solubility pH
(g/cm3) temperature at 25 °C, stability
Formula Ca/P (°C)b g/L range in
molar (− logKa) aqueous
ratio solutions
at 25 °C

Monocalcium Ca(H2PO4)2·H2Oc 0.5 a = 5.6261(5), 2.23 2 Triclinic ~ 100 ~ 18 0.0–2.0


phosphate b = 11.889(2), transforms into (1.14)
monohydrate c = 6.4731(8) Å, MCPA
(MCPM) α = 98.633(6)°,
β = 118.262(6)°,
γ = 83.344(6)°
Monocalcium Ca(H2PO4)2 0.5 a = 7.5577(5), 2.58 2 Triclinic ~ 100 ~ 17 2
phosphate b = 8.2531(6), (1.14)
anhydrous c = 5.5504(3) Å,
(MCPA or α = 109.87(1)°,
MCP) β = 93.68(1)°,
γ = 109.15(1)°
Dicalcium CaHPO4.2H2O 1.0 a = 5.812(2), 2.32 4 Monoclinic ~ 80 ~ 0.088 2.0–6.0
phosphate b = 15.180(3), (6.59)
dihydrate c = 6.239(2) Å,
(DCPD), β = 116.42(3)°
mineral
brushite
Dicalcium CaHPO4 1.0 a = 6.910(1), 2.89 4 Triclinic ~ 100 ~ 0.048 1
phosphate b = 6.627(2), Sinter ~ 300 (6.90)
anhydrous c = 6.998(2) Å,
(DCPA or α = 96.34(2)°,
DCP), mineral β = 103.82(2)°,
monetite γ = 88.33(2)°
Octacalcium Ca8(HPO4)2(PO4)4.5H2O 1.33 a = 19.692(4), 2.61 1 Triclinic An unstable ~ 0.0081 5.5–7.0
phosphate b = 9.523(2), transient (96.6)
(OCP) c = 6.835(2) Å, intermediate
α = 90.15(2)°,
β = 92.54(2)°,
γ = 108.65(1)°
α-Tricalcium α-Ca3(PO4)2 1.5 a = 12.887(2), 2.86 24 Monoclinic – ~ 0.0025 3
phosphate (α- b = 27.280(4), (25.5)
TCP) c = 15.219(2) Å,
β = 126.20(1)°

β-Tricalcium β-Ca3(PO4)2 1.5 a = b = 10.4183(5), 3.08 21d Rhombohedral ~ 1125 ~ 0.0005 3


phosphate (β- c = 37.3464(23) Å, transforms into (28.9)
TCP) γ = 120° α-TCPe

Amorphous CaxHy(PO4)z.nH2O, n = 3–4.5, 1.2–2.2 – – – – – 5 ~5–12f


calcium 15–20% H2O
phosphates
(ACP)
Calcium-deficient Ca10 − x(HPO4)x(PO4)6 − x(OH)2 − x 1.5–1.67 – – – – ~ 700 ~ 0.0094 6.5–9.5
hydroxyapatite (0 < x < 1) transforms into (385)
(CDHA or Ca- β-TCP
def HA)g
Hydroxyapatite Ca10(PO4)6(OH)2 1.67 a = 9.84214(8), 3.16 4 Monoclinic ~ 250, ~ 0.0003 9.5–12
(HA, HAp or b = 2a, c = 6.8814(7) monoclinic to (116.8)
OHAp) Å, γ = 120° 2 hexagonal Sinter
(monoclinic) ~ 1000
a = b = 9.4302(5),
c = 6.8911(2) Å,
γ = 120° (hexagonal) Hexagonal

Fluorapatite (FA or Ca10(PO4)6F2 1.67 a = b = 9.367, 3.20 2 Hexagonal – ~ 0.0002 7–12


FAp) c = 6.884 Å, (120)
γ = 120°

(continued on next page)

71
M. Farokhi et al. Biotechnology Advances 36 (2018) 68–91

Table 1 (continued)

Chemical name Composition Lattice parameters Density Za Lattice structure Transformation Solubility pH
(g/cm3) temperature at 25 °C, stability
Formula Ca/P (°C)b g/L range in
molar (− logKa) aqueous
ratio solutions
at 25 °C

Oxyapatite (OA, Ca10(PO4)6O 1.67 a = b = 9.432, ~ 3.2 1 Hexagonal – ~ 0.087 3


OAp or OXA)h c = 6.881 Å, (369)
α = 90.3°,
β = 90.0°,γ = 119.9°

Tetracalcium Ca4(PO4)2O 2.0 a = 7.023(1), 3.05 4 Monoclinic – ~ 0.0007 3


phosphate b = 11.986(4), (38–44)
(TTCP or c = 9.473(2) Å,
TetCP), β = 90.90(1)°
mineral
hilgenstockite
Enamel – 1.63 a = 9.441, – – – – –
c = 6.880 Å
Dentine – 1.61 a = 9.421, – – – – –
c = 6.887 Å
Bone – 1.71 a = 9.41, c = 6.89 Å – – – – –

a
Number of formula units per unit cell.
b
Stable at temperatures above 100 °C.
c
Cannot be precipitated from aqueous solutions.
d
Per the hexagonal unit cell.
e
Cannot be measured precisely. However, the following values were found: 25.7 ± 0.1 (pH 7.40), 29.9 ± 0.1 (pH 6.00), 32.7 ± 0.1 (pH 5.28). The comparative extent of dis-
solution in acidic buffer is: ACP > > α-TCP > > β-TCP > CDHA > > HA > FA.
f
Always metastable.
g
Occasionally referred to as “precipitated HA” (PHA).
h
The existence of OA remains questionable.

2008). These minerals have lattice parameters a = 0.9432 nm and minerals in bone and teeth. For many years, it was assumed that the
c = 0.6881 nm and crystallize into hexagonal rhombic prisms (Park, inorganic parts of calcified tissues were HAp. However, studies focused
2009; Park and Lakes, 2007). Synthetic or pure HAp is introduced by on synthetic carbonated and biological apatites showed that the bone
Ca10(PO4)6(OH)2 or Ca10 − z(HPO4)z(PO4)6 − z(OH)2 − z [0 < z < 1] as mineral did not consist of pure apatite but contained carbonated apatite
a general formula. The chemical behavior of HAp depends on stoi- (Kokubo, 2008) and some ions such as K+, Mg2 +, Na+, F− and CO32–
chiometry status, which is basic for stoichiometric apatite (z = 1) and (Li et al., 2008; Webster et al., 2004). Moreover, bioactivity of bone
acidic for the others (Faria et al., 2008). mineral is higher than synthetic HAp (Fathi et al., 2008). Therefore,
Inorganic materials coexist with organic molecules in a well-ordered non-stoichiometric HAp substituted with suitable ions during synthesis
manner in some tissues such as bone, cartilage and skin (Wang et al., would be useful to mimic the function of bony apatite and os-
2004a). Natural bone is a nanocomposite (Andiappan et al., 2013; teoinductive activity when compared to pure HAp (Ren et al., 2009).
Bhattacharjee et al., 2016b) composed of organic components (~ 90% Different methodologies have been used to prepare HAp-based
type I collagen, ~5% NCPs,~2% lipids by weight), water and inorganic products in micro- or nanoscale quantities. Co-precipitation (Cao et al.,
CaP (Boskey, 2013; Šupová, 2015; Wu et al., 2014; Zhao et al., 2009). 2013; Ming et al., 2014a), chemical precipitation (He et al., 2012a;
Recently, NCPs have attracted attention due to their effect on bone Kweon et al., 2011; Wang et al., 2002), solid-state reaction (Ming et al.,
tissue behavior, including cell proliferation and attachment, HAp and 2014a; Wang et al., 2002), hydrothermal methods (Cao et al., 2013;
Ca2 + binding and degree of mineralization (Young et al., 1992). Os- Kweon et al., 2011; Ming et al., 2014a; Wang et al., 2002), sol-gel
teocalcin, sialoprotein, alkaline phosphatase (ALP) and matrix Gla synthesis (Cao et al., 2013; Kweon et al., 2011; Ming et al., 2014a), and
protein are the most abundant NCPs that are a main focus of research; wet mechanochemical routes (Cao et al., 2013; Nemoto et al., 2001;
however, it is not well-understood that which of these proteins may Wang et al., 2002) are common methods used for HAp synthesis. Mi-
have a crucial role in bone behavior (Fig. 4). It is reported that the croemulsions, spray drying and plasma melting are also applied for
combination of these proteins may influence bone functions synergis- synthesizing spherical HAp powders (Liu et al., 2013a). Structural
tically more than each of them alone. It should be noted that some features of the synthesized powders are mainly related to the corre-
features such as gender, age, health status and ethnicity affect the sponding preparation method (Kweon et al., 2011). Control of size
proportion of these proteins in bone structure of an individual patient. distribution and morphology of nano-sized HAp powders (nHAps) is
Therefore, the quantity and quality of NCPs can influence the char- critical for tissue engineering and drug delivery systems, which are
acteristics of bone tissue in terms of structure and function (Boskey, often ignored in some methods (He et al., 2012a; Huang et al., 2014).
1989, 2013). The different methods suitable for synthesis of nHAps are summarized
The inorganic phase of bone consists of HAp nanocrystals (about in Table 2.
70 wt% (Andiappan et al., 2013; Wang et al., 2004a)) dispersed in
collagen type I (Col I) (Andiappan et al., 2013; Wang et al., 2004a). This
mineral phase is crystallized along the long axis of collagen fibrils (Hu 2.2. Signaling responses of osteoblast cells to hydroxyapatite
et al., 2015) and integrates together in a tight hierarchical organization
(Fig. 4) [28]. The X-ray diffraction (XRD) spectrum of sintered bone and Dynamic interactions between osteoblasts and their extracellular
mineral apatite showed that HAp and fluorapatite are the main matrix (ECM) regulate their behavior and differentiation potential.
Understanding the principal ECM cues promoting osteogenic

72
M. Farokhi et al. Biotechnology Advances 36 (2018) 68–91

Fig. 4. Ultrastructure of bone tissue. The extracellular matrix of bone is highly specialized. The organic part of bone is composed of collagen, noncollagenous proteins (NCPs) and lipids.
The inorganic phase of bone mainly consists of hydroxyapatite (HAp) nanocrystals. This mineral phase is crystallized along the long axis of collagen fibrils and integrates together in a
tight hierarchical organization. These crystals associate with the collagen fibers, making bone hard and strong. This matrix is organized into numerous thin layers, known as lamellae.

differentiation can be pivotal for both bone tissue engineering and re- interaction between HAp and osteoblast integrins such as BMP/Smad
generative medicine. HAp can trigger signaling cascades; however, the (Liu et al., 2013b), Wnt (Thorfve et al., 2014), TGF-β, MAPK, and Notch
mechanism of action is not well understood. The bioactivity of any signaling pathways (Lü et al., 2014).
material, here defined as the ability of osteoblasts to adhere and spread
upon it, depends on these signaling mechanisms (Kokubo, 1991;
3. Structure and characteristic of silk fibroin
Zambuzzi et al., 2011a). The signaling pathway involved in osteoblasts
adhering on HAp scaffolds was recently evaluated (Gemini-Piperni
Silk protein is found in the gland of silk producing arthropods such
et al., 2014). Adhesion of osteoblast cells on HAp substrates actives
as spiders, silkworms, mites, scorpions, and bees. During metamor-
such signaling pathway that leads to activation of protein kinase C
phosis, silk is spun to fibers. Cocoon silk, as a Food and Drug
(PKC), protein kinase A (PKA), Adducin 1 (ADD1) and vascular en-
Administration (FDA) approved biomaterial for some medical applica-
dothelial growth factor (VEGF).
tions, is produced by silkworm B. mori (Kaplan et al., 1994). B. mori silk
Moreover, kinase activation leads to increased phosphorylation of
is composed of two major types of proteins, silk fibroin (SF) and sericin.
Ser-421 in histone deacetylase 1 (HDAC1), a cyclin-dependent kinase 5
SF is located in the core of B. mori silk fibers and consists of light
(CDK5) substrate. The surface chemistry of HAp is crucial for promoting
(~ 26 kDa) and heavy (~ 390 kDa) chains; while, sericin coats the SF
the differentiation of osteoblasts via phosphorylation (Gemini-Piperni
chains. Silk fibers contain about 70–75% SF filaments and 25–30%
et al., 2014). In response to HAp, other signaling molecules such as
sericin. The primary sequence of sericin consists of some repeat motifs
extracellular regulated kinases (ERK) and SOX9 are also activated (Song
(Sutherland et al., 2010). The structure of the SF heavy chain is similar
et al., 2008b). The integrin receptors mediate the interaction between
to amphiphilic block copolymers because it contains several hydro-
osteoblast cells cultured on HAp substrates with some domains of ECM
phobic and hydrophilic domains or blocks. The hydrophobic blocks of
proteins that can activate the cytoskeletal and intracellular signaling
SF are comprised of highly conserved (GAGAGS) and less conserved
cascades like focal adhesion kinase (FAK). FAK is associated with Shc
(GAGAGX) repeats (where X is either valine or tyrosine) that are re-
protein activating Ras that stimulates the ERK signaling cascade
sponsible for the crystalline structure and β-sheet conformation of SF.
(Miyamoto et al., 1996). Activated ERK affects the expression level of
The hydrophilic part of SF core consists of short and non-repetitive
different osteoblast genes such as osteocalcin and Col I (Fig. 5) (Mimori
segments in comparison with hydrophobic parts (Bini et al., 2004;
et al., 2007; Xiao et al., 2002). Therefore, ERK signaling is essential for
Kaplan and McGrath, 2012). Silk fibers have useful mechanical prop-
cytoskeletal rearrangement and intracellular signaling (Zambuzzi et al.,
erties, including a tensile strength of 0.5 GPa, breaking elongation of
2011b). Various signaling cascades are also triggered upon the
15% and 62,104 J kg− 1 toughness (Li et al., 2012). The Young's

73
Table 2
Different methods for nHAps fabrication and main properties based on the materials used (El Briak-BenAbdeslam et al., 2008; Gedanken, 2004; Sadat-Shojai et al., 2013; Sasikumar and Vijayaraghavan, 2008, 2010; Tas, 2000).

Synthesis method Raw materials Advantages Disadvantages Synthetic conditions Shape Size-Size distribution Product features
M. Farokhi et al.

categories1

Solid-state reaction Few Ca and PO4 containing Relatively simple Weak precise control over 900–1300 °C, usually with water 1, 3 Usually micron-Wide Diverse morphology
chemicals Mass production product features vapor flowing Very high crystallinity
Presynthesized CaP salts Using organic templates to Heterogeneous composition Usually low purity
control morphology of powder Variable Ca/P
Low cost Irregular shape
Cannot be exploited for
biomimetic synthesis
Mechanochemical Few Ca and PO4 containing Simplicity – Wet or dry medium 1, 2, 3 Nano-Usually wide Diverse morphology
(Mechanosynthesis) chemicals Reproducibility Reaction activated Very high crystallinity
e.g.: CaHPO4.2H2O + CaO Production of by mechanical milling Low purity
CaCO3 + NH4H2PO4 nanocrystalline alloys and Effective factors: reagents, Non-stoichiometric Ca/P
Ca3(PO4)2 + Ca(OH)2 ceramics milling medium, atmosphere,
Ca(OH)2 + P2O5 Well-defined structure of rotational speed, diameter of
CaO + Ca(OH)2 + P2O5 powder milling ball
Mass production
Low cost
Chemical precipitation Frequently few Ca and PO4 The simplest route Precise control over Usually at room temperature 1, 2, 3, 4, 5, 7, Usually nano- Variable Diverse morphology
containing reagents Using organic templates conditions necessary to Atmospheric pressure 8, 11 Frequently low
and molecules to control minimize defects Aqueous medium crystallinity
morphology Phase impurities (DCPAa, pH 7–12 Variable purity
Low cost OCPb) Usually non-
stoichiometric Ca/P
Hydrolysis Few chemical substances Distinct method to convert Usually high cost pH 6–7 for acidic phases (DCPD, 1, 3, 5, 7 Variable- Variable Diverse morphology
Other CaP phase (DCPDc, DCPA a CaP into HAp DCPA) Variable crystallinity

74
TCP) Method to modify a Conversion depends on pH, T and Usually high purity
prepared HAp other ions Stoichiometric Ca/P
Sol-gel Variable chemical reagents Molecular level mixing of High cost of some raw Aqueous or organic media 1, 2, 3 Nano-Narrow Diverse morphology
Ca(OEt)2, Ca(NO3)2, P(OEt)3, P2O5, reactants materials Aging at room temperature Variable crystallinity
(NH4)2HPO4 Improving chemical Possible impurities (CaO, Drying at ~ 100–150 °C Variable purity
homogeneity Ca2P2O7, Ca3(PO4)2, CaCO3) Heat treatment at ~ 300–900 °C Stoichiometric Ca/P
Low temperature for Long aging to complete Higher bioresorbability
various steps reaction than chemical
Stoichiometric structure Thermal treatment to get precipitation
pure HAp
Hydrothermal Variable reagents Using organic modifiers to Expensive equipment for Aqueous solution 1, 2, 3, 4, 5, 6, Nano and micron- Frequently needle-like
Wet chemically prepared HAp, control process high temperature and High temperature and pressure 8, 9 Wide Usually irregular
other calcium phosphates, seeding Regulating more pressure 100–200 °C (1–2 MPa), Very high crystallinity
large crystals predictable reactions as Poor control over 300–600 °C (1–2 kbar) Usually high purity
crystal nucleation, morphology and size Stoichiometric Ca/P
growth, and aging distribution
Emulsion Many raw materials Simplicity High cost Low temperature 1, 2, 3, 5 Nano-Narrow Frequently needle-like
More effective control over Different surfactants Frequently low
morphology and Mild conditions crystallinity
microstructure Variable purity
Agglomerate-free Non-stoichiometric Ca/P
Sonochemical Few Ca ad PO4 containing Increasing rate of crystal – Aqueous environment 1, 2, 3 Nano-Usually narrow Diverse morphology
chemicals growth Chemical reactions activated by (usually needle-like)
e.g.: Ca(OH)2 + Ca(H2PO4)2 More uniform, smaller and powerful ultrasound radiation Variable crystallinity
purer crystals Heterogeneous Usually high purity
Minimal agglomeration reactions between liquid and Variable Ca/P
Using organic modifiers to solid reactants
control or change process
Usually low cost
(continued on next page)
Biotechnology Advances 36 (2018) 68–91
M. Farokhi et al. Biotechnology Advances 36 (2018) 68–91

1-Irregular sphere, 2-(micro and nano)Sphere or ball-shaped, 3-rod- or needle-shaped or strip-like, 4-plate or sheet, 5-self-assembled or bundled nanorod, 6-dendelion or rosette, 7-bundled sheets or leaves, 8-Flower, 9-porous microsphere or
modulus of silk obtained from forced silking of silkworm in the la-

Usually stoichiometric Ca/


boratory is about 12.4–17.9 GPa, with an ultimate tensile strength of

Variable crystallinity
Diverse morphology

Diverse morphology
360–530 MPa and 18–21% breaking elongation (Pérez-Rigueiro et al.,

Usually high purity


(usually irregular)

High crystallinity
2001). Silk fibers have been used for many years as sutures and for
Product features

Variable purity
Variable Ca/P
textile engineering due to their high toughness (Nova et al., 2010).
Some studies have reported comparable mechanical behavior for silk
filaments with nylons (Lucas, 1964). These unique mechanical char-

P
acteristics make SF a useful polymer candidate for load-bearing appli-
cations, such as for bone tissue engineering in composite systems with
Size-Size distribution

embedded in micron
aggregates- Variable
HAp.
Usually nano-Wide

Nano particles

3.1. Osteogenic signaling responses to silk fibroin

While silk scaffolds do not inherently contain any cell-specific sig-


naling epitopes in the primary sequence, silk-based scaffolds provide
suitable substrates for supporting the proliferation and adhesion of
mesenchymal stem cells (MSCs) (Mandal and Kundu, 2009; Mauney
categories1

1, 2, 3, 5

1, 2, 3, 5

et al., 2007). The ability of hydrolyzed silk protein to improve the ALP
Shape

activity of osteoblast cells and their osteogenic differentiation has also


been reported (Seo et al., 2011). Moreover, B. mori silk induces mineral
deposition from rat bone marrow stromal cells (BMSCs) during 12 days
Particles formed from aqueous

Heat sources: flame or furnace

of culture in vitro (Nikbakht Dastjerdi, 2006). This study demonstrated


combustion of a fuel and
solution via spontaneous

that by using silk, no dexamethasone additives are required for pre-


osteogenic factor, although the compound can accelerate the outcomes.
Synthetic conditions

oxidiser at elevated

Dexamethasone triggers Wnt/β-catenin signaling dependent Runx2


expression (Langenbach and Handschel, 2013). Furthermore, other
temperatures

signaling factors like Wnt, Notch, fibroblast growth factor (FGF), bone
morphogenetic protein/transforming growth factor beta (BMP/TGFβ),
insulin-like growth factor (IGF), and platelet derived growth factor
(PDGF) regulate the osteogenic differentiation of stem cells on silk-
TCP at high temperature of
HAp decomposition into α-

based substrates (Midha et al., 2016). Based on pharmacologic and


Secondary aggregations

molecular studies, Notch signaling was responsible for differentiation of


leading to decrease in
temperature reaction
Mixed phases due to

Particle morphology

specific surface area


uncontrollable high

mesenchymal progenitor cells into osteoblast cells. Silk protein pro-


dependent on fuel

voked the upregulation of some osteoblast differentiation markers (e.g.


Disadvantages

ALP, osteorix and Runx2) by inhibiting Notch signaling (Jung et al.,


2013). Additionally, the mass of trabecular bone in the skeletogenic
flame
front

used

mesenchyme of mice limbs was significantly increased by inhibiting the


Notch signaling pathway in adolescent mice. Severe osteopenia was
Well chemical homogeneity

observed in mice during aging period when they were without me-
Inexpensive raw materials

senchymal progenitors in their bone marrow. The inhibition of osteo-


Fairly low temperature

continuous production
Single step operation
Quick production of

genic differentiation by Notch signaling is triggered by Hes and Hey


Easily scaled-up for
Relatively simple

Usually low cost

Usually low cost

proteins, which can suppress Runx2 activity through physical interac-


tion (Fig. 6a) (Hilton et al., 2008). Thus, for optimal osteogenic in-
Advantages

of powder

duction of silk scaffolds, suppression of the Notch signaling pathway is


powder

advantageous (Fig. 6b).

3.2. Silk fibroin-based scaffolds for bone tissue engineering


sucrose, citric acid, succinic acid)
Oxidants (Ca(NO3)2, (NH4)2HPO4

The outstanding properties of SF such as biocompatibility, biode-


Organic fuels (glycine, urea,

e.g.: Ca(NO3)2, (NH4)2HPO4


Variable chemical materials

gradability, low immunogenicity, and suitable processability make it a


useful scaffold system for various tissue engineering applications, in-
Few raw materials

cluding bone constructs. Another advantage of SF as a bone substitute is


Raw materials

its high mechanical strength (Kundu et al., 2013; Mottaghitalab et al.,


mesoporous sphere, 10-Bowknot, 11-Dumbbell.
and HNO3)

2015b; Nourmohammadi et al., 2017); it can bear the force produced in


vivo (Nazarov et al., 2004). These properties of SF have attracted many
researchers to use this biomaterial in bone tissue engineering applica-
Anhydrous dicalcium phosphate.

Dicalcium phosphate dihydrate.

tions. Recent investigations on the SF-based scaffold as a bone construct


are summarized in Table 3.
Self-propagating combustion

Octacalcium phosphate.

4. Silk fibroin/hydroxyapatite for bone tissue engineering


Synthesis method
Table 2 (continued)

4.1. Silk fibroin/hydroxyapatite composites for bone tissue engineering


synthesis

Pyrolysis

Bone tissue engineering offers suitable constructs for restoring and


repairing bone defects (Ma et al., 2005). Mimicking the structure and
1

b
a

composition of natural tissues by engineering biomaterials, such as

75
M. Farokhi et al. Biotechnology Advances 36 (2018) 68–91

Fig. 5. Signaling pathway of osteoblast cells seeded on hydroxyapatite (HAp) scaffold. Adhesion of osteoblast cells on HAp substrates triggers some signaling pathways like extracellular
regulated kinases (ERK). Due to interaction of integrin receptors with such extracellular matrix of bone e.g. fibronectin, focal adhesion kinase (FAK) as an intracellular signaling cascade
can be activated. This causes activation of ERK pathway and then expression of some osteoblast specific genes e.g. osteocalcin and collagen type I.

developing inorganic-organic composites, has grown in interest and calvarial defects in rat model that was stably biodegraded without
focus over the past decades. Some typical ceramics such as bioactive provoking the immunogenic responses (Jo et al., 2017).
glass, HAp, β-TCP, and calcium citrate are suitable candidates for bone The homogenous dispersion and interfacial bonding between poly-
tissue engineering applications (Chouzouri and Xanthos, 2007; Maquet meric matrix and HAp is still challenging in fabricating composites. The
et al., 2004). The most common organic material used as bone con- interfacial bonding between HAp and the organic phase is dependent to
structs is collagen. This protein polymer can be obtained from some the density of contact numbers and the chemical interactions. In this
animals (e.g. skin of pig and cow) or generated through transgenic context, pretreatment of SF fibrils with an alkali increased the exposed
techniques. However, collagen is not cost-effective, provoke immune active sites on the surface that further improved the interactions be-
responses, and retains a risk for contamination (Du et al., 2009). For tween the organic and mineral phases. The alkali pretreatment effected
these limitations, some researchers have tried to use other natural the crystallographic characteristics of HAp in comparison to pretreat-
biomaterials such as silk for bone tissue engineering. ment with proteinase K, a protease that digests the SF. Applying both
Blending silk with HAp enhanced crystal formation of HAp along pretreatments enhanced the exposed active sites to interact with the
the c-axis and the coordinative effect on the structure and properties mineral phases. It was suggested that SF surface fibrils were mostly
between SF and HAp was found during the composite film fabrication. disentangled by alkali, while the enzymatic pretreatment was effective
The nucleation of HAp could enhance the molecular orientation and in disentanglement of the SF blocks into smaller fragments (Wang et al.,
crystallinity of SF (Du et al., 2009). Recently, Jo et al. evaluated algi- 2007). An increase in number of contact sites enhanced the interactions
nate/HAp/SF composites as a bone replacement (Jo et al., 2017). Four between organic and inorganic phases to improve microhardness
weeks after implantation, the rate of bone formation in the defected site (> 50%) (Wang et al., 2007; Wang et al., 2004b; Wang et al., 2005).
was significantly higher using alginate/HAp/SF scaffold than alginate However, the remaining sodium ions from the alkali reagent may have
and control (unfilled defects) groups. The suitable biocompatibility of unexpected side effects on crystallographic properties of HAp due to
alginate/HAp/SF scaffold led to less immunogenic response or forma- replacement of calcium with sodium ions.
tion of giant cells around the degraded grafts. Significantly lower ex- During the past decade, many efforts have been attempted to de-
pression level of tumor necrosis factor-alpha (TNF-α) was also observed velop hybrid biomaterials based on organic and inorganic materials to
in alginate/HAp/SF group compared with alginate and alginate/HAp mimic the structure and function of natural biomaterials. The existence
groups; while, the osteogenic markers such as Runx2, osteoprotegerin, of organic materials like polymers and proteins in the structure of hy-
and FGF-23 showed higher expression rate. The alginate/HAp/SF brid materials regulate the nucleation of inorganic crystals, and en-
scaffold was introduced as a potential bone construct for repairing hance the physicochemical properties of the microstructure. As

76
M. Farokhi et al. Biotechnology Advances 36 (2018) 68–91

Fig. 6. The role of Notch signaling pathway in osteoblast differentiation of stem cells. a). Notch signaling blocks RUNX2 expression and consequently inhibition of osteocalcin and
osteopontin. b). Blocking of Notch signaling cascade after seeding of stem cells on silk fibroin based scaffold.
EGF: epidermal growth factor, LNR: Lin-12/Notch repeats, HD domain: heterodimerization domain, TMD: transmembrane domain, RAM: RBPjκ association module, ANK: ankyrin, NLS:
nuclear localization sequences, TAD: transactivation domain, PEST: proline (P), glutamic acid (E), serine (S), threonine (T) rich motif, NICD: Notch intracellular domain, ADAM: a
disintegrin and metalloproteinase, MAMAL: Mastermind-like, HES: hairy and enhancer of split, HEY: HES-related with YRPW motif.

mentioned above, adding HAp to SF improve the biological and me- similar crystallographic and chemical characteristics to the natural
chanical properties of SF. HAp/SF composites are so similar to the apatite in bone (Murugan and Ramakrishna, 2005). Composites con-
natural bone structure that mimic the inorganic and organic phases of taining nHAp induced both osteogenesis and angiogenesis by stimu-
native bone tissue. Thus, combination of inorganic and organic mate- lating the proliferation, adhesion, and differentiation of osteoblasts;
rials in a unique hybrid composite can improve the flexibility, me- therefore, these composites are potent for bone tissue engineering ap-
chanical strength, and the toughness of the bone tissue scaffolds. Some plications (Kilian et al., 2008; Nageeb et al., 2012; Wei and Ma, 2004).
applications of SF/HAp composites for bone tissue engineering are Besides, in many studies SF protein extracted from B.mori alone or
listed in Table 4. in blended with CaP has been used for bone tissue engineering.
However, there is no RGD motifs or sufficient arginine amino acids in
4.2. Silk fibroin/nano-hydroxyapatite composites for bone tissue the structure of SF derived B.mori to support cellular adhesion (Sen and
engineering Babu, 2004). Some reports suggesting the presence of RGD epitopes and
high arginine content in the structure of non-mulberry silk fibroins
Recently, nanoparticles have attracted attention in biomedical fields (NSF) that reduced the cytotoxicity and inflammatory responses (Datta
due to their exceptional properties such as high surface to volume ratio, et al., 2001; Mandal and Kundu, 2008). Moreover, presence of hydro-
tunable structural properties in terms of surface chemistry, size and philic amino acids in the structure of NSF promotes cellular attachment
shape (Brannon-Peppas and Blanchette, 2012). Today, different appli- to the surface of the materials (Meinel et al., 2005b; Sen and Babu,
cations are considered for nanoparticles including theranostics therapy, 2004; Zhang et al., 2005b). The in situ reinforcement of NSF-nHAp
molecular imaging, drug delivery and cell labeling (Hassani Besheli scaffold and the NSF scaffold with external nHAp deposition showed
et al., 2017; Kim et al., 2014a; Mottaghitalab et al., 2015a; higher mechanical properties and biocompatibility in comparison to
Mottaghitalab et al., 2017). It is also observed that the alterations in NSF-nHAp reinforced fibroin scaffolds. Both composite scaffolds also
nanoscale features can affect cellular responses. For example, nano- showed less immunogenicity by using osteoblast-macrophage co-cul-
patterned structures were potent in directing stem cell differentiation ture model (Behera et al., 2017). However, no significant anti-in-
without the using of exogenous biochemical agents (Kim et al., 2012a, flammatory differences were observed through the modified scaffold
2012b). Controlling the cellular behavior by using nano-patterned indicating the native anti-inflammatory properties of fibroin structures.
structures is a key application for nanotechnology in biomedical fields. It was also observed that by culturing the macrophages on the modified
In the orthopedic field, HAp nanocrystals were sintered and improved scaffolds, the level of interleukin 1 beta (IL1-β) increased that might be
densification because of the high surface area. The high surface area related to the role of HAp (Loi et al., 2016). Similarly, nHAp/SF pow-
can also increase fracture toughness and mechanical strength of HAp ders were synthesized using a co-precipitation method, added to SF
nanocrystals. However, agglomeration is the most challenging issue in solution in order to fabricate nHAp/SF composite material (Liu et al.,
using nano-powders (LeGeros, 1993). nHAp showed better bioactivity 2011). These composites showed improved compressive properties. The
in comparison to coarser crystals that broaden its applications in tissue- good compressive properties were due to the uniform dispersion of
engineered implants when compared to other constructs (Stupp and nHAp/SF in SF solution, the high surface area of needle-like nHAp/SF
Ciegler, 1992). Moreover, nHAp have high bioresorption properties and powders, and good molecular interaction between SF molecules in the

77
M. Farokhi et al.

Table 3
SF based scaffolds for bone tissue engineering applications.

Material Processing method Scaffold structure Cell Key findings Ref.

TSFa/PLAb Electrospinning Nanofibers mMSCsc Inducing the proliferation and attachment of MSC, increasing the rate of ALPd production, (Shao et al., 2016a)
(90:10) osteogenesis, and bone mineralization
NSFe/PCLf Electrospinning Nanofibers – Stimulating the early bone formation, infiltration of implants and formation of strong bonds at the (Bhattacharjee et al., 2016a)
interface of bone-implant, activating the initial immune reaction which reaching to normal after
4 weeks post-implantation
SFg/CMCh Electrospinning Nanofibers hMSCsi Increased proliferation, adhesion and differentiation of hMSCs confirmed by ALP production, (Singh et al., 2016)
transcription of RUNX2, and expressions of osteocalcin and type1 collagen
SF/Decellularized pulp/Collagen/ Freeze-drying 3D scaffold MG-63 Capability of the modified scaffold for alveolar bone repair due to outstanding biofunctionality (Sangkert et al., 2016)
Fibronectin
SF/Diopside Freeze-drying 3D scaffold MC3T3-E1 Increased wettability, suitable porosity, high mechanical strength, and good biocompatibility by (Ghorbanian et al., 2013)
incorporating diopside nanopowders into the SF matrix
j k l
SF/CS /Nano ZrO2 Freeze-drying 3D scaffold HGF Formation of a porous scaffold with appropriate pore size for cell infiltration, improved compressive (Teimouri et al., 2015)
strength and water uptake, and decreased porosity after adding zirconia
SF/CaPm/PLGAn Electrospinning/Freeze- 3D scaffold hOBso Providing an optimal microenvironment for the biological behavior of osteoblast cells due to (Farokhi et al., 2014b)
drying acceptable porosity, chemical and physical properties with ability to sustain the release rate of VEGFp
within 28 days, stimulating the new bone formation at the site of injury 10 weeks after implantation
SF Casting Film BMSCq More efficient delivery of BMP-2r after immobilizing this biomolecule on the surface of biomaterial in (Karageorgiou et al., 2004)
comparison to soluble delivery, retained in vitro biological activity of BMP-2 based on its potential to
induce the osteogenic markers of BMSCs
SF – Mesh OECs/hOBs Extracellular matrix deposition by hOBs during 4 weeks, enhancing expression of osteogenic markers, (Fuchs et al., 2009)
retained functionality of both cell types in provoking the formation of capillary like structures and

78
differentiation of hOBs
SF Salt leaching 3D scaffold hMSC Increased bone formation in calvarial critical size defects in mice during 5 weeks by using tissue- (Meinel et al., 2005a)
engineered bone implants, detecting less bone formation by using stem cell loaded SF scaffolds and
scaffolds alone

a
Tussah silk fibroin.
b
Polylactic acid.
c
Mouse mesenchymal stem cells.
d
Alkaline phosphatase.
e
Non-mulberry silk fibroin.
f
Silk fibroin.
g
Poly(ε-caprolactone).
h
Carboxymethyl cellulose.
i
Human mesenchymal stem cells.
j
Chitosan.
k
Zirconia.
l
Human gingival fibroblast.
m
Calcium phosphate.
n
Poly (lactic-co-glycolic acid).
o
Human osteoblast cell.
p
Vascular endothelial growth factor.
q
Bone marrow mesenchymal stem cells.
r
Bone morphogenetic protein-2.
s
Outgrowth endothelial cells.
Biotechnology Advances 36 (2018) 68–91
Table 4
Applications of SF/HApcomposites for bone tissue engineering.

Materials Processing method Cell Structural properties In vitro/in vivo key findings Ref.
M. Farokhi et al.

a b c −1 −1
SF /HAp MAPLE SaOs2 Exhibition of 1540 cm amide II, 1654 cm Improved proliferation of cells with normal (Miroiu et al., 2010)
amide I, 1243 cm− 1 amide III peaks in the FTIR morphology after 72 h culture on SF and HAp/SF
spectrum of SF, and observing of 1027 cm− 1 only films
for HAp and HAp/SF composite groups
SF/HAp Layer solvent casting/Freeze BMSCsd Decreasing pore average size with increasing SF No variation in the proliferation rate of BMSCs on (Gholipourmalekabadi et al., 2015)
drying content from 5% (350 ± 67 μm) to 10% HAp/SF and no toxicity, good in vivo
(112 ± 89 μm), more uniform pores in the biocompatibility of SF containing 5% HAp, no
scaffolds containing 5% HAp significant increase in the average number of
lymphocytes in long and short periods
SF/HAp Casting – Exhibition of silk II conformation in the secondary – (Ming et al., 2014b)
structure of SF films, no impact of HAp content in
the solution on the structure of SF films, Significant
effect of HAp content on rheological behavior of SF
solution
NSFe/HAp Electrospinning and soaking Osteoblast Higher crystallinity and thermal stability of Higher cell viability on electrospun NSF/HAp (Sekar et al., 2016)
electrospun NSF/HAp scaffold compared with pure scaffold than pure NSF and NSF/HAp formed by
NSF and NSF/HAp formed by soaking method soaking method
CSf/SF/nHAp Salt fractionation MG-63 Scaffolds exhibited 90.5–96.1% porosity with Suitable for increasing the proliferation and (Qi et al., 2014)
100–300 μm pore size, better mechanical behavior adhesion of cells, higher efficacy of CS/SF/nHA
of tri-component scaffold compared with bi- composite porous scaffold than other groups
component scaffolds
CS/SF/HAp Co-precipitation – Characteristics of primary HAp crystals: needle- – (Wang and Li, 2007)
like shape, 20–50 nm long, and ~ 10 nm wide,
Chemical interactions between amino group of
chitosan or amid bands of SF with Ca2 + ions

79
SF/COLg/HAp Particulate leaching MG-63 Displaying porous structure with adjustable Significantly higher biocompatibility at days 7 and (Mou et al., 2013)
macropores with high interconnectivity, uniform 14 compared with day 3, suitable features of
dispersion of SF and collagen within the scaffold scaffold in inducing the cellular proliferation, and
potential of the scaffold to stimulate the cellular
infiltration and migration
PVAg/SF/HAph Freezing and thawing – No agglomeration and uniform distribution of silk – (Zhang et al., 2012)
in the composite hydrogel, high impact of silk on
elastic modulus of PVA-HAp-Silk composite
hydrogel

a
Silk fibroin.
b
Hydroxyapatite.
c
Matrix Assisted Pulsed Laser Evaporation.
d
Bone marrow stromal cells.
e
Non-mulberry silk fibroin.
f
Chitosan.
g
Collagen.
h
Polyvinyl alcohol.
Biotechnology Advances 36 (2018) 68–91
M. Farokhi et al. Biotechnology Advances 36 (2018) 68–91

structure of nHAp/SF powder and SF matrix. The good interfacial in- bone tissue (Vallet-Regi and González-Calbet, 2004). nHAp possess
teractions between the nHAp/SF powders and the SF matrix promoted great characteristics compared with HAp such as high sinterability and
cellular proliferation, attachment, and differentiation (Liu et al., 2011). densification as a result of high surface area that lead to improved
The superior biological behavior of nHAp powders when compared to mechanical strength and toughness (LeGeros, 1993). Additionally, the
micro-scaled particles provides options for many biomedical applica- appropriate biocompatibility, osteoconductivity, the ability to promote
tions. Recently, nHAp particles 30–100 nm were homogeneously dis- the angiogenesis, and inducing the attachment, migration, and differ-
persed in a collagen/silk fibroin (Col-SF) matrix and the electrostatic entiation of osteoblast make it useful for bone regeneration (Hassan
interaction between Ca2 + of the inorganic phase and carboxyl or amino et al., 2016). However, nHAp is fragile, less flexible with intrinsic
groups of organic phase generated nanocomposites (Col-SF/HAp) with hardness that make it difficult to shape into the desired form which
suitable structural properties and biocompatibility. Additionally, the limits its applications in repairing load-bearing bone defects. In order to
existence of SF in the structure of nanocomposites enhanced elastic hamper the aforementioned drawbacks, nHAp in blended with other
modulus in comparison to Col/nHAp (Chen et al., 2014b). In another polymers such as SF to form nanocomposites with high osteo-
study, it was observed that SF affected the mineralization of HAp na- conductivity suitable for orthopedic surgery (Sun et al., 2011).
nocrystals. SF was able to significantly increase the growth of HAp
nanocrystals at pH 8 in room temperature. The strong chemical inter- 4.3. Silk nanofibers containing hydroxyapatite for bone tissue engineering
action between HAp and the SF protein shifted the amide II peak from application
1517 to 1539 cm− 1 in FTIR spectra, which indicated that HAp crystals
were replaced by carbonated HAp (Kong et al., 2004). SF not only in- The high surface area of nanofiber scaffolds makes them useful for
fluenced the morphological aspects of HAp crystals but also affected the protein absorption and binding to cell membrane receptors. The ab-
water dispersibility of HAp. Moreover, SF surrounded HAp nanocrystals sorbed proteins can change the structural conformation of the na-
to form a layer with a negative charge to prevent aggregation of nHAp noscale scaffolds along with generating more binding sites that make
particles in aqueous solution (Huang et al., 2014). Previous studies also them more suitable for tissue engineering. Electrospinning is one of the
confirmed the regulatory activity of SF on biomineralization of calcium most useful techniques for fabricating nanofibrous scaffolds.
salts (Takeuchi et al., 2005; Wang et al., 2012). However, the distance Electrospun mats are 3D structures with appropriate porosity and high
between SF and HAp nanocrystals played an important role in the in- surface area to volume ratio that mimic to some extent the structure of
tensity of their interactions. At longer distances, the interactions be- the ECM. Therefore, electrospun nanofibers are considered as potential
came weaker (Zhang et al., 2013). Thus, SF and nHAp particle ratios are scaffolds for tissue engineering applications due to their ability to
important factors in controlling the growth of nanocrystals and crystal support cellular proliferation, adhesion, and differentiation (Agarwal
morphology. Accordingly, regular SF nanostructures were implemented et al., 2008; Liang et al., 2007). However, using electrospun SF nano-
in order to evaluate their effect on the formation and morphology of fibers is still limited for bone tissue applications because of their low
HAp nanocrystals. SF nanospheres induced the formation of rice-like mechanical properties. In order to overcome this limitation, inorganic
nHAp, while SF nanofibers stimulated the formation of HAp nanofibers ceramics can be incorporated into the polymeric matrix. However, the
(He et al., 2012a). Therefore, it was suggested that different nanos- aggregation of a ceramic particle in the SF matrix can interrupt the
tructures of SF can be considered as templates for the formation of electrospinning process. Therefore, further investigations are needed to
different morphologies of HAp. Silk fibroin/sodium alginate (SF/SA) fabricate appropriate SF/HAp nanofibers for bone tissue engineering
hydrogels induced the formation of HAp nanorods with rectangular purposes. In several studies, the aggregation of HAp particles in SF/HAp
column morphologies at room temperature. The strong interaction be- nanocomposites was reported (Kim et al., 2008; Li et al., 2006; Zhang
tween SF/SA nanofibers and Ca2 + ions in HAp had significant effect on et al., 2010). The aggregation can reduce the mechanical strength of the
the morphology of nanocrystals (Ming et al., 2014a). Generally, the SF/HAp composites (Kim, 2007; Song et al., 2008a) and can result in
phosphorylated motifs on the surface of SF play an important role in non-uniform physical and chemical structures in these scaffolds.
triggering the nucleation and formation of apatite crystals. Each of A useful strategy for enhancing the uniform dispersion of HAp
apatite crystals grew in size to form a thick mineralized layer during particles in the SF matrix is the surface modification of the SF polymer
mineralization. Adding foreign molecules may trigger the initial for- by γ-glycidoxypropyltrimethoxysilane (GPTMS). The epoxide func-
mation of HAp crystals; however, mineral proliferation may be delayed tional groups of GPTMS are able to attach to amino groups in the side
due to inhibition of the growth sites (Boskey, 1998; LeGeros, 1990). chains of SF. Concomitantly, another side group of GPTMS can interact
Furthermore, one of the main challenges of using biomaterials to induce with the hydroxyl groups on the HAp surface (Arafat et al., 2011; Kim
the process of bone regeneration is the bacterial resistance before et al., 2014b). Electrospun SF nanofibers containing < 20% of well-
complete bone repair. Thus, the biomaterials should have the ability to dispersed nHAp modified by GPTMS showed high mechanical strength
stimulate the regeneration before the adhesion of bacterial species. For for load bearing applications. However, higher concentrations of HAp
this purpose, SF/nHAp hydrogen was modified with gold nanoparticles (> 20%) disrupted the polymeric chains that affect the mechanical
(AuNPs) and AgNPs in order to induce the antimicrobial effect (Ribeiro behavior of the scaffolds (Kim et al., 2014b). Similarly, using > 20%
et al., 2017). The obtained results revealed that the hydrogel containing HAp content showed decreased bending modulus due to local ag-
high concentration of AgNPs had strong capability to reduce the gregation of nHAp (Yang et al., 2016).
number of different gram positive and gram negative bacterial strains Electrospinning of SF/HAp blends faces another limitation such as
such as S. aureus (MSSA), S. aureus (MRSA), S. epidermidis, Escherichia immediate precipitation of nHAp before ejection. A three-way stopcock
coli (E. coli), and P. aeruginosa. However, the hydrogen comprising connector was developed for rapid mixing of various polymeric solu-
AuNPs did not show a linear antimicrobial effect against S. epidermidis. tions and then rapidly injecting under the electric field to form blended
The hydrogels containing AuNPs ≥ 0.5% had suitable antibacterial nanofibers (Sheikh et al., 2013). Based on FTIR analysis, the chemical
activity toward MRSA and P. aeruginosa; while, those containing interaction between SF and nHAp changed the conformation of SF from
AuNPs ≥ 0.1% reduced the number of MSSA and E. coli. It seems that random coil to β-sheet, which improved the mechanical strength of SF
the higher surface activity of AgNPs in comparison to AuNPs might be nanofibers (Sheikh et al., 2013). It was also showed that combining
responsible for this observation (Ribeiro et al., 2017). electrospun SF nanofibers with high pure HAp and silver nanoparticles
Taken together, the biological HAp existed in the natural bone has (AgNPs) produced a suitable scaffold with antimicrobial properties for
nanometric thickness and length with nanoscopic plate-like or rod-like tissue engineering (Sheikh et al., 2014). AgNPs enhanced the anti-
shape. Thus, applying nHAp would be a good bone substitute with high microbial activity of nanofibrous structures. SF nanofibers containing
regeneration capacity that highly mimic the mineral phase of natural 0.5% silver nitrate showed no toxicity toward NIH 3 T3 fibroblast cells;

80
M. Farokhi et al. Biotechnology Advances 36 (2018) 68–91

however, using higher concentrations (1.0 or 1.5%) of silver nitrate 4.4. Silk fibroin/hydroxyapatite for drug delivery in bone tissue engineering
revealed typical toxicity (Sheikh et al., 2014). New biocomposites based
on electrospun SF nanofibers containing mesoporous bioactive glass/ To date, many attempts have been made to develop an appropriate
hydroxyapatite nanocomposite (MGHA) were reported (Liu et al., carrier based on novel materials for drug delivery purposes. These
2014). Recently, mesoporous bioactive glass (MBGs) has been in- systems control the release rate of drugs in a more sustained manner
troduced as potential candidates for drug delivery and bone tissue en- that can be helpful for decreasing side effects. Different structures such
gineering applications. The ternary SiO2–CaO–P2O5 structure of MBG as membranes, micro- and nanoparticles, fibers and three-dimensional
possess useful properties such as high surface area, degradability, and scaffolds have been introduced as carriers for different biomolecules
large pore volume in comparison to non-mesoporous bioactive glasses (Balmayor, 2015; Zarrabi et al., 2014). Most of these structures are
that make them suitable for inducing bone formation (Wu et al., 2013; mainly comprised of polymers; however, using inorganic materials can
Xia and Chang, 2006). Moreover, adding MGHA particles into SF so- be also useful.
lution increased fiber diameter related to conductivity affecting fiber Bioceramics are attractive inorganic materials to incorporate
diameter and particle dispersion. The decrease in the conductivity of bioactive molecules or even cells, without denaturation or loss of
the blend solution was controlled by using higher MGHA content and bioactivity during preparation or implantation (Ginebra et al., 2012).
less SF solution in the matrix. Using MGHA as a reinforcing component Large surface area, narrow pore size distribution, and well-ordered pore
in the structure of SF-based nanocomposites improved swelling, surface systems make CaPs useful candidates for drug delivery (Bose and
hydrophilicity, and pore size, while tensile strength decreased. SF/ Tarafder, 2012; Vallet-Regí et al., 2007). Moreover, CaP is useful for
MGHA scaffolds induced the formation of larger bones than pure SF controlling the local release rate of antibiotics (Arcos and Vallet-Regí,
after 4 and 8 weeks post-implantation. Additionally, in both implants, 2013; Verron et al., 2010), anticancer, anti-inflammatory, and analgesic
new bones containing osteocytes were directly formed near the scaf- drugs (Ginebra et al., 2006a) and growth factors (Bose and Tarafder,
folds. The capability of SF/MGHA composites in bone regeneration may 2012; Ginebra et al., 2006b). Oral administration of CaPs is impossible
be related to the chemical interactions between the implant and host because of rapid degradation at low pH of the gastric system. Therefore,
tissue, along with their surface bioactivity (Liu et al., 2014). drug delivery systems for the intestinal system (pH 8–9) with a con-
In a recent study, coaxial electrospinning was applied in order to trolled and targeted release profile are useful. However, current in-
prepare a core-shell structure consisting of HAp and tussah silk fibroin vestigations have been focused more on drug delivery to locally or di-
(TSF) (Shao et al., 2016b). The crystalline region of TSF has a higher rectly to bone tissue or as coatings (Kolmas et al., 2016).
amount of alanine (Ala) in comparison to B. mori silk; therefore, Ala CaPs also absorb many chemical molecules on their surface, which
repeats are the main sequences in the structure of TSF, while domes- is useful for applications such as chromatography, and different bio-
ticated silk is mostly comprised of Gly-Ala-Gly-Ser repeats. In addition, molecules isolation and purification (Ginebra et al., 2006a; Tiselius
the existence of an Arg-Gly-Asp motif in TSF structure induced specific et al., 1956). For instance, this criterion has been used for purification
cellular attachment, useful for biomedical purposes (He et al., 2013; of bone growth factor through HAp chromatography (Urist et al.,
Zhang et al., 2009). An increased content of nanoparticles in the core of 1984). The high affinity of HAp to different biomolecules can provide
prepared fibers increased the tensile strength. In other words, the options for developing novel matrixes for drug delivery systems.
content of core nanoparticles can be used to gradually change the However, applying HAp-based carriers faces challenges due to their low
mechanical properties of the composite from soft to rigid. A similar capacity for loading drugs and the inability to control the release ki-
trend was also seen in the Young's modulus of the composites. This netics of such biomolecules. In contrast, silk is a useful biomaterial with
scaffold showed promising properties for bone tissue engineering ap- high drug loading capacity and capability in controlling the release rate
plications due to potential in inducing cellular proliferation, attach- of drugs. Different drugs and biological macromolecules can be also
ment, ALP production, and biomineralization (Shao et al., 2016b). It loaded on silk-based structures without changing their bioactivity
was also found that the robust interaction between nHAp and SF na- (Kundu et al., 2010; Shi et al., 2013a). For this reason, considerable
nofibrills induced the formation of flower-like structure that potentially efforts have been devoted to developing combinations of the intrinsic
facilitated the mineralization of HAp in situ. Moreover, the good me- bone regeneration potential of HAp blended with SF, with the ability to
chanical properties and thixotropy with storage modulus (G’) make SF/ incorporate drugs or other active molecules relevant for different
HAp hydrogel useful to recover to over 85% in 50 s after applying a therapeutic needs.
large shearing strain (5000%) (Mi et al., 2016). Electrospun SF fibers containing nHAp/bone morphogenetic protein
Generally, it is necessary to use nanofibrous structure for bone 2 (BMP-2) were used to induce bone formation in vitro using human
tissue engineering applications due to the nanometric nature of ECM bone marrow-derived mesenchymal stem cells (hMSCs) (Li et al., 2006).
components existed in the natural bone. The bone is a natural nano- The authors found that it was impossible to electrspin the SF solution
composite comprising nanoscale HAp with about 20–80 nm in length with the concentration > 18% because the insufficient viscosity and
that are dispersed in an organic collagenous matrix (Rho et al., 1998). surface tension of this solution. Higher concentrations of SF solution
The individual collagen chains have the length about 10 nm which induced the gel formation during the electrospinning process. Fur-
reach to 500 nm after collagen fiber formation. Thus, using nanoscale thermore, nHAp had a tendency to form aggregations after suspending
bone constructs not only mimic the structure of natural bone but also in aqueous media and to overcome to this issue the SF and polyethylene
can provoke the cellular activities in terms of adhesion, growth, dif- glycol (PEG) solutions were prepared in 0.001 M phosphate buffer
ferentiation, and expression of various proteins (Jiang et al., 2015). (pH 6.8) that reduced the aggregation. It was observed that the nHAp
Despite the potential of electrospun mats in resembling the structure of were effectively embedded in the SF/PEO fiber scaffold. The nano-
collagenous fibers, the weak mechanical behavior compared with particles showed different fashions in dispersing in the polymeric ma-
porous materials might be challenging for bone tissue engineering ap- trix in which they were well-oriented along the fiber axis in some re-
plications. Therefore, it is crucial to find a way for increasing the me- gions while agglomerated in the other parts. It seems that high viscosity
chanical strength of fibrous SF nanofibers to reach to adequate ten- of SF/PEO blend solution led to non-homogenous dispersion of in-
sional or compressional stresses similar to skeletal bone (Fu et al., 2011; dividual particles. Moreover, this study revealed that co-processed
Wei and Ma, 2004). It seems that addition of ceramic phases in SF fibers BMP-2 supported more calcium deposition and more upregulation of
are good strategies to improve the mechanical properties of fibrous bone-specific markers (BMP-2, Col I and bone sialoprotein-II) in com-
scaffolds (Ito et al., 2005). parison to control groups. This observation suggested that silk nanofi-
bers were useful for BMP-2 delivery, which broadened the application
of this biomaterial in bone tissue engineering applications. Based on

81
M. Farokhi et al. Biotechnology Advances 36 (2018) 68–91

XRD spectra, the crystallinity of apatite on SF/BMP-2 scaffolds was PCL for dual delivery of rhBMP-2 and TGF-β (Bhattacharjee et al.,
higher than bulk SF scaffolds. Additionally, the incorporation of nHAp 2016b). The 4-methacryloyloxyethyl trimellitate anhydride (4-META)
with SF/BMP-2 during electrospinning induced bone formation (Li was used to modify the NSF-PCL matrix and the nHAp were deposited
et al., 2006). In another study, silk was used as a surface stabilizer and on them by electrodeposition at two separate potentials (3 V and 5 V).
template in order to enhance surface properties of nHAp (Ding et al., The strong ionic interaction between nHAp and 4-META promoted the
2016a, 2016b, 2016c). The core-shell structure of these composite deposition of nHAp on the NSF-PCL matrix. The rhBMP-2 and TGF-β
materials facilitated the interaction between silk on the outer surface of with a 1:1 ratio were loaded on the surfaces of the scaffold. After
nanoparticles with BMP-2 resulting increase the loading capacity of 4 weeks, rhBMP-2 and TGF-β exhibited about 20.02% and 20.93% re-
nHAp. This study showed that removing silk from nHAp had no effect lease kinetics from NSF-PCL scaffolds, respectively. It seemed that the
on the morphological aspects of nHAp; however, aggregation particles covalent interaction between the growth factors and HAp nuclei re-
were observed because of the elimination of repulsive forces from SF. duced the release kinetics. This structure had many favorable properties
Several studies confirmed that the high loading capacity of silk was for bone tissue engineering because it contained SF with RGD motifs
related to electrostatic interactions with BMP-2 (Shi et al., 2013a; Shi with the ability to induce cellular attachment and proliferation, calcium
et al., 2013b). Moreover, HAp particles alone showed 90% release of phosphate deposition with high potential in promoting the proliferation
BMP-2 during 2 h, it was about 50% and 70% for silk alone and silk- of osteoblasts, and the ability for dual growth factor delivery. The re-
nHAp composite groups after 21 days, respectively. Despite the faster sults also revealed that the prepared structures increased ALP produc-
release kinetics of BMP-2 from silk-nHAp compared to silk particles, the tion, intensified mineralization and improved the expression of bone-
trend was linear. Moreover, BMP-2 loaded silk-nHAp showed more related genes. Random actin-stress fibers with dense cell colony de-
potential for inducing osteogenesis from bone mesenchymal stem cells posits were observed across the nanofibers loaded with dual growth
(BMSCs) in comparison to BMP-2 loaded HAp particles or silk scaffolds factors via confocal laser microscopy. However, these outcomes were
(Ding et al., 2016a, 2016b, 2016c). lower in the single growth factor-loaded scaffolds, which confirmed the
In a study, a BCP consisting of HAp/β-TCP (60/40% ratio) and potential utility of dual growth factor delivery in bone tissue en-
rhBMP-2/SF microsphere was developed which were then integrated gineering applications (Bhattacharjee et al., 2016b).
into a BCP/rhBMP-2/SF composite (Chen et al., 2014a, 2014b). Based Similarly, chemical deposition method was used for preparing nHAp
on SEM, HAp particles showed a needle-like structure, while β-TCP had particles. Electrospun PCL nanofibers were prepared with different
hexahedral-like morphology. The rhBMP-2/SF had the mean diameter contents of nHAp (0%, 25% and 50%) (Bhattacharjee et al., 2016c).
of 398.7 ± 99.86 nm with loading capacity of 4.53 ± 0.08%. For NSF was then grafted on the nHAp-PCL composites. It was observed
pore formation, different concentrations of sodium chloride (NaCl; 50, that initial addition of nHAp had a positive effect in increasing the
100, 150, and 200 w/v) were applied. It was observed that using NaCl mechanical strength of NSF grafted PCL nanofibers. However, the
with 50, 100 and 150% (w/v) concentrations had no effect on the higher tensile strength, toughness, and ductility was detected in those
structure materials during the desalination process; while, 200% NaCl matrices containing 25% nHAp. Although the samples containing 50%
easily damaged and degraded into particles or powder. Moreover, nHAp showed higher mechanical properties and toughness than NSF-
rhBMP-2 showed a dual-phase release profile, with both fast and slow PCL nanaofibers without nHAp, but it was weaker than samples con-
rates, from the blended composites during 28 days. In the first three taining 25% nHAp. In addition, rhBMP-2 and TGF-β were coupled to
days, a burst and the fast release of protein (about 47.3 ± 3.1%) were the scaffolds by using carbodiimide coupling. The release rate of
observed. Afterwards, the cumulative release was about 72.2 ± 4.6% rhBMP-2 and TGF-β was about 21.34% and 22.6% after 4 weeks, re-
detected between days 3 to 14, and about 90.4 ± 5.3 between days 14 spectively. These low release rates may be related to the strong inter-
and 28 (Chen et al., 2014a, 2014b). The addition of organic materials actions between the functional groups of growth factors with active
into inorganic phases exhibits many advantages, such as increasing binding sites of SF. In the first 3 days, a burst release profile of both
degradability, promoting migration and recruitment of osteoblast cells growth factors was observed, which may be due to the presence of non-
on the surface of the materials, and increasing rates of mineralization. conjugated or physically conjugated growth factors (about 22%). After
The efficacy of BCP/rhBMP-2/SF in a sheep lumbar fusion model 3 days, the remaining growth factors that were covalently bonded to
was also evaluated (Chen et al., 2015a, 2015b). For this purpose, BCP the scaffolds were slowly released during 4 weeks. The fabricated
and rhBMP-2/SF microspheres were blended to form BCP/rhBMP-2/SF scaffolds containing both growth factors provided optimal micro-
composites that were implanted into the disc spaces of 30 sheep at the environments for cellular migration, proliferation, differentiation, and
levels of L1/2, L3/4 and L5/6, randomly. The growth factor exhibited enhanced expression of bone specific markers (Bhattacharjee et al.,
an initial burst release about 39.1 ± 2.8% during the first 4 days, and 2016c). The cross-talk between signaling of these growth factors and
then had sustained release over 28 days (cumulative release different signaling cascades such as Wnt, Hedgehog, Notch, MAP, and
rate ~ 81.3 ± 4.9%). After implantation, the BCP/rhBMP-2/SF FGF play important roles in triggering the differentiation of osteoblasts
showed significantly better histological aspects such as in semi-quan- and consequently, bone formation (Chen et al., 2012). As mentioned
titative CT scores, fusion rates, histologic scores and fusion stiffness in earlier, multiple biomolecules can help to improve the bone healing
bending in all directions than the BCP or BCP/rhBMP-2 study groups. process. In addition to growth factor delivery, other biomolecules such
The results recommended that using low doses of rhBMP-2 in the as some drugs would be useful in bone tissue engineering applications.
structure of SF microspheres improved bone fusion in sheep via BCP Recently, silk–HAp films were prepared using casting method for in-
constructs (Chen et al., 2015a, 2015b). corporation of anti-osteoporotic drugs like clodronate (non-nitrogenous
Despite the beneficial characteristics of controlled release systems in bisphosphonate) and alendronate (nitrogenous bisphosphonate)
biomedical fields, the important parameters for mimicking the natural (Hayden et al., 2014). Biological aspects of THP-1 human acute
process of bone repair remain challenging to emulate. In many ex- monocytic leukemia cell line-derived osteoclasts and human mesench-
periments, a single growth factor with controlled release kinetics is ymal stem cell-derived osteoblasts such as calcium deposition, meta-
used; however, developing a system with the ability to control the re- bolic activity and ALP production on the prepared films were evaluated
lease of multiple growth factors is more useful for bone regeneration during 12 weeks. Clodronate-complexed scaffolds showed higher me-
(Kolambkar et al., 2011). Therefore, attempts have been pursued to tabolic activity and roughness after culturing osteoblasts and co-cul-
develop delivery systems containing multiple growth factors for the tures of osteoblasts with osteoclasts. Although the low dose of alen-
synergistic effect on bone healing (Chen et al., 2010; Farokhi et al., dronate enhanced the metabolic activity and calcium deposition of
2016b; Shah et al., 2011). osteoblasts (Fig. 7), some toxicity was observed using the high doses
The electrospinning method was used to produce nanofibrous NSF/ against osteoclasts, osteoblasts and co-cultures. These data may result

82
M. Farokhi et al. Biotechnology Advances 36 (2018) 68–91

Fig. 7. Effects of bisphosphonates on calcium deposition. Calcium content of films was measured after 4, 8 and 12 weeks of culture and expressed as mean ± SD. Top left: osteoblasts
cultured on clodronate-loaded films. Top right: osteoblasts cultured on alendronate-loaded films. Bottom left: co-cultures cultured on clodronate-loaded films. Bottom right: co-cultures
cultured on alendronate-loaded films. ∗P < 0.05, ∗∗ P < 0.01, ∗∗∗ P < 0.001. Copyright © 2014 Elsevier B.V. Reprinted with permission from (Hayden et al., 2014).

from different mechanisms of action of nitrogenous and non‑ni- challenges of using ESCs such as ethical, religious and immune rejection
trogenous bisphosphonates. The toxicity may be related to drug class. or the high costs of isolating and culturing iPSCs. However, iPSCs and
Nitrogenous bisphosphonates cause apoptosis in various cell types at ESCs are still good candidates for biomedical and tissue engineering
concentrations similar to those that cause osteoclast apoptosis, and by applications due to their beneficial characteristics (Nori et al., 2011;
the same mechanisms (Idris et al., 2008). Moreover, the inhibitory ef- Song et al., 2010; Takahashi et al., 2007). In recent years, using dif-
fect of nitrogenous bisphosphonates on mineralization is independent ferent stem cell sources for different kinds of tissue engineering pur-
of the mechanism, which leads to cell apoptosis (Idris et al., 2008). poses has been introduced. Besides, it is more cost-effective to promote
Additionally, the ALP production was not changed using clodronate; the differentiation of stem cells using appropriate scaffolds without
however, alendronate decreased the ALP activity in a dose-dependent growth factors for optimal tissue repair. For example, silk/HAp com-
manner (Hayden et al., 2014). posies have been introduced as a potential scaffold for promoting the
Generally, drugs can be loaded in CaPs such as HAp by tailoring the stem cell differentiation. Accordingly, the NSF/nHAp scaffold showed
porosity with large surface area during synthesis (Arcos and Vallet- better biological and physiochemical properties for stimulating the
Regí, 2013; Vallet-Regí et al., 2007). Although high porosity may de- differentiation of bone marrow-derived stem cells (CBhMSC) into an
crease mechanical strength, pores with an appropriate size can act as a osteogenic lineage in comparison to NSF electrospun scaffolds because
host for drug molecules (Arcos and Vallet-Regí, 2013). The release of the existence of well-dispersed nHAp over NSF matrix. After 24 h, the
drug can be controlled by altering the amount of porosity of CaP cells were attached and aligned along the axis of fibers that form or-
structures (Kim and Tabata, 2015). Furthermore, a homogeneous dis- iented morphology. The ability of the scaffold to promote the osteo-
tribution of drug can be achieved in well-ordered pores (Vallet-Regí genic differentiation of CBhMSC cells were confirmed by the slight
et al., 2008). In addition to the beneficial features of CaPs as drug decrease in the proliferation rate and the increase in the expression
carriers, nHAp retains an advantage as a major constituent of natural level of ALP and osteogenic genes (Panda et al., 2014). Recently, NSF
bones. The degradation of CaP produces Ca2 + and PO43 − ions that are composites containing silk fibers and HAp was fabricated as bone
found in high concentrations in the bloodstream. Moreover, CaPs, substitutes (Gupta et al., 2016). The HAp was uniformly dispersed in
particularly HAp, are relatively insoluble at pH 7.4, thus there is no the structure of highly porous and interconnected composite scaffold
immunogenic response, and no cytotoxic degradation products (Bose that formed an appropriate substrate for inducing the proliferation and
and Tarafder, 2012). migration of cells by preparing the suitable environment for transfer-
ring oxygen and nutrients. Moreover, the great mechanical behavior of
the composite provided a supportive matrix for growth and differ-
4.5. Silk fibroin/hydroxyapatite for stem cell differentiation entiation of MG63 and hBMSCs cells due to high surface area/rough-
ness and osteoconductivity. The osteogenic potential of the scaffolds
Several tissues are important source of therapeutically relevant were further confirmed by assessment the expression level of osteogenic
differentiated cells but the inevitable difficulties are in harvesting suf- genes like Col I, osteopontin, osteocalcin, and sialoprotein, ALP pro-
ficient cells for implantation. Lineages such as neurons and cardiac duction, and ECM deposition (Gupta et al., 2016).
cells, being terminally differentiated and non-regenerative, impose the Another study showed that adding NSF (10 wt%) into PLGA/gra-
biggest challenge. Among different stem cell lineages, pluripotent cells phene oxide (GO) nanofibers and then mineralization in simulated body
have been focus in biomedical applications due to high self-renewal fluid (SBF) can significantly enhance the hemocompatibility, protein
capacity, prolonged undifferentiated options, and the ability to differ- absorption, surface hydrophilicity, and the mechanical properties.
entiate into multiple cell types under suitable stimuli (Zhao et al., Besides, the conductivity of the electrospinning solution was increased
2013). Adult stem cells are other potential cells useful for tissue en- from 0.39 to 0.78 mS/cm after adding 1 wt% GO into PLGA/NSF na-
gineering because they can fully regenerate damaged tissue and im- nofibers that significantly decreased the fiber diameter from 321 nm to
prove the process of tissue repair. Recently, autologous adult stem cells, 89 nm. Moreover, biological tests confirmed that scaffolds not only
such as cardiac stem cells (CSCs), were found favorable when compared support hMSC adhesion and proliferation, but also promote osteogen-
to induced pluripotent stem cells (iPSCs) and embryonic stem cells esis and ALP production (Shao et al., 2016c).
(ESCs) for treating myocardial infarction (MI) and chronic heart failure Similarly, GO-HAp/SF composite was prepared by biomineralizing
(CHF) (Shafiq et al., 2016). Moreover, adult stem cells do not face the

83
M. Farokhi et al. Biotechnology Advances 36 (2018) 68–91

carboxylated GO sheets, blending with SF, and freeze-drying for eval- osteogenic properties, the silk-HAp composite with silk nanofibers in
uating the differentiation of MSCs. The GO-HAp (1:4)/SF composites hydrogels was prepared. For this, an injectable system was fabricated
showed significantly higher mechanical strength in comparison to by blending thixotropic silk nanofiber hydrogels with water dispersible
HAp/SF scaffolds that had better compressive strength (85 MPa) and silk-HAp nanoparticles to form injectable nanoscale systems. The
compressive modulus (938 MPa). The protein absorption of GO-HAp composites had about 60% w/w HAp content for optimally mimic the
(1:4)/SF were also 1.8-fold higher than HAp/SF scaffolds. Furthermore, bone environment. The physical strength of the system was also im-
the GO-HAp/SF composite was able to potentially improve the adhesion proved to 21 kPa by adding of HAp that was suitable for stimulating the
and proliferation of MSCs, increase the expression level of osteocalcin, osteo-differentiation. The osteogenesis was more enhanced by using
and differentiation of osteoblast cells (Wang et al., 2017). composite hydrogels in comparison to silk nanofiber hydrogel without
Aqueous-derived porous SF scaffolds were fabricated and con- HAp. The silk-HAp composite hydrogel was able to treat irregular bone
sidered as templates to deposit apatite on the pore surfaces via mixing defects after injection (Ding et al., 2017).
with polyaspartic acid (PA) during processing, followed by miner- The main purpose of tissue engineering is to develop functional
alization with CaCl2 and Na2HPO4 (Kim et al., 2008). The salt-leaching materials with suitable biocompatibility that highly mimics the char-
method using granular NaCl with particle size of 850 ∼ 1000 μm was acteristics of the target tissue. For this, using engineered biomaterials is
used to prepare the scaffolds. Moreover, the ratios of SF-PA (w/w) of interest due to rational design, adequate biocompatibility and bio-
blends were as 100/0 (group I), 95/5 (group II), 90/10 (group III) and degradability, tunable structural properties, and functionality.
80/20 (group IV). The scaffolds were then used for seeding hMSCs Recombinant silk fusion proteins are an example of engineered mate-
under osteogenic condition in the presence or absence of BMP-2 for rials that have these features. Mineral-binding sequences are rich in
6 weeks. After 24 h, the aqueous solution of SF/PA converted to hy- acidic residues (aspartate, glutamate, phosphoserines) that can increase
drogel and formed a porous and water-stable structure. The apatite the net negative charge of engineered materials. These sequences can
crystals were deposited using salt deposition because PA produced interact with positively charged HAp nanocrystals (Addison et al.,
nucleation sites at the interfaces of the hydrophobic silk structures. It 2010). Recently, an organic-inorganic hybrid system was developed
was found that the increased polyaspartic content increased HAp de- based on genetically engineered spider silks. The organic phase of
position on the scaffolds. Accordingly, it is well-established that pro- spider silk inspired domain (SGRGGLGGQG AGAAAAAGGA GQGGYG-
teins containing aspartic acid-rich motifs are able to interact with cal- GLGSQGT)15 maintained the stability of the scaffold and facilitated
cium salts in a specific fashion (Addadi and Weiner, 1985; Fujisawa different modes of processing; while, the inorganic domain of SF
et al., 1996). It was also showed that the calcium deposition and ALPase VTKHLNQISQSY (VTK) for HAp binding was involved in regulating the
activity were almost the same in mineralized silk scaffolds in the ab- osteogenesis and biomineralization. The VTK peptide localized on re-
sence of BMP-2. Moreover, the osteogenic media, whether using BMP-2 combinant fusion protein was able to provoke the calcification but it
or not, also stimulated similar ALPase activity of MSCs after 6 weeks in did not any effect in improving the physical strength of the recombinant
groups I, II, and III. The group IV showed higher ALPase activity in the structure. It was also observed that adding VTK peptide on both c-
presence of BMP-2. All the mineralized scaffolds revealed more calcium terminal and N-terminal sides of recombinant silk could significantly
deposition with BMP-2, while no calcium deposition was detected in enhance the formation of crystalline HAp. All the prepared constructs
those scaffolds without mineralization (control groups) after 6 weeks. was able to provoke the attachment, proliferation, and differentiation
Based on van Kossa staining, more calcium deposition led to higher of hMSCs (Dinjaski et al., 2017). Salt-leached SF scaffolds were also
osteoconductivity of the scaffolds. Immunocytochemistry demonstrated prepared followed by functionalization and mineralization in CaCl2
Col I was expressed in all groups. Additionally, the expression of Col I ethanol solution/K2HPO4 aqueous solution. The scaffolds were then
was increased in groups II and III in the presence of BMP-2 (Fig. 8). It mineralized by immersing in the simulated body fluid (SBF) for 4 days.
was concluded that higher amounts of apatite in the structure of porous The authors claimed that the duration of soaking in SBF would be de-
silk constructs along with the existence of BMP-2 increased the osteo- creased by using this method that consequently protect the SF-nHAp
conductivity of the scaffolds (Kim et al., 2008). scaffold from degradation during the mineralization process. The mi-
Despite the effectiveness of using silk/HAp as bone constructs, the crospores of nHAp with a size range of 100 to 250 nm were formed on
injectable form of bone materials is less investigated. The main draw- the surface wall of SF. Incorporating of nHAp particles with the SF
back of bone substitutes is the lack of suitable structure for filling the scaffolds significantly improved the compressive strength and stiffness
irregular bone defects. To provide an injectable biomaterials with of SF (Liu et al., 2015a, 2015b). Furthermore, the cDNA microarray

Fig. 8. Immunohistochemical staining of collagen type I. Extracellular matrices (*) stained positive for Col I within the scaffold pores. The Col I matrix covered the entire mineralized
aqueous-based silk lattice (#) after 6 weeks of culture with BMP-2 (100 ng/mL; b, d, f, h) or without BMP-2 (a, c, e, g). Copyright © 2008 Elsevier B.V. Reprinted with permission from
(Kim et al., 2008).

84
M. Farokhi et al. Biotechnology Advances 36 (2018) 68–91

demonstrated changes in 23 calcium ion binding genes and 17 cell promoted the proliferation of rBMSCs and upregulated mRNA levels of
adhesion genes of BMSCs seeded on nHAp-SF scaffolds compared with ALP and Runx2 more than the other groups. Moreover, rBMSCs showed
bulk SF constructs. Among these genes, about 15 genes corresponded to higher levels of ALP expression when cells were cultured on SF scaffolds
cellular adhesion and 20 genes were attributed to calcium ion binding containing 30% HAp in comparison to 0% HAp-SF scaffolds. Higher
were decreased in expression level. Based on these observations, the calcium deposition was found by Alizarian red assay when 30% HAp-SF
cultured BMSCs on nHAp-SF scaffolds had round shapes due to less scaffolds were used versus 0% HAp-SF scaffolds (Ding et al., 2016c).
surface adhesion. The nHAp-SF scaffolds showed higher os- Therefore, it was suggested that SF scaffolds containing 30% HAp were
teoinductivity in vitro, significantly higher ALP activity on day 14, and potent structures to promote bone formation due to their ability to
more osteogenesis related to signaling pathway genes on day 7 com- induce mineral deposition and consequently, cell differentiation.
pared with the SF scaffolds. Moreover, the nHAp-SF scaffolds demon- Co-cultures of MSCs between ligament and bone cells on suitable
strated higher capability to induce bone regeneration in cranial bone sections of hybrid silk scaffolds induced cellular differentiation into
defects than bulk SF scaffolds. It was also shown that soluble IL-1a fibrocartilage lineage (He et al., 2012b). The section of the scaffold for
factor and biomolecules on the surface of BMSCs were played an im- seeding osteoblast cells was pre-coated with HAp. Real-Time PCR re-
portant role in provoking osteogenesis. Upon interaction between vealed that both osteoblast and fibroblast cells maintained their normal
BMSCs and nHAp, BMSCs regulated the activity of IL-1a by IL IR2. This phenotype during co-culture. However, co-cultured BMSCs differ-
study helped to understand better the possible mechanism of os- entiated toward the fibrocartilage lineage. The co-cultured BMSCs on
teoinductivity by nHAp (Liu et al., 2015a, 2015b). the hybrid SF scaffolds exhibited up-regulation of Col II, Sox9, aggrecan
Electrospun SF nanofibers containing nHAp and BMP-2 were pre- and fibrocartilage gene markers in comparison to mono-cultured
pared as bone construct with the ability to induce the differentiation of BMSCs. The data were confirmed by histological analysis and im-
BMSCs (Niu et al., 2017). The nHAp particles were successfully dis- munohistochemistry. The immunohistochemistry staining revealed that
persed in SF fibers. The increase in the content of HAp resulted in more aggrecan was the most prominent marker in the enthesis (the con-
aggregations of this particles in the SF matrix. The mechanical prop- nective tissue between tendon or ligament and bone). In addition, the
erties of the composite containing 20% HAp such as tensile strength, trilineage co-culture was supplemented with 10 ng/mL TGF-β3 and the
yield strength and Young's modulus were almost two-fold more than fibrocartilaginous differentiation of BMSCs by the TGF-β3-treated
pure SF scaffolds. BMSCs showed significantly more ALP activity on SF/ group was similar to those groups without TGF-β3. No calcium de-
HAp/BMP-2 scaffolds compared with SF scaffolds. Thus, SF/HAp/BMP- position was observed in the TGF-β3-treated group and it was suggested
2 scaffolds were considered as a potential structure in inducing osteo- that TGF-β3 could postpone the process of mineralization in chon-
genesis (Niu et al., 2017). It is assumed that synergistic effect of HAp drocytes (He et al., 2012b).
and BMP-2 improves the osteogenic capacity of the composite scaffold. In another study, leaching paraffin microsphere with the modified
This issue was also confirmed by Lu et al. (2015) or Liu et al. (2015a, temperature gradient-guided thermal-induced phase separation (TIPS)
2015b)) claimed that degradation of SF can increase the exposure of methods were applied to fabricate SF/HAp scaffolds with the potential
HAp to the cells during time. to mimic the integrated trilayered osteochondral substrates (Ding et al.,
Similarly, the bioactivity and the sustained release rate of BMP-2 2014). Based on the Micro-CT analysis, the consecutive tri-layer
was preserved in the nHAp-embedded SF scaffold by adjusting the ratio structure consisting a chondral (top) layer, intermediate layer, and
of loaded BMP-2 on SF and nHAp. The suitable BMP-2 release behavior bony (bottom) layer was observed in the integrated scaffold. Using TIPS
inside the structure of SF-nHAp scaffold increase the expression level of method resulted in the formation of oriented microtubularlike porous
Col I, ALP and Runx2 protein that resulted in higher osteo-conductivity. structure in the chondral layer and the paraffin-sphere leaching led to
Based on the results from in vitro and in vivo, the release of BMP-2 had formation of an interconnected macropore structure in the bony layer.
positive effect in inducing osteogenesis of the scaffold (Ding et al., The chondral and bony layers had porosities of 85.30% ± 1.80% and
2016a, 2016b, 2016c). 90.25% ± 2.05%, respectively. Adipose-derived stem cells (ADSCs)
Shen et al. fabricated porous SF/nHAp composites containing BMP- were used under chondrogenic (with TGF-β1 and IGF) and osteogenic
2 and SDF-1 loaded SF microspheres to synthesis the cell-free scaffold. (with dexamethasone) conditions in chondral and bony layers of the
Laminar jet break-up technology was used for SF microsphere pre- scaffold in order to evaluate the potential of these cells for differ-
paration which provides optimal condition for high encapsulation ef- entiation to cartilage and bone lineages. In the chondral layer which
ficiency (Shen et al., 2016). It was reported by Wenk et al. that pre- was previously seeded with ADSCs, chondrocyte-like cells were ob-
paring SF microspheres using this method can preserved the bioactivity served, which were enclosed by ECM components, especially collagen
of IGF-I and controlling the release rate of growth factor in a sustained type II and glycosaminoglycan (GAG). The amounts of ECM compo-
manner (Wenk et al., 2008). SDF-1 showed a rapid release rate during nents were also increased during 21 days in the chondral layer. More-
the first few days, while it presented a sustained profile in the following over, chondrogenic-related genes were also up-regulated during cul-
three weeks. The scaffold significantly provoked the osteogenic differ- ture. Generally, it was proposed that ADSCs seeded chondral layer
entiation and recruitment of BMSCs (Shen et al., 2016). provided suitable environments for inducing chondrogenesis under
In another study, electrospun SF nanofibers containing HAp were chondrogenic conditions. The ADSCs also showed higher expression of
fabricated for periosteum regeneration (Ding et al., 2016c). Periosteum osteogenic related genes and bone ECM components such as Col I and
is the main source of osteo-progenitor cells around bone, therefore, this calcium. Therefore, osteogenic differentiation also could be promoted
tissue plays a crucial role in the initial phase of bone formation and under osteogenic conditions. Consequently, the pores were filled with
regeneration (Tiyapatanaputi et al., 2004; Zhang et al., 2005a). The more cartilage-related and bone-related matrices as the culture time
HAp particles were seen on the surfaces of SF nanofibers containing increased, resulting in an increase in the compressive elastic modulus
10%, 20%, and 30% HAp. Moreover, more HAp was observed on the (Ding et al., 2014). The results showed promise for the use of novel 3D
scaffolds with higher content of HAp. Some aggregations of HAp par- integrated SF scaffolds in osteochondral tissue engineering in vitro.
ticles were observed on the surface of interconnected pores of SF/HAp However, whether ADSCs can differentiate into cartilage and bone in
nanofibers which may be responsible for increasing the diameter of the corresponding chondral and bony layers of integrated scaffolds
10%, 20%, and 30% SF/HAp nanofibers. The Young's modulus of the needs to be further investigated through in vivo implantation. The
scaffolds was also increased with the increase in the content of HAp. isolating role of the intermediate layer needs to be confirmed in vivo.
Moreover, the results demonstrated that SF scaffolds containing 30% Other important studies using SF/HAp for stem cell differentiation are
HAp induced the metabolic activity of rBMSCs more than those scaf- listed in Table 5.
folds containing 0% HAp after 4 days. SF scaffolds containing 30% HAp Totally, primary osteoblast cells are the mostly used cells for bone

85
Table 5
M. Farokhi et al.

Simultaneous use of SF/HAp composites and stem cell for bone tissue engineering.

Material Processing method Structure and properties of the scaffold Cell Key findings Ref.

SFa/HApb Direct-write assembly Formation of 3D scaffold with an interpenetrating gradient network of hMSCsc/ Formation of complex network of hMSCs matrix within the 3D scaffold ((Sun et al., 2012)
pore channels, suitable mechanical integrity hMMECsd by hMSCs and hMMECs based on histological analysis, observation of
vascular-like structures only in hMSCs and hMMECs co-culture groups
SF/HAp Freeze-drying/co- Formation of dense flake-like HAp crystals on the surface and the pore BMSCse Increasing the surface roughness and proliferation of BMSCs by (Jiang et al., 2013)
precipitation walls of the 3D silk structure incorporating HAp in SF matrix, improved in vitro differentiation of
BMSCs toward osteoblast cells
SF/nHAp Graft polymerization High and uniform coupling of nHAp on the surface of nano-HAp/SF MMCsf Good cellular proliferation and attachment, higher ALPg and bone- (Tanaka et al., 2007)
sheet specific osteocalcin production under osteogenic condition after
14 days of cell cultured on nano-HAp/SF sheets compared with controls
SF/HAp Salt-leaching Altered surface chemistry, increased surface roughness and stiffness hMSCs Higher surface osteoconductivity of SF scaffold by using HAp and (Bhumiratana et al.,
after incorporating HAp in the structure of 3D scaffold suggesting HAp microparticles as nucleation sites for directing 2011)
mineralization, enhancing the formation of trabecular structure,
increasing the mechanical behavior and conductivity of tissue grafts
SF/Gelatin/ Salt-leaching/ Formation of 3D scaffold with BMSCs/ Higher proliferation rate of mouse MC3T3-E1 cell line seeded on the (Sinlapabodin et al.,
HAp co-precipitation approximately 52 ± 5.5% porosity and pore size of 52.3 ± 5.5 MC3T3E1 scaffolds with perfusion flow rate of 1 mL/min than the static culture, 2016)
increased differentiation of BMSCs into osteoblast using perfusion flow
rate of 3 mL/min compared to 1 and 5 mL/min flow rates, statistically
higher Ca/P ratio, ALP activity, and calcium contents in flow rate of
3 mL/min compared to other flow rates
Ti/SF/nHAp Co-precipitation/ Induction of HAp nanocrystals aggregation in the presence of SF, intact MG-63 Better cellular attachment and osteogenic differentiation based on ALP (Lin et al., 2015)
electrostatic spray crystalline structure of HAp in the structure of SF/nHAp composites activity, BGPh contents, and Col Ii of cells in SF/nHAp compared with
deposition nHAp groups

86
SF/HAp Freeze-drying Formation of 3D structure with interconnected pores, varied pore size hDPCsj/ Remaining mostly SF scaffold after 4 weeks at the defected site, (Park et al., 2015)
between 20 and 80 μm with the diameter of 65 μm hPDLCsk supporting cellular attachment on SF scaffold and producing
extracellular matrix, inducing new bone formation in the edges of
defects by using SF fibers
SF/PHBVl/HAp Electrospinning Reduction in the tensile Young's modulus of the composite fibers with MC3T3-E1 Increased cellular proliferation and attachment on composites (Paşcu et al., 2016)
the increase in the content of the nHAp and SF phase from 2% to 5% compared with PHBV control, higher expansion and elongation of cells
on the surface of fiber composites after 1 day of culturing, incomplete
elongation of cells seeded on PHBV scaffolds
SF/HAp Salt-leaching – PMSCsm Secretion of calcium crystals by exposing PMSCs into osteogenic (Jin et al., 2014)
medium, secreting extracellular matrix when cell cultured on SF/HAp
scaffolds, increasing tissue regeneration by using PMSCs seeded SF/
HAp scaffolds in radius segmental bone of rabbit

a
Silk fibroin
b
Hydroxyapatite
c
Human bone marrow derived mesenchymal stem cells
d
Human mammary microvascular endothelial cells
e
Bone marrow derived mesenchymal stromal cells
f
Marrow mesenchymal cells
g
Alkaline phosphatase
h
Bone gla protein
i
Collagen type I
j
Human dental pulp cells
k
Human periodontal ligament cells
l
Polyhydroxybutyrate–polyhydroxyvalerate
m
Placenta-derived mesenchymal stem cells.
Biotechnology Advances 36 (2018) 68–91
M. Farokhi et al. Biotechnology Advances 36 (2018) 68–91

regeneration; however, their limited number and dedifferentiation References


properties might limit applications. Using potential stem cells can
overcome these drawbacks (Wang et al., 2006). Stem cells have been Abadi, M., Ghasemi, I., Khavandi, A., Shokrgozar, M., Farokhi, M., Homaeigohar, S.S.,
largely investigated for creating functional tissues and organs but it is Eslamifar, A., 2010. Synthesis of nano β-TCP and the effects on the mechanical and
biological properties of β-TCP/HDPE/UHMWPE nanocomposites. Polym. Compos. 31
still at infant stage. Bone marrow is the favored source of MSCs for both (10), 1745–1753.
experimental and clinical application. The biological features of stem Aboudzadeh, N., Imani, M., Shokrgozar, M.A., Khavandi, A., Javadpour, J., Shafieyan, Y.,
cell such as self-renewal and differentiation capacity are regulated by Farokhi, M., 2010. Fabrication and characterization of poly (D, L-Lactide-co-glyco-
lide)/hydroxyapatite nanocomposite scaffolds for bone tissue regeneration. J.
ECM components. ECM also serves as a supporting structure for various Biomed. Mater. Res. A 94 (1), 137–145.
growth factors and biomolecules which can act as structural support Addadi, L., Weiner, S., 1985. Interactions between acidic proteins and crystals: stereo-
with the ability to control signal transduction (Kundu and Kundu, chemical requirements in biomineralization. Proc. Natl. Acad. Sci. 82 (12),
4110–4114.
2010). The suitable morphological aspects of SF/HAp composites make Addison, W.N., Miller, S.J., Ramaswamy, J., Mansouri, A., Kohn, D.H., McKee, M.D.,
it an appropriate structure for stimulating the differentiation of stem 2010. Phosphorylation-dependent mineral-type specificity for apatite-binding pep-
cell into desired lineage and thus promoting the regeneration of de- tide sequences. Biomaterials 31 (36), 9422–9430.
Agarwal, S., Wendorff, J.H., Greiner, A., 2008. Use of electrospinning technique for
fected tissue.
biomedical applications. Polymer 49 (26), 5603–5621.
Andiappan, M., Sundaramoorthy, S., Panda, N., Meiyazhaban, G., Winfred, S.B.,
Venkataraman, G., Krishna, P., 2013. Electrospun eri silk fibroin scaffold coated with
5. Conclusions and future perspectives hydroxyapatite for bone tissue engineering applications. Prog. Biomater. 2 (1), 1–11.
Aoki, H., 1991. Science and Medical Applications of Hydroxyapatite. Ishiyaku
Silk scaffolds are biocompatible structures with tunable properties Euroamerica.
Arafat, M.T., Lam, C.X., Ekaputra, A.K., Wong, S.Y., He, C., Hutmacher, D.W., Li, X.,
in terms of morphology, degradability, and conformation that make
Gibson, I., 2011. High performance additive manufactured scaffolds for bone tissue
them useful candidates for tissue engineering applications. Recently, engineering application. Soft Matter 7 (18), 8013–8022.
investigations have been performed to better understand the structure Arcos, D., Vallet-Regí, M., 2013. Bioceramics for drug delivery. Acta Mater. 61 (3),
and processing of silk-based structures that broadened the applications 890–911.
Azadi, M., Teimouri, A., Mehranzadeh, G., 2016. Preparation, characterization and bio-
of different types of silk in regenerative medicine. The unique char- compatible properties of -chitin/silk fibroin/nanohydroxyapatite composite scaffolds
acteristics of silk such as exceptional mechanical and physical behavior, prepared by freeze-drying method. RSC Adv. 6, 7048–7060.
and versatile processability to form different structures make it poten- Baino, F., Vitale-Brovarone, C., 2014. Bioceramics in ophthalmology. Acta Biomater. 10
(8), 3372–3397.
tially useful for ligament, bone and musculoskeletal tissue engineering. Bajaj, A.K., Wongworawat, A.A., Punjabi, A., 2003. Management of alveolar clefts. J.
Silk has many suitable properties for various applications; however, it Craniofacial Surg. 14 (6), 840–846.
has some limitations. For example, the properties of silk protein, as a Balmayor, E.R., 2015. Targeted delivery as key for the success of small osteoinductive
molecules. Adv. Drug Deliv. Rev. 94, 13–27.
natural polymer, are different based on species source. Moreover, in- Bauer, T.W., Geesink, R., Zimmerman, R., McMahon, J.T., 1991. Hydroxyapatite-coated
consistence degumming processes and different fabrication methods femoral stems. Histological analysis of components retrieved at autopsy. JBJS 73
may produce silk structures with various properties if the process is not (10), 1439–1452.
Behera, S., Naskar, D., Sapru, S., Bhattacharjee, P., Dey, T., Ghosh, A.K., Mandal, M.,
tightly controlled. It is well established that genetically modified silk Kundu, S.C., 2017. Hydroxyapatite reinforced inherent RGD containing silk fibroin
proteins may overcome the mentioned limitations pending scale up and composite scaffolds: promising platform for bone tissue engineering. Nanomedicine.
cost efficiencies. In order to enhance the potential of silk-based scaf- Bhattacharjee, P., Naskar, D., Maiti, T.K., Bhattacharya, D., Das, P., Nandi, S.K., Kundu,
S.C., 2016a. Potential of non-mulberry silk protein fibroin blended and grafted poly
folds for bone tissue engineering applications, this polymer can be
(Є-caprolactone) nanofibrous matrices for in vivo bone regeneration. Colloids Surf. B:
blended with other materials such as HAp to enhance osteogenic po- Biointerfaces 143, 431–439.
tential of the scaffolds. HAp has a similar chemical structure to mi- Bhattacharjee, P., Naskar, D., Maiti, T.K., Bhattacharya, D., Kundu, S.C., 2016b. Non-
neralized bone of human tissue. This ceramic material demonstrates mulberry silk fibroin grafted poly (ε-caprolactone) nanofibrous scaffolds mineralized
by electrodeposition: an optimal delivery system for growth factors to enhance bone
affinity to hard tissues. More investigations are to be carried out in regeneration. RSC Adv. 6 (32), 26835–26855.
order to understand better the application of nanoscale constructs that Bhattacharjee, P., Naskar, D., Maiti, T.K., Bhattacharya, D., Kundu, S.C., 2016c. Non-
can be degraded over time and replaced by endogenous hard tissues. It mulberry silk fibroin grafted poly (Є-caprolactone)/nano hydroxyapatite nanofibrous
scaffold for dual growth factor delivery to promote bone regeneration. J. Colloid
is also necessary to focus on modifying the surfaces of biomaterials with Interface Sci. 472, 16–33.
different biomolecules with the aim of improving their biological be- Bhumiratana, S., Grayson, W.L., Castaneda, A., Rockwood, D.N., Gil, E.S., Kaplan, D.L.,
havior and interfacial functions. There are many studies that describe Vunjak-Novakovic, G., 2011. Nucleation and growth of mineralized bone matrix on
silk-hydroxyapatite composite scaffolds. Biomaterials 32 (11), 2812–2820.
the benefits of using nHAp-SF scaffolds for bone tissue engineering Bini, E., Knight, D.P., Kaplan, D.L., 2004. Mapping domain structures in silks from insects
applications in vitro and in vivo. However, there is no comprehensive and spiders related to protein assembly. J. Mol. Biol. 335 (1), 27–40.
study that shows the adverse effects of using these structures. It is es- Bose, S., Tarafder, S., 2012. Calcium phosphate ceramic systems in growth factor and
drug delivery for bone tissue engineering: a review. Acta Biomater. 8 (4), 1401–1421.
sential to develop effective constructs for clinical use despite advanta-
Boskey, A.L., 1989. Noncollagenous matrix proteins and their role in mineralization. Bone
geous properties of SF/HAp in this area. It is noteworthy that many Mineral 6 (2), 111–123.
investigations have confirmed the potential of SF scaffolds to induce Boskey, A.L., 1998. Biomineralization: conflicts, challenges, and opportunities. J. Cell.
Biochem. 72 (S30–31), 83–91.
osteogenesis in vivo. Most of these studies are performed on small an-
Boskey, A.L., 2013. Bone composition: relationship to bone fragility and antiosteoporotic
imal models, thus the results may not fully be extended for human use. drug effects. BoneKEy Rep. 2.
More investigations are needed to be performed in the future to identify Brannon-Peppas, L., Blanchette, J.O., 2012. Nanoparticle and targeted systems for cancer
the requirements of using silk-based scaffolds in clinical trials and therapy. Adv. Drug Deliv. Rev. 64, 206–212.
Cai, Y., Jin, J., Mei, D., Xia, N., Yao, J., 2009. Effect of silk sericin on assembly of hy-
constructing suitable commercialized structures for bone tissue re- droxyapatite nanocrystals into enamel prism-like structure. J. Mater. Chem. 19 (32),
generation. 5751–5758.
Cao, H., Chen, X., Yao, J., Shao, Z., 2013. Fabrication of an alternative regenerated silk
fibroin nanofiber and carbonated hydroxyapatite multilayered composite via layer-
Acknowledgments by-layer. J. Mater. Sci. 48 (1), 150–155.
Chen, F.-M., Zhang, M., Wu, Z.-F., 2010. Toward delivery of multiple growth factors in
tissue engineering. Biomaterials 31 (24), 6279–6308.
This work is supported by Pasteur Institute of Iran (grant number Chen, G., Deng, C., Li, Y.-P., 2012. TGF-beta and BMP signaling in osteoblast differ-
984). entiation and bone formation. Int. J. Biol. Sci. 8 (2), 272–288.
Chen, L., Gu, Y., Feng, Y., Zhu, X.-S., Wang, C.-Z., Liu, H.-L., Niu, H.-Y., Zhang, C., Yang,
H.-L., 2014a. Bioactivity of porous biphasic calcium phosphate enhanced by re-
Disclosure statement combinant human bone morphogenetic protein 2/silk fibroin microsphere. J. Mater.
Sci. Mater. Med. 25 (7), 1709–1719.
Chen, L., Hu, J., Ran, J., Shen, X., Tong, H., 2014b. Preparation and evaluation of
No potential conflict of interest was reported by the authors.

87
M. Farokhi et al. Biotechnology Advances 36 (2018) 68–91

collagen-silk fibroin/hydroxyapatite nanocomposites for bone tissue engineering. Int. Gemini-Piperni, S., Milani, R., Bertazzo, S., Peppelenbosch, M., Takamori, E.R., Granjeiro,
J. Biol. Macromol. 65, 1–7. J.M., Ferreira, C.V., Teti, A., Zambuzzi, W., 2014. Kinome profiling of osteoblasts on
Chen, L., Hu, J., Ran, J., Shen, X., Tong, H., 2015a. A novel nanocomposite for bone tissue hydroxyapatite opens new avenues on biomaterial cell signaling. Biotechnol. Bioeng.
engineering based on chitosan–silk sericin/hydroxyapatite: biomimetic synthesis and 111 (9), 1900–1905.
its cytocompatibility. RSC Adv. 5 (69), 56410–56422. Gholipourmalekabadi, M., Mozafari, M., Gholipourmalekabadi, M., Nazm Bojnordi, M.,
Chen, L., Liu, H.-L., Gu, Y., Feng, Y., Yang, H.-L., 2015b. Lumbar interbody fusion with Hashemi-soteh, M.B., Salimi, M., Rezaei, N., Sameni, M., Samadikuchaksaraei, A.,
porous biphasic calcium phosphate enhanced by recombinant bone morphogenetic Ghasemi Hamidabadi, H., 2015. In vitro and in vivo evaluations of three-dimensional
protein-2/silk fibroin sustained-released microsphere: an experimental study on hydroxyapatite/silk fibroin nanocomposite scaffolds. Biotechnol. Appl. Biochem. 62
sheep model. J. Mater. Sci. Mater. Med. 26 (3), 1–12. (4), 441–450.
Chouzouri, G., Xanthos, M., 2007. In vitro bioactivity and degradation of poly- Ghorbanian, L., Emadi, R., Razavi, S.M., Shin, H., Teimouri, A., 2013. Fabrication and
caprolactone composites containing silicate fillers. Acta Biomater. 3 (5), 745–756. characterization of novel diopside/silk fibroin nanocomposite scaffolds for potential
Clavero, J., Lundgren, S., 2003. Ramus or chin grafts for maxillary sinus inlay and local application in maxillofacial bone regeneration. Int. J. Biol. Macromol. 58, 275–280.
onlay augmentation: comparison of donor site morbidity and complications. Clin. Giannoudis, P.V., Dinopoulos, H., Tsiridis, E., 2005. Bone substitutes: an update. Injury 36
Implant. Dent. Relat. Res. 5 (3), 154–160. (3), S20–S27.
Cunniff, P.M., Fossey, S.A., Auerbach, M.A., Song, J.W., Kaplan, D.L., Adams, W.W., Eby, Ginebra, M.-P., Traykova, T., Planell, J., 2006a. Calcium phosphate cements as bone drug
R.K., Mahoney, D., Vezie, D.L., 1994. Mechanical and thermal properties of dragline delivery systems: a review. J. Control. Release 113 (2), 102–110.
silk from the spider Nephila clavipes. Polym. Adv. Technol. 5 (8), 401–410. Ginebra, M.-P., Traykova, T., Planell, J.A., 2006b. Calcium phosphate cements: compe-
Datta, A., Ghosh, A.K., Kundu, S.C., 2001. Differential expression of the fibroin gene in titive drug carriers for the musculoskeletal system? Biomaterials 27 (10), 2171–2177.
developmental stages of silkworm, Antheraea mylitta (Saturniidae). Comp. Biochem. Ginebra, M.-P., Canal, C., Espanol, M., Pastorino, D., Montufar, E.B., 2012. Calcium
Physiol. B: Biochem. Mol. Biol. 129 (1), 197–204. phosphate cements as drug delivery materials. Adv. Drug Deliv. Rev. 64 (12),
Ding, X., Zhu, M., Xu, B., Zhang, J., Zhao, Y., Ji, S., Wang, L., Wang, L., Li, X., Kong, D., 1090–1110.
2014. Integrated trilayered silk fibroin scaffold for osteochondral differentiation of Gosline, J., Guerette, P., Ortlepp, C., Savage, K., 1999. The mechanical design of spider
adipose-derived stem cells. ACS Appl. Mater. Interfaces 6 (19), 16696–16705. silks: from fibroin sequence to mechanical function. J. Exp. Biol. 202 (23),
Ding, Z., Fan, Z., Huang, X., Bai, S., Song, D., Lu, Q., Kaplan, D., 2016a. Bioactive natural 3295–3303.
protein–hydroxyapatite nanocarriers for optimizing osteogenic differentiation of Gupta, P., Adhikary, M., Kumar, M., Bhardwaj, N., Mandal, B.B., 2016. Biomimetic, os-
mesenchymal stem cells. J. Mater. Chem. B 4 (20), 3555–3561. teoconductive non-mulberry silk fiber reinforced tricomposite scaffolds for bone
Ding, Z., Fan, Z., Huang, X., Lu, Q., Xu, W., Kaplan, D.L., 2016b. Silk–hydroxyapatite tissue engineering. ACS Appl. Mater. Interfaces 8 (45), 30797–30810.
nanoscale scaffolds with programmable growth factor delivery for bone repair. ACS Hassan, M.N., Mahmoud, M.M., El-Fattah, A.A., Kandil, S., 2016. Microwave-assisted
Appl. Mater. Interfaces 8 (37), 24463–24470. preparation of Nano-hydroxyapatite for bone substitutes. Ceram. Int. 42 (3),
Ding, X., Wu, C., Ha, T., Wang, L., Huang, Y., Kang, H., Zhang, Y., Liu, H., Fan, Y., 2016c. 3725–3744.
Hydroxyapatite-containing silk fibroin nanofibrous scaffolds for tissue-engineered Hassani Besheli, N., Mottaghitalab, F., Eslami, M., Gholami, M., Kundu, S.C., Kaplan, D.L.,
periosteum. RSC Adv. 6 (23), 19463–19474. Farokhi, M., 2017. Sustainable release of vancomycin from silk fibroin nanoparticles
Ding, Z., Han, H., Fan, Z., Lu, H., Sang, Y., Yao, Y., Cheng, Q., Lu, Q., Kaplan, D.L., 2017. for treating severe bone infections in a rat tibia osteomyelitis model. ACS Appl.
Nanoscaled Silk-Hydroxyapatite Hydrogels for Injectable Bone Biomaterials. ACS Mater. Interfaces 9 (6), 5128–5138.
Appl. Mater. Interfaces 9 (20), 16913–16921. Hayden, R.S., Vollrath, M., Kaplan, D.L., 2014. Effects of clodronate and alendronate on
Dinjaski, N., Plowright, R., Zhou, S., Belton, D.J., Perry, C.C., Kaplan, D.L., 2017. osteoclast and osteoblast co-cultures on silk–hydroxyapatite films. Acta Biomater. 10
Osteoinductive recombinant silk fusion proteins for bone regeneration. Acta (1), 486–493.
Biomater. 49, 127–139. He, X., Huang, X., Lu, Q., Bai, S., Zhu, H., 2012a. Nanoscale control of silks for regular
Dorozhkin, S.V., 2010. Bioceramics of calcium orthophosphates. Biomaterials 31 (7), hydroxyapatite formation. Prog. Nat. Sci. 22 (2), 115–119.
1465–1485. He, P., Ng, K.S., Toh, S.L., Goh, J.C.H., 2012b. In vitro ligament–bone interface re-
Dorozhkin, S.V., 2012. Calcium Orthophosphates: Applications in Nature, Biology, and generation using a trilineage coculture system on a hybrid silk scaffold.
Medicine. CRC Press. Biomacromolecules 13 (9), 2692–2703.
Du, C., Jin, J., Li, Y., Kong, X., Wei, K., Yao, J., 2009. Novel silk fibroin/hydroxyapatite He, J., Wang, D., Cui, S., 2012c. Novel hydroxyapatite/tussah silk fibroin/chitosan bone-
composite films: structure and properties. Mater. Sci. Eng. C 29 (1), 62–68. like nanocomposites. Polym. Bull. 68 (6), 1765–1776.
Du, N., Yang, Z., Liu, X.Y., Li, Y., Xu, H.Y., 2011. Structural origin of the strain-hardening He, J., Cheng, Y., Li, P., Zhang, Y., Zhang, H., Cui, S., 2013. Preparation and character-
of spider silk. Adv. Funct. Mater. 21 (4), 772–778. ization of biomimetic tussah silk fibroin/chitosan composite nanofibers. Iran. Polym.
Ducheyne, P., Healy, K., Hutmacher, D.E., Grainger, D.W., Kirkpatrick, C.J., 2015. J. 22 (7), 537–547.
Comprehensive Biomaterials. Newnes. Hench, L., Jones, J., 2005. Biomaterials, Artificial Organs and Tissue Engineering.
El Briak-BenAbdeslam, H., Ginebra, M., Vert, M., Boudeville, P., 2008. Wet or dry me- Elsevier.
chanochemical synthesis of calcium phosphates? Influence of the water content on Hilton, M.J., Tu, X., Wu, X., Bai, S., Zhao, H., Kobayashi, T., Kronenberg, H.M.,
DCPD–CaO reaction kinetics. Acta Biomater. 4 (2), 378–386. Teitelbaum, S.L., Ross, F.P., Kopan, R., 2008. Notch signaling maintains bone marrow
Eshtiagh-Hosseini, H., Housaindokht, M.R., Chahkandi, M., 2007. Effects of parameters of mesenchymal progenitors by suppressing osteoblast differentiation. Nat. Med. 14 (3),
sol–gel process on the phase evolution of sol–gel-derived hydroxyapatite. Mater. 306–314.
Chem. Phys. 106 (2), 310–316. Hu, J.-x., Cai, X., Mo, S.-b., Chen, L., Shen, X.-y., 2015. Fabrication and characterization
Faria, R.M., César, D.V., Salim, V.M., 2008. Surface reactivity of zinc-modified hydro- of chitosan-silk fibroin/hydroxyapatite composites via in situ precipitation for bone
xyapatite. Catal. Today 133, 168–173. tissue engineering. Chin. J. Polym. Sci. 33 (12), 1661–1671.
Farokhi, M., Mottaghitalab, F., Hadjati, J., Omidvar, R., Majidi, M., Amanzadeh, A., Huang, X., Liu, X., Liu, S., Zhang, A., Lu, Q., Kaplan, D.L., Zhu, H., 2014.
Azami, M., Tavangar, S.M., Shokrgozar, M.A., Ai, J., 2014a. Structural and functional Biomineralization regulation by nano-sized features in silk fibroin proteins: synthesis
changes of silk fibroin scaffold due to hydrolytic degradation. J. Appl. Polym. Sci. of water-dispersible nano-hydroxyapatite. J. Biomed. Mater. Res. B Appl. Biomater.
131 (6). 102 (8), 1720–1729.
Farokhi, M., Mottaghitalab, F., Shokrgozar, M.A., Ai, J., Hadjati, J., Azami, M., 2014b. Idris, A.I., Rojas, J., Greig, I.R., van't Hof, R.J., Ralston, S.H., 2008.
Bio-hybrid silk fibroin/calcium phosphate/PLGA nanocomposite scaffold to control Aminobisphosphonates cause osteoblast apoptosis and inhibit bone nodule formation
the delivery of vascular endothelial growth factor. Mater. Sci. Eng. C 35, 401–410. in vitro. Calcif. Tissue Int. 82 (3), 191–201.
Farokhi, M., Mottaghitalab, F., Shokrgozar, M.A., Kaplan, D.L., Kim, H.-W., Kundu, S.C., Ito, Y., Hasuda, H., Kamitakahara, M., Ohtsuki, C., Tanihara, M., Kang, I.-K., Kwon, O.H.,
2016a. Prospects of peripheral nerve tissue engineering using nerve guide conduits 2005. A composite of hydroxyapatite with electrospun biodegradable nanofibers as a
based on silk fibroin protein and other biopolymers. Int. Mater. Rev. 1–25. tissue engineering material. J. Biosci. Bioeng. 100 (1), 43–49.
Farokhi, M., Mottaghitalab, F., Shokrgozar, M.A., Ou, K.-L., Mao, C., Hosseinkhani, H., Jiang, J., Hao, W., Li, Y., Yao, J., Shao, Z., Li, H., Yang, J., Chen, S., 2013.
2016b. Importance of dual delivery systems for bone tissue engineering. J. Control. Hydroxyapatite/regenerated silk fibroin scaffold-enhanced osteoinductivity and os-
Release 225, 152–169. teoconductivity of bone marrow-derived mesenchymal stromal cells. Biotechnol. Lett.
Fathi, M., Hanifi, A., Mortazavi, V., 2008. Preparation and bioactivity evaluation of bone- 35 (4), 657–661.
like hydroxyapatite nanopowder. J. Mater. Process. Technol. 202 (1), 536–542. Jiang, T., Carbone, E.J., Lo, K.W.-H., Laurencin, C.T., 2015. Electrospinning of polymer
Figueiredo, M., Fernando, A., Martins, G., Freitas, J., Judas, F., Figueiredo, H., 2010. nanofibers for tissue regeneration. Prog. Polym. Sci. 46, 1–24.
Effect of the calcination temperature on the composition and microstructure of hy- Jin, J., Wang, J., Huang, J., Huang, F., Fu, J., Yang, X., Miao, Z., 2014. Transplantation of
droxyapatite derived from human and animal bone. Ceram. Int. 36 (8), 2383–2393. human placenta-derived mesenchymal stem cells in a silk fibroin/hydroxyapatite
Fu, Q., Saiz, E., Rahaman, M.N., Tomsia, A.P., 2011. Bioactive glass scaffolds for bone scaffold improves bone repair in rabbits. J. Biosci. Bioeng. 118 (5), 593–598.
tissue engineering: state of the art and future perspectives. Mater. Sci. Eng. C 31 (7), Jo, Y.-Y., Kim, S.-G., Kwon, K.-J., Kweon, H., Chae, W.-S., Yang, W.-G., Lee, E.-Y., Seok,
1245–1256. H., 2017. Silk fibroin-alginate-hydroxyapatite composite particles in bone tissue en-
Fuchs, S., Jiang, X., Schmidt, H., Dohle, E., Ghanaati, S., Orth, C., Hofmann, A., Motta, A., gineering applications in vivo. Int. J. Mol. Sci. 18 (4), 858.
Migliaresi, C., Kirkpatrick, C.J., 2009. Dynamic processes involved in the pre-vas- Jung, S.-R., Song, N.-J., Yang, D.K., Cho, Y.-J., Kim, B.-J., Hong, J.-W., Yun, U.J., Jo, D.-
cularization of silk fibroin constructs for bone regeneration using outgrowth en- G., Lee, Y.M., Choi, S.Y., 2013. Silk proteins stimulate osteoblast differentiation by
dothelial cells. Biomaterials 30 (7), 1329–1338. suppressing the Notch signaling pathway in mesenchymal stem cells. Nutr. Res. 33
Fujisawa, R., Wada, Y., Nodasaka, Y., Kuboki, Y., 1996. Acidic amino acid-rich sequences (2), 162–170.
as binding sites of osteonectin to hydroxyapatite crystals. Biochim. Biophys. Acta Kaplan, D., McGrath, K., 2012. Protein-based materials. Springer Sci. Bus. Media.
1292 (1), 53–60. Kaplan, D., Adams, W.W., Farmer, B., Viney, C., 1994. Silk polymers(materials science
Gedanken, A., 2004. Using sonochemistry for the fabrication of nanomaterials. Ultrason. and biotechnology). In: A. C. S. Symposium Series. American Chemical Society.
Sonochem. 11 (2), 47–55. Karageorgiou, V., Meinel, L., Hofmann, S., Malhotra, A., Volloch, V., Kaplan, D., 2004.

88
M. Farokhi et al. Biotechnology Advances 36 (2018) 68–91

Bone morphogenetic protein-2 decorated silk fibroin films induce osteogenic differ- of HA/PMMA bone cement. Colloids Surf. B: Biointerfaces 131, 39–46.
entiation of human bone marrow stromal cells. J. Biomed. Mater. Res. A 71 (3), Liu, H., Xu, G.W., Wang, Y.F., Zhao, H.S., Xiong, S., Wu, Y., Heng, B.C., An, C.R., Zhu,
528–537. G.H., Xie, D.H., 2015b. Composite scaffolds of nano-hydroxyapatite and silk fibroin
Kilian, O., Wenisch, S., Karnati, S., Baumgart-Vogt, E., Hild, A., Fuhrmann, R., Jonuleit, enhance mesenchymal stem cell-based bone regeneration via the interleukin 1 alpha
T., Dingeldein, E., Schnettler, R., Franke, R.-P., 2008. Observations on the micro- autocrine/paracrine signaling loop. Biomaterials 49, 103–112.
vasculature of bone defects filled with biodegradable nanoparticulate hydro- Loi, F., Córdova, L.A., Zhang, R., Pajarinen, J., Lin, T.-h., Goodman, S.B., Yao, Z., 2016.
xyapatite. Biomaterials 29 (24), 3429–3437. The effects of immunomodulation by macrophage subsets on osteogenesis in vitro.
Kim, H.W., 2007. Biomedical nanocomposites of hydroxyapatite/polycaprolactone ob- Stem Cell Res Ther 7 (1), 15.
tained by surfactant mediation. J. Biomed. Mater. Res. A 83( (1), 169–177. Lü, X., Wang, J., Li, B., Zhang, Z., Zhao, L., 2014. Gene expression profile study on os-
Kim, Y.-H., Tabata, Y., 2015. Dual-controlled release system of drugs for bone re- teoinductive effect of natural hydroxyapatite. J. Biomed. Mater. Res. A 102 (8),
generation. Adv. Drug Deliv. Rev. 94, 28–40. 2833–2841.
Kim, H.J., Kim, U.-J., Kim, H.S., Li, C., Wada, M., Leisk, G.G., Kaplan, D.L., 2008. Bone Lu, Z., Roohani-Esfahani, S.-I., Li, J., Zreiqat, H., 2015. Synergistic effect of nanomaterials
tissue engineering with premineralized silk scaffolds. Bone 42 (6), 1226–1234. and BMP-2 signalling in inducing osteogenic differentiation of adipose tissue-derived
Kim, D.-H., Provenzano, P.P., Smith, C.L., Levchenko, A., 2012a. Matrix nanotopography mesenchymal stem cells. Nanomedicine 11 (1), 219–228.
as a regulator of cell function. J. Cell Biol. 197 (3), 351–360. Lucas, F., 1964. Spiders and their silks. Discovery 25 (1), 20–26.
Kim, D.-H., Smith, R.R., Kim, P., Ahn, E.H., Kim, H.-N., Marbán, E., Suh, K.-Y., Levchenko, Ma, Z., Kotaki, M., Inai, R., Ramakrishna, S., 2005. Potential of nanofiber matrix as tissue-
A., 2012b. Nanopatterned cardiac cell patches promote stem cell niche formation and engineering scaffolds. Tissue Eng. 11 (1–2), 101–109.
myocardial regeneration. Integr. Biol. 4 (9), 1019–1033. Mandal, B.B., Kundu, S.C., 2008. Non-bioengineered silk gland fibroin protein: char-
Kim, E.-S., Ahn, E.H., Dvir, T., Kim, D.-H., 2014a. Emerging nanotechnology approaches acterization and evaluation of matrices for potential tissue engineering applications.
in tissue engineering and regenerative medicine. Int. J. Nanomedicine 9 (Suppl. 1), 1. Biotechnol. Bioeng. 100 (6), 1237–1250.
Kim, H., Che, L., Ha, Y., Ryu, W., 2014b. Mechanically-reinforced electrospun composite Mandal, B.B., Kundu, S.C., 2009. Cell proliferation and migration in silk fibroin 3D
silk fibroin nanofibers containing hydroxyapatite nanoparticles. Mater. Sci. Eng. C scaffolds. Biomaterials 30 (15), 2956–2965.
40, 324–335. Mandal, B.B., Grinberg, A., Gil, E.S., Panilaitis, B., Kaplan, D.L., 2012. High-strength silk
Kim, H.H., Park, J.B., Kang, M.J., Park, Y.H., 2014c. Surface-modified silk hydrogel protein scaffolds for bone repair. Proc. Natl. Acad. Sci. 109 (20), 7699–7704.
containing hydroxyapatite nanoparticle with hyaluronic acid–dopamine conjugate. Maquet, V., Boccaccini, A.R., Pravata, L., Notingher, I., Jérôme, R., 2004. Porous poly (α-
Int. J. Biol. Macromol. 70, 516–522. hydroxyacid)/Bioglass® composite scaffolds for bone tissue engineering. I: prepara-
Kokubo, T., 1991. Bioactive glass ceramics: properties and applications. Biomaterials 12 tion and in vitro characterisation. Biomaterials 25 (18), 4185–4194.
(2), 155–163. Marelli, B., Ghezzi, C.E., Alessandrino, A., Barralet, J.E., Freddi, G., Nazhat, S.N., 2012.
Kokubo, T., 2008. Bioceramics and their Clinical Applications. Elsevier. Silk fibroin derived polypeptide-induced biomineralization of collagen. Biomaterials
Kolambkar, Y.M., Boerckel, J.D., Dupont, K.M., Bajin, M., Huebsch, N., Mooney, D.J., 33 (1), 102–108.
Hutmacher, D.W., Guldberg, R.E., 2011. Spatiotemporal delivery of bone morpho- Mauney, J.R., Nguyen, T., Gillen, K., Kirker-Head, C., Gimble, J.M., Kaplan, D.L., 2007.
genetic protein enhances functional repair of segmental bone defects. Bone 49 (3), Engineering adipose-like tissue in vitro and in vivo utilizing human bone marrow and
485–492. adipose-derived mesenchymal stem cells with silk fibroin 3D scaffolds. Biomaterials
Kolmas, J., Krukowski, S., Laskus, A., Jurkitewicz, M., 2016. Synthetic hydroxyapatite in 28 (35), 5280–5290.
pharmaceutical applications. Ceram. Int. 42(2, 2472–2487. Meinel, L., Fajardo, R., Hofmann, S., Langer, R., Chen, J., Snyder, B., Vunjak-Novakovic,
Kong, X., Cui, F., Wang, X., Zhang, M., Zhang, W., 2004. Silk fibroin regulated miner- G., Kaplan, D., 2005a. Silk implants for the healing of critical size bone defects. Bone
alization of hydroxyapatite nanocrystals. J. Cryst. Growth 270 (1), 197–202. 37 (5), 688–698.
Kundu, B., Kundu, S.C., 2010. Osteogenesis of human stem cells in silk biomaterial for Meinel, L., Hofmann, S., Karageorgiou, V., Kirker-Head, C., McCool, J., Gronowicz, G.,
regenerative therapy. Prog. Polym. Sci. 35 (9), 1116–1127. Zichner, L., Langer, R., Vunjak-Novakovic, G., Kaplan, D.L., 2005b. The inflammatory
Kundu, J., Chung, Y.-I., Kim, Y.H., Tae, G., Kundu, S., 2010. Silk fibroin nanoparticles for responses to silk films in vitro and in vivo. Biomaterials 26 (2), 147–155.
cellular uptake and control release. Int. J. Pharm. 388 (1), 242–250. Mi, R., Liu, Y., Chen, X., Shao, Z., 2016. Structure and properties of various hybrids
Kundu, B., Rajkhowa, R., Kundu, S.C., Wang, X., 2013. Silk fibroin biomaterials for tissue fabricated by silk nanofibrils and nanohydroxyapatite. Nano 8 (48), 20096–20102.
regenerations. Adv. Drug Deliv. Rev. 65 (4), 457–470. Midha, S., Murab, S., Ghosh, S., 2016. Osteogenic signaling on silk-based matrices.
Kweon, H., Lee, K.-G., Chae, C.-H., Balázsi, C., Min, S.-K., Kim, J.-Y., Choi, J.-Y., Kim, S.- Biomaterials 97, 133–153.
G., 2011. Development of nano-hydroxyapatite graft with silk fibroin scaffold as a Mimori, K., Komaki, M., Iwasaki, K., Ishikawa, I., 2007. Extracellular signal-regulated
new bone substitute. J. Oral Maxillofac. Surg. 69 (6), 1578–1586. kinase 1/2 is involved in ascorbic acid-induced osteoblastic differentiation in peri-
Langenbach, F., Handschel, J., 2013. Effects of dexamethasone, ascorbic acid and β- odontal ligament cells. J. Periodontol. 78 (2), 328–334.
glycerophosphate on the osteogenic differentiation of stem cells in vitro. Stem Cell Ming, J., Bie, S., Jiang, Z., Wang, P., Zuo, B., 2014a. Novel hydroxyapatite nanorods
Res Ther 4 (5), 117. crystal growth in silk fibroin/sodium alginate nanofiber hydrogel. Mater. Lett. 126,
Lee, K.Y., Yuk, S.H., 2007. Polymeric protein delivery systems. Prog. Polym. Sci. 32 (7), 169–173.
669–697. Ming, J., Liu, Z., Bie, S., Zhang, F., Zuo, B., 2014b. Novel silk fibroin films prepared by
LeGeros, R.Z., 1990. Calcium phosphates in oral biology and medicine. Monogr. Oral Sci. formic acid/hydroxyapatite dissolution method. Mater. Sci. Eng. C 37, 48–53.
15, 1–201. Miroiu, F., Socol, G., Visan, A., Stefan, N., Craciun, D., Craciun, V., Dorcioman, G.,
LeGeros, R.Z., 1993. Biodegradation and bioresorption of calcium phosphate ceramics. Mihailescu, I., Sima, L., Petrescu, S., 2010. Composite biocompatible hydro-
Clin. Mater. 14 (1), 65–88. xyapatite–silk fibroin coatings for medical implants obtained by matrix assisted
Li, C., Vepari, C., Jin, H.-J., Kim, H.J., Kaplan, D.L., 2006. Electrospun silk-BMP-2 scaf- pulsed laser evaporation. Mater. Sci. Eng. B 169 (1), 151–158.
folds for bone tissue engineering. Biomaterials 27 (16), 3115–3124. Miyamoto, S., Teramoto, H., Gutkind, J.S., Yamada, K.M., 1996. Integrins can collaborate
Li, L., Wei, K.-M., Lin, F., Kong, X.-D., Yao, J.-M., 2008. Effect of silicon on the formation with growth factors for phosphorylation of receptor tyrosine kinases and MAP kinase
of silk fibroin/calcium phosphate composite. J. Mater. Sci. Mater. Med. 19 (2), activation: roles of integrin aggregation and occupancy of receptors. J. Cell Biol. 135
577–582. (6), 1633–1642.
Li, R., Chen, G.M., Ma, X.L., Chen, Q.Y., Xu, G.W., Huang, Y.P., 2011. Mineralization of Mottaghitalab, F., Farokhi, M., Zaminy, A., Kokabi, M., Soleimani, M., Mirahmadi, F.,
HA crystals regulated by terephthaloyl chloride-modified silk fibroin films. Chin. Shokrgozar, M.A., Sadeghizadeh, M., 2013. A biosynthetic nerve guide conduit based
Chem. Lett. 22 (9), 1107–1110. on silk/SWNT/fibronectin nanocomposite for peripheral nerve regeneration. PLoS
Li, G., Liu, H., Li, T., Wang, J., 2012. Surface modification and functionalization of silk One 8 (9), e74417.
fibroin fibers/fabric toward high performance applications. Mater. Sci. Eng. C 32 (4), Mottaghitalab, F., Farokhi, M., Shokrgozar, M.A., Atyabi, F., Hosseinkhani, H., 2015a.
627–636. Silk fibroin nanoparticle as a novel drug delivery system. J. Control. Release 206,
Liang, D., Hsiao, B.S., Chu, B., 2007. Functional electrospun nanofibrous scaffolds for 161–176.
biomedical applications. Adv. Drug Deliv. Rev. 59 (14), 1392–1412. Mottaghitalab, F., Hosseinkhani, H., Shokrgozar, M.A., Mao, C., Yang, M., Farokhi, M.,
Lin, L., Hao, R., Xiong, W., Zhong, J., 2015. Quantitative analyses of the effect of silk 2015b. Silk as a potential candidate for bone tissue engineering. J. Control. Release
fibroin/nano-hydroxyapatite composites on osteogenic differentiation of MG-63 215, 112–128.
human osteosarcoma cells. J. Biosci. Bioeng. 119 (5), 591–595. Mottaghitalab, F., Kiani, M., Farokhi, M., Kundu, S.C., Reis, R.L., Gholami, M., Bardania,
Liu, L., Liu, J., Kong, X., Cai, Y., Yao, J., 2011. Porous composite scaffolds of hydro- H., Dinarvand, R., Geramifar, P., Beiki, D., 2017. Targeted delivery system based on
xyapatite/silk fibroin via two-step method. Polym. Adv. Technol. 22 (6), 909–914. gemcitabine loaded silk fibroin nanoparticles for lung cancer therapy. ACS Appl.
Liu, J., Liu, Y., Kong, Y., Yao, J., Cai, Y., 2013a. Formation of vaterite regulated by silk Mater. Interfaces 9 (37), 31600–31611.
sericin and its transformation towards hydroxyapatite microsphere. Mater. Lett. 110, Mou, Z.-L., Duan, L.-M., Qi, X.-N., Zhang, Z.-Q., 2013. Preparation of silk fibroin/col-
221–224. lagen/hydroxyapatite composite scaffold by particulate leaching method. Mater. Lett.
Liu, H., Peng, H., Wu, Y., Zhang, C., Cai, Y., Xu, G., Li, Q., Chen, X., Ji, J., Zhang, Y., 105, 189–191.
2013b. The promotion of bone regeneration by nanofibrous hydroxyapatite/chitosan Murugan, R., Ramakrishna, S., 2005. Aqueous mediated synthesis of bioresorbable na-
scaffolds by effects on integrin-BMP/Smad signaling pathway in BMSCs. Biomaterials nocrystalline hydroxyapatite. J. Cryst. Growth 274 (1), 209–213.
34 (18), 4404–4417. Nageeb, M., Nouh, S.R., Bergman, K., Nagy, N.B., Khamis, D., Kisiel, M., Engstrand, T.,
Liu, T., Ding, X., Lai, D., Chen, Y., Zhang, R., Chen, J., Feng, X., Chen, X., Yang, X., Zhao, Hilborn, J., Marei, M.K., 2012. Bone engineering by biomimetic injectable hydrogel.
R., 2014. Enhancing in vitro bioactivity and in vivo osteogenesis of organic–inorganic Mol. Cryst. Liq. Cryst. 555 (1), 177–188.
nanofibrous biocomposites with novel bioceramics. J. Mater. Chem. B 2 (37), Nazarov, R., Jin, H.-J., Kaplan, D.L., 2004. Porous 3-D scaffolds from regenerated silk
6293–6305. fibroin. Biomacromolecules 5 (3), 718–726.
Liu, Z., Tang, Y., Kang, T., Rao, M., Li, K., Wang, Q., Quan, C., Zhang, C., Jiang, Q., Shen, Nemoto, R., Nakamura, S., Isobe, T., Senna, M., 2001. Direct synthesis of hydroxyapatite-
H., 2015a. Synergistic effect of HA and BMP-2 mimicking peptide on the bioactivity silk fibroin nano-composite sol via a mechanochemical route. J. Sol-Gel Sci. Technol.

89
M. Farokhi et al. Biotechnology Advances 36 (2018) 68–91

21 (1–2), 7–12. glycolic acid)/graphene oxide/Tussah silk fibroin nanofiber scaffolds with multiple
Nikbakht Dastjerdi, M., 2006. Induction of mineralized nodule formation in rat bone orthogonal layers enhance osteoblastic differentiation of mesenchymal stem cells.
marrow stromal cell cultures by silk fibroin. Iran. Biomed. J. 10 (3), 133–138. ACS Biomater. Sci. Eng. 3 (7), 1370–1380.
Niu, B., Li, B., Gu, Y., Shen, X., Liu, Y., Chen, L., 2017. In vitro evaluation of electrospun Sheikh, F.A., Ju, H.W., Moon, B.M., Park, H.J., Kim, J.H., Lee, O.J., Park, C.H., 2013. A
silk fibroin/nano-hydroxyapatite/BMP-2 scaffolds for bone regeneration. J. Biomater. novel approach to fabricate silk nanofibers containing hydroxyapatite nanoparticles
Sci. Polym. Ed. 28 (3), 257–270. using a three-way stopcock connector. Nanoscale Res. Lett. 8 (1), 1.
Nori, S., Okada, Y., Yasuda, A., Tsuji, O., Takahashi, Y., Kobayashi, Y., Fujiyoshi, K., Sheikh, F.A., Woo Ju, H., Mi Moon, B., Jung Park, H., Kim, J.H., Joo Lee, O., Hum Park,
Koike, M., Uchiyama, Y., Ikeda, E., 2011. Grafted human-induced pluripotent stem- C., 2014. Facile and highly efficient approach for the fabrication of multifunctional
cell–derived neurospheres promote motor functional recovery after spinal cord injury silk nanofibers containing hydroxyapatite and silver nanoparticles. J. Biomed. Mater.
in mice. Proc. Natl. Acad. Sci. 108 (40), 16825–16830. Res. A 102 (10), 3459–3469.
Nourmohammadi, J., Roshanfar, F., Farokhi, M., Nazarpak, M.H., 2017. Silk fibroin/ Shen, X., Zhang, Y., Gu, Y., Xu, Y., Liu, Y., Li, B., Chen, L., 2016. Sequential and sustained
kappa-carrageenan composite scaffolds with enhanced biomimetic mineralization for release of SDF-1 and BMP-2 from silk fibroin-nanohydroxyapatite scaffold for the
bone regeneration applications. Mater. Sci. Eng. C 76, 951–958. enhancement of bone regeneration. Biomaterials 106, 205–216.
Nova, A., Keten, S., Pugno, N.M., Redaelli, A., Buehler, M.J., 2010. Molecular and na- Shi, P., Abbah, S.A., Saran, K., Zhang, Y., Li, J., Wong, H.-K., Goh, J.C., 2013a. Silk fi-
nostructural mechanisms of deformation, strength and toughness of spider silk fibrils. broin-based complex particles with bioactive encrustation for bone morphogenetic
Nano Lett. 10 (7), 2626–2634. protein 2 delivery. Biomacromolecules 14 (12), 4465–4474.
Omenetto, F.G., Kaplan, D.L., 2010. New opportunities for an ancient material. Science Shi, P., Chen, K., Goh, J.C., 2013b. Efficacy of BMP-2 delivery from natural protein based
329 (5991), 528–531. polymeric particles. Adv. Healthc. Mater. 2 (7), 934–939.
Panda, N., Bissoyi, A., Pramanik, K., Biswas, A., 2014. Directing osteogenesis of stem cells Shokrgozar, M., Farokhi, M., Rajaei, F., Bagheri, M., Azari, S., Ghasemi, I., Mottaghitalab,
with hydroxyapatite precipitated electrospun eri–tasar silk fibroin nanofibrous scaf- F., Azadmanesh, K., Radfar, J., 2010. Biocompatibility evaluation of HDPE-UHMWPE
fold. J. Biomater. Sci. Polym. Ed. 25 (13), 1440–1457. reinforced β-TCP nanocomposites using highly purified human osteoblast cells. J.
Park, J., 2009. Bioceramics: properties, characterizations, and applications. Springer Sci. Biomed. Mater. Res. A 95( (4), 1074–1083.
Bus. Media 741. Singh, B., Panda, N., Mund, R., Pramanik, K., 2016. Carboxymethyl cellulose enables silk
Park, J., Lakes, R.S., 2007. Biomaterials: an introduction. Springer Sci. Bus. Media. fibroin nanofibrous scaffold with enhanced biomimetic potential for bone tissue en-
Park, J.-Y., Yang, C., Jung, I.-H., Lim, H.-C., Lee, J.-S., Jung, U.-W., Seo, Y.-K., Park, J.-K., gineering application. Carbohydr. Polym. 151, 335–347.
Choi, S.-H., 2015. Regeneration of rabbit calvarial defects using cells-implanted Sinlapabodin, S., Amornsudthiwat, P., Damrongsakkul, S., Kanokpanont, S., 2016. An
nano-hydroxyapatite coated silk scaffolds. Biomater. Res. 19 (1), 1. axial distribution of seeding, proliferation, and osteogenic differentiation of MC3T3-
Paşcu, E.I., Cahill, P.A., Stokes, J., McGuinness, G.B., 2016. Towards functional 3D- E1 cells and rat bone marrow-derived mesenchymal stem cells across a 3D Thai silk
stacked electrospun composite scaffolds of PHBV, silk fibroin and nanohydrox- fibroin/gelatin/hydroxyapatite scaffold in a perfusion bioreactor. Mater. Sci. Eng. C
yapatite: mechanical properties and surface osteogenic differentiation. J. Biomater. 58, 960–970.
Appl (0885328215626047). Sofia, S., McCarthy, M.B., Gronowicz, G., Kaplan, D.L., 2001. Functionalized silk-based
Pérez-Rigueiro, J., Elices, M., Llorca, J., Viney, C., 2001. Tensile properties of silkworm biomaterials for bone formation. J. Biomed. Mater. Res. 54 (1), 139–148.
silk obtained by forced silking. J. Appl. Polym. Sci. 82 (8), 1928–1935. Song, J.-H., Kim, H.-E., Kim, H.-W., 2008a. Electrospun fibrous web of collagen–apatite
Qi, X.N., Mou, Z.L., Zhang, J., Zhang, Z.Q., 2014. Preparation of chitosan/silk fibroin/ precipitated nanocomposite for bone regeneration. J. Mater. Sci. Mater. Med. 19 (8),
hydroxyapatite porous scaffold and its characteristics in comparison to bi-component 2925–2932.
scaffolds. J. Biomed. Mater. Res. A 102 (2), 366–372. Song, J.-H., Kim, J.-H., Park, S., Kang, W., Kim, H.-W., Kim, H.-E., Jang, J.-H., 2008b.
Ratner, B.D., Hoffman, A.S., Schoen, F.J., Lemons, J.E., 2004. Biomaterials Science: An Signaling responses of osteoblast cells to hydroxyapatite: the activation of ERK and
Introduction to Materials in Medicine. Academic press. SOX9. J. Bone Miner. Metab. 26 (2), 138–142.
Ren, F., Xin, R., Ge, X., Leng, Y., 2009. Characterization and structural analysis of zinc- Song, H., Yoon, C., Kattman, S.J., Dengler, J., Massé, S., Thavaratnam, T., Gewarges, M.,
substituted hydroxyapatites. Acta Biomater. 5 (8), 3141–3149. Nanthakumar, K., Rubart, M., Keller, G.M., 2010. Interrogating functional integration
Rezwan, K., Chen, Q., Blaker, J., Boccaccini, A.R., 2006. Biodegradable and bioactive between injected pluripotent stem cell-derived cells and surrogate cardiac tissue.
porous polymer/inorganic composite scaffolds for bone tissue engineering. Proc. Natl. Acad. Sci. 107 (8), 3329–3334.
Biomaterials 27 (18), 3413–3431. Stupp, S.I., Ciegler, G.W., 1992. Organoapatites: materials for artificial bone. I. Synthesis
Rho, J.-Y., Kuhn-Spearing, L., Zioupos, P., 1998. Mechanical properties and the hier- and microstructure. J. Biomed. Mater. Res. 26 (2), 169–183.
archical structure of bone. Med. Eng. Phys. 20 (2), 92–102. Sugihara, A., Sugiura, K., Morita, H., Ninagawa, T., Tubouchi, K., Tobe, R., Izumiya, M.,
Ribeiro, M., Ferraz, M.P., Monteiro, F.J., Fernandes, M.H., Beppu, M.M., Mantione, D., Horio, T., Abraham, N.G., Ikehara, S., 2000. Promotive effects of a silk film on epi-
Sardon, H., 2017. Antibacterial silk fibroin/nanohydroxyapatite hydrogels with silver dermal recovery from full-thickness skin wounds. Exp. Biol. Med. 225 (1), 58–64.
and gold nanoparticles for bone regeneration. Nanomedicine 13 (1), 231–239. Sun, F., Zhou, H., Lee, J., 2011. Various preparation methods of highly porous hydro-
Sadat-Shojai, M., Khorasani, M.-T., Dinpanah-Khoshdargi, E., Jamshidi, A., 2013. xyapatite/polymer nanoscale biocomposites for bone regeneration. Acta Biomater. 7
Synthesis methods for nanosized hydroxyapatite with diverse structures. Acta (11), 3813–3828.
Biomater. 9 (8), 7591–7621. Sun, L., Parker, S.T., Syoji, D., Wang, X., Lewis, J.A., Kaplan, D.L., 2012. Direct-write
Sangkert, S., Kamonmattayakul, S., Chai, W.L., Meesane, J., 2016. A biofunctional- assembly of 3D silk/hydroxyapatite scaffolds for bone co-cultures. Adv. Healthc.
modified silk fibroin scaffold with mimic reconstructed extracellular matrix of de- Mater. 1 (6), 729–735.
cellularized pulp/collagen/fibronectin for bone tissue engineering in alveolar bone Šupová, M., 2015. Substituted hydroxyapatites for biomedical applications: a review.
resorption. Mater. Lett. 166, 30–34. Ceram. Int. 41 (8), 9203–9231.
Sasikumar, S., Vijayaraghavan, R., 2008. Solution combustion synthesis of bioceramic Sutherland, T.D., Young, J.H., Weisman, S., Hayashi, C.Y., Merritt, D.J., 2010. Insect silk:
calcium phosphates by single and mixed fuels—a comparative study. Ceram. Int. 34 one name, many materials. Annu. Rev. Entomol. 55, 171–188.
(6), 1373–1379. Swetha, M., Sahithi, K., Moorthi, A., Srinivasan, N., Ramasamy, K., Selvamurugan, N.,
Sasikumar, S., Vijayaraghavan, R., 2010. Synthesis and characterization of bioceramic 2010. Biocomposites containing natural polymers and hydroxyapatite for bone tissue
calcium phosphates by rapid combustion synthesis. J. Mater. Sci. Technol. 26 (12), engineering. Int. J. Biol. Macromol. 47 (1), 1–4.
1114–1118. Takahashi, K., Tanabe, K., Ohnuki, M., Narita, M., Ichisaka, T., Tomoda, K., Yamanaka, S.,
Sekar, K., Balan, K.K., Sundaramoorthy, S., 2016. Comparision of electro spun tassar silk 2007. Induction of pluripotent stem cells from adult human fibroblasts by defined
fibroin-hydroxyapatite composite scaffold prepared by soaking and in-situ methods. factors. Cell 131 (5), 861–872.
Mater. Today 3 (6), 1330–1337. Takeuchi, A., Ohtsuki, C., Miyazaki, T., Kamitakahara, M., Ogata, S.-i., Yamazaki, M.,
Sen, K., Babu, K., 2004. Studies on Indian silk. I. Macrocharacterization and analysis of Furutani, Y., Kinoshita, H., Tanihara, M., 2005. Heterogeneous nucleation of hy-
amino acid composition. J. Appl. Polym. Sci. 92 (2), 1080–1097. droxyapatite on protein: structural effect of silk sericin. J. R. Soc. Interface 2 (4),
Seo, C.-W., Um, I.C., Rico, C.W., Kang, M.Y., 2011. Antihyperlipidemic and body fat- 373–378.
lowering effects of silk proteins with different fibroin/sericin compositions in mice Tanaka, T., Hirose, M., Kotobuki, N., Ohgushi, H., Furuzono, T., Sato, J., 2007. Nano-
fed with high fat diet. J. Agric. Food Chem. 59 (8), 4192–4197. scaled hydroxyapatite/silk fibroin sheets support osteogenic differentiation of rat
Shafiq, M., Jung, Y., Kim, S.H., 2016. Insight on stem cell preconditioning and instructive bone marrow mesenchymal cells. Mater. Sci. Eng. C 27 (4), 817–823.
biomaterials to enhance cell adhesion, retention, and engraftment for tissue repair. Tas, A.C., 2000. Combustion synthesis of calcium phosphate bioceramic powders. J. Eur.
Biomaterials 90, 85–115. Ceram. Soc. 20 (14), 2389–2394.
Shah, N.J., Macdonald, M.L., Beben, Y.M., Padera, R.F., Samuel, R.E., Hammond, P.T., Teimouri, A., Ebrahimi, R., Emadi, R., Beni, B.H., Chermahini, A.N., 2015. Nano-com-
2011. Tunable dual growth factor delivery from polyelectrolyte multilayer films. posite of silk fibroin–chitosan/Nano ZrO 2 for tissue engineering applications: fab-
Biomaterials 32 (26), 6183–6193. rication and morphology. Int. J. Biol. Macromol. 76, 292–302.
Shao, Z., Vollrath, F., 2002. Materials: surprising strength of silkworm silk. Nature 418 Thorfve, A., Lindahl, C., Xia, W., Igawa, K., Lindahl, A., Thomsen, P., Palmquist, A.,
(6899), 741. Tengvall, P., 2014. Hydroxyapatite coating affects the Wnt signaling pathway during
Shao, W., He, J., Han, Q., Sang, F., Wang, Q., Chen, L., Cui, S., Ding, B., 2016a. A bio- peri-implant healing in vivo. Acta Biomater. 10 (3), 1451–1462.
mimetic multilayer nanofiber fabric fabricated by electrospinning and textile tech- Tiselius, A., Hjerten, S., Levin, Ö., 1956. Protein chromatography on calcium phosphate
nology from polylactic acid and Tussah silk fibroin as a scaffold for bone tissue en- columns. Arch. Biochem. Biophys. 65 (1), 132–155.
gineering. Mater. Sci. Eng. C 67, 599–610. Tiyapatanaputi, P., Rubery, P.T., Carmouche, J., Schwarz, E.M., O'Keefe, R.J., Zhang, X.,
Shao, W., He, J., Sang, F., Ding, B., Chen, L., Cui, S., Li, K., Han, Q., Tan, W., 2016b. 2004. A novel murine segmental femoral graft model. J. Orthop. Res. 22 (6),
Coaxial electrospun aligned tussah silk fibroin nanostructured fiber scaffolds em- 1254–1260.
bedded with hydroxyapatite–tussah silk fibroin nanoparticles for bone tissue en- Urist, M.R., Huo, Y.K., Brownell, A.G., Hohl, W.M., Buyske, J., Lietze, A., Tempst, P.,
gineering. Mater. Sci. Eng. C 58, 342–351. Hunkapiller, M., DeLange, R., 1984. Purification of bovine bone morphogenetic
Shao, W., He, J., Wang, Q., Cui, S., Ding, B., 2016c. Biomineralized poly (l-lactic-co- protein by hydroxyapatite chromatography. Proc. Natl. Acad. Sci. 81 (2), 371–375.

90
M. Farokhi et al. Biotechnology Advances 36 (2018) 68–91

Vallet-Regi, M., González-Calbet, J.M., 2004. Calcium phosphates as substitution of bone tissue engineering. Mater. Sci. Eng. R. Rep. 80, 1–36.
tissues. Prog. Solid State Chem. 32, 1), 1–31. Xia, W., Chang, J., 2006. Well-ordered mesoporous bioactive glasses (MBG): a promising
Vallet-Regí, M., Balas, F., Colilla, M., Manzano, M., 2007. Bioceramics and pharmaceu- bioactive drug delivery system. J. Control. Release 110 (3), 522–530.
ticals: a remarkable synergy. Solid State Sci. 9 (9), 768–776. Xiao, G., Gopalakrishnan, R., Jiang, D., Reith, E., Benson, M.D., Franceschi, R.T., 2002.
Vallet-Regí, M., Balas, F., Colilla, M., Manzano, M., 2008. Bone-regenerative bioceramic Bone morphogenetic proteins, extracellular matrix, and mitogen-activated protein
implants with drug and protein controlled delivery capability. Prog. Solid State kinase signaling pathways are required for osteoblast-specific gene expression and
Chem. 36 (3), 163–191. differentiation in MC3T3-E1 cells. J. Bone Miner. Res. 17 (1), 101–110.
Verron, E., Khairoun, I., Guicheux, J., Bouler, J.-M., 2010. Calcium phosphate bioma- Yang, D., Kim, H., Lee, J., Jeon, H., Ryu, W., 2016. Direct modulus measurement of single
terials as bone drug delivery systems: a review. Drug Discov. Today 15 (13), composite nanofibers of silk fibroin/hydroxyapatite nanoparticles. Compos. Sci.
547–552. Technol. 122, 113–121.
Vetsch, J.R., Paulsen, S.J., Müller, R., Hofmann, S., 2015. Effect of fetal bovine serum on Young, M.F., Kerr, J.M., Ibaraki, K., Heegaard, A.-M., Robey, P.G., 1992. Structure, ex-
mineralization in silk fibroin scaffolds. Acta Biomater. 13, 277–285. pression, and regulation of the major noncollagenous matrix proteins of bone. Clin.
Wang, L., Li, C., 2007. Preparation and physicochemical properties of a novel hydro- Orthop. Relat. Res.(281), 275-294.
xyapatite/chitosan–silk fibroin composite. Carbohydr. Polym. 68 (4), 740–745. Zambuzzi, W.F., Coelho, P.G., Alves, G.G., Granjeiro, J.M., 2011a. Intracellular signal
Wang, L., Nemoto, R., Senna, M., 2002. Microstructure and chemical states of hydro- transduction as a factor in the development of “smart” biomaterials for bone tissue
xyapatite/silk fibroin nanocomposites synthesized via a wet-mechanochemical route. engineering. Biotechnol. Bioeng. 108 (6), 1246–1250.
J. Nanopart. Res. 4 (6), 535–540. Zambuzzi, W.F., Ferreira, C.V., Granjeiro, J.M., Aoyama, H., 2011b. Biological behavior
Wang, L., Nemoto, R., Senna, M., 2004a. Changes in microstructure and physico-chemical of pre-osteoblasts on natural hydroxyapatite: a study of signaling molecules from
properties of hydroxyapatite–silk fibroin nanocomposite with varying silk fibroin attachment to differentiation. J. Biomed. Mater. Res. A 97( (2), 193–200.
content. J. Eur. Ceram. Soc. 24 (9), 2707–2715. Zarrabi, A., Shokrgozar, M.A., Vossoughi, M., Farokhi, M., 2014. In vitro biocompatibility
Wang, L., Nemoto, R., Senna, M., 2004b. Effects of alkali pretreatment of silk fibroin on evaluations of hyperbranched polyglycerol hybrid nanostructure as a candidate for
microstructure and properties of hydroxyapatite–silk fibroin nanocomposite. J. nanomedicine applications. J. Mater. Sci. Mater. Med. 25 (2), 499–506.
Mater. Sci. Mater. Med. 15 (3), 261–265. Zhang, X., Xie, C., Lin, A.S., Ito, H., Awad, H., Lieberman, J.R., Rubery, P.T., Schwarz,
Wang, L., Ning, G.-L., Senna, M., 2005. Microstructure and gelation behavior of hydro- E.M., O'Keefe, R.J., Guldberg, R.E., 2005a. Periosteal progenitor cell fate in segmental
xyapatite-based nanocomposite sol containing chemically modified silk fibroin. cortical bone graft transplantations: implications for functional tissue engineering. J.
Colloids Surf. A Physicochem. Eng. Asp. 254 (1), 159–164. Bone Miner. Res. 20 (12), 2124–2137.
Wang, Y., Kim, H.-J., Vunjak-Novakovic, G., Kaplan, D.L., 2006. Stem cell-based tissue Zhang, Y.-Q., Zhou, W.-L., Shen, W.-D., Chen, Y.-H., Zha, X.-M., Shirai, K., Kiguchi, K.,
engineering with silk biomaterials. Biomaterials 27 (36), 6064–6082. 2005b. Synthesis, characterization and immunogenicity of silk fibroin-L-asparaginase
Wang, L., Li, C., Senna, M., 2007. High-affinity integration of hydroxyapatite nano- bioconjugates. J. Biotechnol. 120 (3), 315–326.
particles with chemically modified silk fibroin. J. Nanopart. Res. 9 (5), 919–929. Zhang, F., Zuo, B.Q., Zhang, H.X., Bai, L., 2009. Studies of electrospun regenerated SF/
Wang, T., Porter, D., Shao, Z., 2012. The intrinsic ability of silk fibroin to direct the TSF nanofibers. Polymer 50 (1), 279–285.
formation of diverse aragonite aggregates. Adv. Funct. Mater. 22 (2), 435–441. Zhang, Y., Wu, C., Friis, T., Xiao, Y., 2010. The osteogenic properties of CaP/silk com-
Wang, Q., Chu, Y., He, J., Shao, W., Zhou, Y., Qi, K., Wang, L., Cui, S., 2017. A graded posite scaffolds. Biomaterials 31 (10), 2848–2856.
graphene oxide-hydroxyapatite/silk fibroin biomimetic scaffold for bone tissue en- Zhang, D., Chen, K., Wu, L., Wang, D., Ge, S., 2012. Synthesis and characterization of
gineering. Mater. Sci. Eng. C 80, 232–242. PVA-HA-silk composite hydrogel by orthogonal experiment. J. Bionic Eng. 9 (2),
Webster, T.J., Massa-Schlueter, E.A., Smith, J.L., Slamovich, E.B., 2004. Osteoblast re- 234–242.
sponse to hydroxyapatite doped with divalent and trivalent cations. Biomaterials 25 Zhang, X., Fan, Z., Lu, Q., Huang, Y., Kaplan, D.L., Zhu, H., 2013. Hierarchical biomi-
(11), 2111–2121. neralization of calcium carbonate regulated by silk microspheres. Acta Biomater. 9
Wei, G., Ma, P.X., 2004. Structure and properties of nano-hydroxyapatite/polymer (6), 6974–6980.
composite scaffolds for bone tissue engineering. Biomaterials 25 (19), 4749–4757. Zhang, J., Liu, W., Schnitzler, V., Tancret, F., Bouler, J.-M., 2014. Calcium phosphate
Wenk, E., Wandrey, A.J., Merkle, H.P., Meinel, L., 2008. Silk fibroin spheres as a platform cements for bone substitution: chemistry, handling and mechanical properties. Acta
for controlled drug delivery. J. Control. Release 132 (1), 26–34. Biomater. 10 (3), 1035–1049.
Wnek, G.E., Bowlin, G.L., Ito, A., Ohgushi, H., 2008. Calcium phosphate ceramics: new Zhao, C., Tan, A., Pastorin, G., Ho, H.K., 2013. Nanomaterial scaffolds for stem cell
generation produced in Japan. In: Encyclopedia of Biomaterials and Biomedical proliferation and differentiation in tissue engineering. Biotechnol. Adv. 31 (5),
Engineering, Second edition. CRC Press, pp. 461–469 Online Version. 654–668.
Wu, C., Fan, W., Chang, J., 2013. Functional mesoporous bioactive glass nanospheres: Zhao, Y., Chen, J., Chou, A.H., Li, G., LeGeros, R.Z., 2009. Nonwoven silk fibroin net/
synthesis, high loading efficiency, controllable delivery of doxorubicin and inhibitory nano-hydroxyapatite scaffold: preparation and characterization. J. Biomed. Mater.
effect on bone cancer cells. J. Mater. Chem. B 1 (21), 2710–2718. Res. A 91( (4), 1140–1149.
Wu, S., Liu, X., Yeung, K.W., Liu, C., Yang, X., 2014. Biomimetic porous scaffolds for bone

91

You might also like