Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Jan 1.
Published in final edited form as: Prog Polym Sci. 2012 Jan 1;37(1):1–17. doi: 10.1016/j.progpolymsci.2011.07.003

Natural and Genetically Engineered Proteins for Tissue Engineering

Sílvia Gomes 1,2,3, Isabel B Leonor 1,2, João F Mano 1,2, Rui L Reis 1,2, David L Kaplan 3,
PMCID: PMC3207498  NIHMSID: NIHMS311808  PMID: 22058578

Abstract

To overcome the limitations of traditionally used autografts, allografts and, to a lesser extent, synthetic materials, there is the need to develop a new generation of scaffolds with adequate mechanical and structural support, control of cell attachment, migration, proliferation and differentiation and with bio-resorbable features. This suite of properties would allow the body to heal itself at the same rate as implant degradation. Genetic engineering offers a route to this level of control of biomaterial systems. The possibility of expressing biological components in nature and to modify or bioengineer them further, offers a path towards multifunctional biomaterial systems. This includes opportunities to generate new protein sequences, new self-assembling peptides or fusions of different bioactive domains or protein motifs. New protein sequences with tunable properties can be generated that can be used as new biomaterials. In this review we address some of the most frequently used proteins for tissue engineering and biomedical applications and describe the techniques most commonly used to functionalize protein-based biomaterials by combining them with bioactive molecules to enhance biological performance. We also highlight the use of genetic engineering, for protein heterologous expression and the synthesis of new protein-based biopolymers, focusing the advantages of these functionalized biopolymers when compared with their counterparts extracted directly from nature and modified by techniques such as physical adsorption or chemical modification.

Keywords: Biomaterials, tissue engineering, natural polymers, recombinant technology, chimeric proteins

1. Introduction

Treatment of injured tissues or organs focuses on the use of autologous and allogenic grafts [1]. However, this practice has significant limitations for the patient and health systems worldwide. Autologous grafts cause donor site morbidity and consequent loss of organ functionality. Allografts are associated with risk of disease transmission and require the use of immunosuppressants with associated side effects [24]. In the field of orthopaedic implants, autologous and allogenic grafts account for 90% of the grafts currently used, with synthetic materials (metals, polymers, ceramics and composite systems) used in 10% of surgery cases [3, 5, 6]. In the three types of grafts there are numerous cases of implant failure as a consequence of undesirable local tissue responses resulting in implant loosening, insufficient osseointegration, osteolysis, inflammation and infection [24]. These complications account for a failure rate of 13 to 30% in the case of autografts and 20 to 40% for allografts [2].Besides autologous and allogenic grafts, synthetic materials have also been used for controlled drug delivery systems, scaffolds design and orthopaedic fixation as screws, pins or rods [7, 8]. Nevertheless, most synthetic polymers are too hydrophobic and need additional bulk or surface modifications to render the material more biocompatible and suitable for implantation [9]. Therefore, there is a need for alternatives to these practices. Tissue engineering and regenerative medicine offer an approach to circumvent the present therapies with new methods of health care treatment with the purpose of improving the quality of life [2, 10]. This improvement can come in the form of new cytocompatible and non-toxic biomaterials for the manufacture of a new generation of scaffolds comprising adequate mechanical and structural support and able to control cell attachment, migration, proliferation and differentiation [11, 12]. Furthermore, this future generation of scaffolds should not behave as a permanent prosthesis but instead should perform as bio-resorbable temporary implants, allowing for the body to heal itself at the same rate as the implant degradation [11, 13].

In recent years a small number of synthetic biodegradable polymers, mainly polyesters containing glycolic (PLG) or lactic (PLL) acids and caprolactone (PCL) were approved by the Food and Drug Administration (FDA) for use in sutures [13]. Epicel™ (autologous keratinocyte skin graft to treat severe burn victims from Genzyme Biosurgery, Cambridge, MA), Carticel® (autologous chondrocyte transplantation to treat cartilage injury from Genzyme Biosurgery, Cambridge, MA) [14], MACI™ for matrix-induced autologous chondrocyte implantation (Genzyme Biosurgery, Cambridge, MA) where chondrocytes are supplied seeded onto a type I/III collagen scaffold secured to the skin injury with fibrin glue [15], and Apligraf (bovine collagen I matrix seeded with keratinocytes for wound care from Organogenesis, Canton, MA) [16], are products for cell therapy also available in the market. Other examples of products already commercially available are Atrigel® (Atrix Laboratories, Fort Collins, Co, USA) a system of biodegradable polymers for drug delivery [17] and the calcium phosphate based products Collagraft (Zimmer, Warsaw, IN; and Collagen Corporation, Palo Alto, CA) and ProOsteon (Interpore international, Irvine, CA) for bone applications [2]. However, since giving a detailed description of these and other products available in the market is not the purpose of this review we advise the reader to address other reports for more information [2, 1317].

Despite the enormous research effort during the last few decades materials scientist have not fully developed a new generation of biocompatible biomaterials [13]. This limitation of tissue engineering to move forward from the laboratory into the clinic is the result of many issues, including legal, the need to develop functional blood vessel networks to nourish the new tissues mainly inside scaffolds, inability of the biomaterials to promote the formation of functional tissues, and many related issues [13]. For these reasons it is critical to develop the next generation of biomaterials that will address the limitations above. New approaches in the fields of bionanotechnology, protein engineering and bionano-fabrication will play a role in the development of these next generation biomaterials [1820].

In this review we address some biopolymers already being used or with potential applications in regenerative medicine and tissue engineering, giving special focus to proteins and protein-based biomaterials. Additionally, we will also focus on the different approaches used for functionalization of these biomaterials in order to improve performance, mechanical efficiency, biocompatibility and degradability, usually with a goal towards control of these processes. This overview will be followed by a description of the novel design approaches, namely genetic engineering, enabling the synthesis of new protein-based biopolymers inspired in nature but without many of the drawbacks of their native counterparts when extracted directly from natural sources. Additional information can be found in recent reviews addressing the use of biomimetic materials in tissue engineering [21], the application of protein templates for tissue engineering [12], the synthetic modification of proteins and peptides [22] and the use of bioengineering for biomaterials design [19, 20].

2. Natural proteins for biomedical applications

The similarity between natural polymers and the macromolecules forming extracellular matrices suggests an innate ability for some of these polymers to interact with the cells and the biomolecules present in host tissues, inducing mild immunological reactions when compared with synthetic materials [11, 23, 24]. Natural polymers such as fibrin, fibronectin, collagen, elastin, silk, keratin, chitosan, alginate, amylose/amylopectin and hyaluronic acid are widely used in tissue engineering [23, 25]. Within the myriad of biopolymers present in nature, proteins are considered to be one of the most sophisticated groups in terms of chemistry [26]. Therefore some proteins with potential use in the biomedical field will be addressed in the next paragraphs. Table 1 addresses some of the basic features of the animal proteins described in this section.

Table 1.

Basic features of some proteins with potential applications in the biomedical field.

Protein Main functions Basic structure Relevant properties
Collagen [31, 34] Structural protein in tissues such as connective tissue, tendon, skin, bone and cartilage Three parallel polypeptide chains formed by GXY (G - glycine, X - usually proline, Y - usually 4-hydroxyproline) repeats and arranged in triple helix Biodegradability, low antigenecity and biocompatibility
Fibronectin [47, 49] Structural support and cell signalling Dimer of two non-identical polypetide chains bonded at the carboxyl end by disulfide bonds Multi-domain protein with cell (RGD motif), collagen and fibrin binding motifs
Elastin [56, 182, 183] Structural protein found predominantly in connective tissue of arteries, ligaments, skin and lung Cross-linked units of tropoelastin formed by hydrophobic (often 3 to 6 repeats of GVGVP, GGVP and GVGVAP) and hydrophilic lysine domains Temperature dependent self-assembly and phase separation behaviour
Fibrin [72, 184] Blood clotting, fibrinolysis, cellular and matrix interactions, inflammation and wound healing Resultant from the polymerization and crosslinking of fibrinogen units after thrombin cleavage Growth factor binding and interaction with cells such as platelets, leucocytes, fibroblasts and endothelial cells
Laminins [80] Major components of basement membranes underlying epithelial and endothelial cells and embedding Schwann, muscle and fat cells Heterotrimers of one β, one α and one γ chain, which represent different gene products Self-assembly and binding to several matrix proteins and integrins
Vitronectin [84] Regulates clot formation and immune response, provides biological cues for cell adhesion, migration and proliferation and extracellular anchoring In human blood is found as a single chain or as a dimer while in the extracellular matrix exists as a disulfide-linked vitronectin multimer Multi-domain protein with an RGD motif to mediate the attachment and spreading of cells and binding motifs for collagen, heparin, plasminogen, glycosaminoglycan and fibrin binding motifs
Keratin [88, 185] Structural protein in the cytoskeletons of vertebrate epithelial cells and epidermis appendages such as hair, nails and wool Formed by α-helical coiled-coil dimers assembled into 10 nm wide filaments Biocompatibility, good cell attachment and growth
Silk [93, 98] Building element of many arthropod nests, cocoons and prey traps Highly repetitive core domain of alternating poly-A hydrophobic and G rich hydrophilic motifs Self-assembly and remarkable mechanical properties
Mussel adhesive proteins (MAPs) [119] Substrate adhesion Repetitive sequence, with molecular weights ranging between 5 and 120 kDa and high presence of 3,4-dihydroxyphenyl-L-alanine (DOPA) Function over a wide range of temperatures, humidity and salinity and form permanent bonds to a wide variety of surfaces

Collagen is synthesized by fibroblasts and other cell types such as chondrocytes [27] and osteoblasts [28] and is the most abundant protein in the mammalian body, accounting for 20–30% of the total protein [29]. Its primary functions in tissues are to provide mechanical support [30] and to control cell adhesion, cell migration and tissue repair [31]. Collagens form a large family of triple helical molecules with about 28 different types described [32]. All collagens share the same triple-helical structure where three parallel polypeptides, α-chains, coil around each other forming a right handed triple helix chain. In animals these collagen triple helices are known as tropocollagen and its hierarchical organization into more complex structures generates the fibers and networks in tissues such as bone, skin tendons, basement membranes and cartilage [33, 34]. Collagen is easy to modify and process and its abundance, nonantigenicity, biodegradability, biocompatibility and plasticity make collagen a promising biopolymer for applications in the medical and pharmaceutical fields and tissue engineering purposes [30]. Reconstituted gels of Type I collagen are widely used for biomedical applications and its main sources are animal tissues such as skin and tendons [25, 35]. Collagen scaffolds have been extensively used for soft tissue repair [36], vascular [37] and dermal tissue engineering [38, 39], bone repair [40] and as a carrier for the delivery of drugs [41] and biologically active molecules [42]. Additionally, collagens can also be used to fabricate microspheres for cell encapsulation [43] and drug loading for controlled release [44].

However, despite the wide range of applications collagens matrices lack the mechanical properties required for hard tissue during initial implantation. For this reason collagen is often blended with other materials, either synthetic [45] or natural [46], to overcome mechanical limitations [12].

Fibronectin is also a component of the extracellular matrix with important functions such as structural support and signalling for cell survival, migration, contractility, differentiation and growth factor signalling [47]. Fibronectin is synthesized by different cell types, such as fibroblasts and is secreted as a dimer with disulfide bonds formed between the 230–270KDa subunits. These subunits are formed by three types of repeating modules named type I, II and III [48]. Fibronectin is a multi-domain glycoprotein with a remarkable number of biological functions, many of which are mediated through interactions with integrins, such as via the RGD sequences present in fibronectin. Besides binding to cell integrins, fibronectin binds to other biologically important molecules such as heparin, collagen/gelatin and fibrin [49]. Since fibronectin is biocompatible and easily recognized by cell integrins, the use of fibronectin or domains of the protein to functionalize scaffolds for tissue engineering is often considered [50]. Polymeric scaffolds of chitosan [51, 52], collagen [53] and hyaluronic acid [54] have been modified with fibronectin to improve cell adhesion and proliferation.

Additionally, fibronectin-mimetic peptide-amphiphiles were used in the fabrication of nanofibers and gels with excellent cell adhesion properties [50]. Another strategy was to prepare fibronectin-terminated multilayer films of poly-lysine and dextran sulphate for the study of the spreading behaviour of human umbilical vein endothelial cells. The cells spread to a greater extent and in a more symmetric manner on the films coated with fibronectin, suggesting that such fibronectin coated films may represent a promising strategy to control cell interactions with the materials in tissue engineering [55].

Together with collagen and fibronectin, elastin is also part of the core architecture supporting cell adhesion and growth [56]. Elastin fibers are mainly present in connective and vascular tissues, the lungs and skin. Elastin is a polymer of tropoelastin monomeric precursor and elastin fibers are an important component of the extracellular matrix to impart elasticity to organs and tissues. Hydrophobic domains present in the elastin sequence are responsible for these elastic properties [57, 58]. Elastin also has chemotactic activity, inducing cell proliferation and regulating cell differentiation, with the specific binding of integrin αvβ3 to the C-terminus in tropoelastin [59]. Due to its characteristics elastin is of interest for drug delivery and tissue engineering and has been used in the fabrication of hybrid materials in combinations with collagen [60], polycaprolactone (PCL) [61] and silk [62] for the production of vascular grafts [63], hydrogels [64], bone repair [65] and for drug delivery [66]. However, the crosslinking that occurs between the water-soluble tropoelastin monomers to form the insoluble and stable elastin fibers limits the use of elastin from animal origin [56]. Therefore artificial proteins incorporating elastin-like peptides have been of interest for the development of new protein-based biomaterials [67, 68] with properties similar to native elastin [69].

Fibrin is another example of a specialized extracellular matrix protein with potential application for tissue engineering. However, unlike collagen, elastin and fibronectin, fibrin networks form mostly during blood clotting. Fibrin is the result of fibrinogen polymerization in the presence of thrombin [70]. Fibrinogen is a 340 kDa protein present in plasma formed by pairs of three different polypeptides, Aα, Bβ and γ, held together by disulfide bridges [71]. Fibrin and fibrinogen are two important components in blood clotting, fibrinolysis, cellular and matrix interactions, inflammation, wound healing and neoplasia [72]. In the particular case of clot formation, thrombin cleavage both Aα and Bβ chains at their N-termini, leading to the exposure of polymerization sites in both chains [73]. Subsequently the combination of these polymerization sites leads to the formation of double-strand twisted fibrils. These fibrin protofibrils undergo lateral aggregation and form branches, producing a three dimensional network [74]. Blood clots are further stabilized by covalent bonds formed by the plasma transglutaminase, factor XIII, making the clot more mechanically stable and less susceptible to enzymatic digestion [75]. Fibrin is a viscoelastic polymer and is used clinically as a medical adhesive; fibrin sealants are FDA approved.

Furthermore, fibrin is also used for skin repair, replacing sutures and staples in fixation of skin grafts promoting a better wound healing [76], and in the transplantation of keratinocytes in burned patients [35]. Fibrin is also a promising biopolymer for applications in tissue engineering, in the repair of damaged tissues [77, 78], and drug delivery, as a carrier for growth factors [79]. Additionally, two proteinaceous components of the extracellular matrix, laminins and vitronectin, are mainly used to coat synthetic and natural polymer-based materials to improve cellular response. Laminins are cell adhesion glycoproteins localized in the extracellular matrix of the basement membrane and are able to bind to other matrix proteins [80]. Recently, lamimin-derived peptides have been used as coatings to induce the adhesion of different cell types such as hepatocytes [81] and human dermal fibroblasts [82]. Also, these peptides are being studied for drug delivery in the development of targeting drug-loaded systems for cancer treatments [83]. Vitronectin is a multifunctional glycoprotein present in the extracellular matrix where it binds to glycosaminoglycans, collagen, plasminogen and urokinase-receptor and its RGD allows it to mediate the adhesion and spreading of cells [84]. This multipartner binding makes vitronectin an attractive biopolymer for tissue engineering and to induce cell attachment when used as a surface coating [85, 86].

The proteins described above are extracellular matrix proteins and have been more commonly used for tissue engineering and regenerative medicine applications. However, in the past few years other proteins have also emerged as potential biopolymers for the fabrication of new biomaterials, such as Keratin [87]. Moreover, since it is a protein shared by all mammals with a highly conserved amino acid sequence it is expected to offer good cell and tissue responses [88]. keratin fibers are hierarchically structured proteins present in hard and filamentous structures, such as hairs, horns and nails [87]. The presence of a LDV cell binding domain in keratin amino acid sequence [87] suggests utility for the fabrication of scaffolds for tissue engineering. Keratin based biomaterials have been used to support adhesion, spread and growth of L929 fibroblast cells [89], and the growth and differentiation of osteoblasts (MC3T3-E1) [90]. Keratin films have an inhibitory effect on the IgE receptor-stimulated histamine release from mast cells, making it suitable for use in antiallergenic materials [91].

As collagen and keratin, silk is another example of a hierarchically structured fibrous protein. Silk is characterized by its outstanding mechanical properties out-competing high performance man made fibers such as Kevlar, nylon and high-tensile steel, and by its self-assembly leading to fibers with a complex hierarchical arrangement [26]. Silk-protein-based fibers are produced by insects [92] and spiders [93] which use it for different ends such as cocoon and nest construction. However, despite the multitude of functions and different protein structures, many silk-based fibers have similar amino acid compositions and high levels of crystallinity. Silkworm silk produced by the silkworm species Bombyx mori is the most well studied silk protein [92]. The silk fiber is formed by two microfilaments embedded in glue-like glycoproteins named sericin which works as a coating. Each microfilament results from the assembly of a hydrophobic ~370 kDa heavy-chain fibroin protein, a relatively hydrophilic ~25 kDa light-chain fibroin and a 30 kDa P25 protein [94]. Spider dragline silk has a slightly different structure with a core filament formed by two spidroin molecules, major ampullate spidroin protein 1 (MaSp1) and 2 (MaSp2), coated by glycoproteins and lipids [95]. The remarkable mechanical features of the different types of silk are in part due to the presence of α-helix and β-turns, responsible for its elastic properties. These elastic domains alternate with β-sheet motifs which confer toughness to silk fibers. The strong molecular cohesion occurring with amide-amide interactions in the β-sheet crystalline regions is thought to be responsible for the remarkable stiffness of silk fibers [96]. In B. mori silk, the hexapeptide repeat GAGAGS is involved in the formation of the β-sheets. In spider silk besides GA sequences there are also poly-Ala blocks and both motifs contribute for the formation of anti-parallel β-sheets [96]. These poly-A and GA motifs are embedded in amorphous regions formed by either GGX (X can be Tyr, Leu or Gln) or GPGXX motifs believed to be responsible for the elastic features [97]. The outstanding mechanic features and biocompatibility are reasons why silk has been used through the millennia in such diverse applications as hunting, fabrication of paper, wound dressing, textiles and sutures [98]. With new technologies in the fields of polymer synthesis and processing, silk continues to be an important topic of research for biomaterial and biomedical research. In the case of B. mori silk, sericulture provides the product used by the textile industry and in medical sutures [93]. Additionally, this silk is being studied for tissue engineering in the form of scaffolds for a range of tissue needs, such as corneal regeneration [99, 100], cartilage repair [101, 102], vascular grafts [103, 104], bone regeneration [105, 106] and drug delivery [107, 108]. As mentioned above B. mori silk is available in large supplies from sericulture, and is therefore most commonly used for the above studies. In the case of spiders, it is difficult to breed spider species due to their cannibalistic behaviour. With the advance of biotechnology tools it is now possible to bioengineer spider silk genes to produce spider silk-like proteins [109], such as for tissue engineering [110], cell culture [111], nerve regeneration [112, 113] and wound dressings [114].

Mussel adhesive proteins (MAPs) are produced by marine mussels and used in the formation of the byssal threads which allow the animal to anchor to substrates. A common feature to all the adhesives produced by mussels is the presence of the amino acid 3,4-dihydroxyphenyl-L-alanine (DOPA). DOPA residues are key elements for the chemisorption to substrates underwater and the crosslinking process within the adhesive molecules [115]. These natural adhesives display outstanding properties in terms of function under harsh marine environments with wide temperature, salinity and humidity fluctuations and the mechanical effects of tides, waves and currents [116]. These remarkable properties make MAPs attractive biomaterials as bioadhesives. MAPs have been used as bioadhesives for cells [117] and as self-adhesive micro-encapsulated drug carriers for biotechnological, tissue engineering and biomedical applications [118]. MAP derivatives were also used in the fabrication of adhesive-coated meshes as wound sealants, replacing tradition sutures, staples and tacks [119].

The proteins addressed above are widely used for tissue engineering and biomedical applications and can be obtained from animal sources. Moreover, the majority of proteins used in the development of new scaffolds for tissue engineering are extracted from natural sources. In this way, in most cases these polymers need further modifications to make them more suitable for different biomedical applications. The next section refers to physical and chemical approaches used for the functionalization of these biomaterials.

3. Techniques for the functionalization of protein-based biomaterials

The properties of protein-based biomaterials can be improved by combining them with bioactive molecules to enhance in vitro and/or in vivo functions. The surface of protein-derived scaffolds can be modified by physical adsorption, physical entrapment (encapsulation) or by chemical modification. These techniques are commonly used to functionalize protein-based biomaterials with different biologically active molecules, such as growth factors and antibiotics, improving cell and tissue responses.

Physical adsorption is a simple immobilization procedure and is frequently used to attach bioactive molecules such as extracellular matrix proteins or growth factors to the surface of scaffolds by dip coating [120]. Adsorption efficiency is dependent on the physical and chemical properties of the material, including wettability, surface topography, functional groups, pH and electrical charge, among other factors [121]. Many biomaterials are hydrophobic, therefore, methods are needed to enhance wettability to make them more hydrophilic. Physical methods such as bombardment with ions, UV light and plasma modification are used to disrupt chemical bonds between carbon and non-carbon atoms generating unsaturated bonds and radicals which react with oxygen, increasing hydrophilicity and enhancing reactivity towards biological molecules [121]. Natural polymers have the advantage of being rich in reactive chemical groups (hydroxyl, carboxyl, amide) which make them more hydrophilic and capable of interacting with bioactive molecules. Collagen and silk are examples of protein-based materials that have been functionalized through adsorption of bioactive molecules, including bone morphogenetic proteins (BMPs) [122, 123], basic fibroblast growth factor (bFGF) [124], vascular endothelial growth factor (VEGF) [125] and therapeutic compounds such as antibiotics [126] and heparin [127] as it is summarized in Table 2. In most of these studies the protein-based scaffolds were soaked in a solution containing the bioactive component. In other cases the proteins were blended with the bioactive molecule in solution and then cast to form scaffolds [128].

Table 2.

Summary of protein based scaffolds functionalized with different bioactive molecules.

Protein matrices Modification process Modifying molecule Application References
Collagen EDC/NHS covalent immobilization VEGF Vascularisation/angiogenesis [186, 187]
FGF/VEGF Vascularisation/angiogenesis [188]
Heparin BMP/FGF/PDGF delivery system [189191]
Traut’s Reagent and Sulfo-SMCC covalent immobilization poly-Histidine antibody BMP delivery system [192]
VEGF Vascularisation/angiogenesis [193]
Adsorption BMP-2 BMP delivery system [194]
FGF Cartilage regeneration/Growth factor delivery [195, 196]
Microsphere encapsulation BMP-7 BMP delivery system [197]
VEGF Vascularisation/angiogenesis [198]
Gelatin EDC covalent immobilization TGF-beta Cartilage regeneration [199]
Adsorption TGF-beta/IGF Cartilage regeneration [200]
FGF Growth factor delivery [195]
Fibronectin Cartilage regeneration [201]
Microsphere encapsulation TGF-beta Chondrogenesis/cartilage regeneration [202, 203]
BMP-2/VEGF Angiogenesis and osteogenesis [204]
BMP-2 Growth factor delivery [205]
Fibrin Microsphere encapsulation FGF Angiogenesis [206]
BMP-2 Bone regeneration [207, 208]
Patterning immobilization FGF-2 Tissue engineering [209]
Heparin Michael type addition BMP-2 Bone/ligament regeneration [210]
HGF Hepatocyte differentiation [211]
Silk Cyanuric chloride immobilization Lactose Hepatocyte attachment [212]
Crosslinking Gelatin Tendon tissue engineering [213]
Adsorption FGF Growth factor delivery [124]
Gelatin Drug/Growth factor delivery [214]
Collagen/chodroitin-6-sulfate/hyaluronan Tendon tissue engineering [215]
Collagen tendon tissue engineering [216, 217]
BMP-2 Bone regeneration [218]
Microsphere encapsulation IGF Drug/Growth factor delivery [219]
Blend Gelatin Tissue engineering [220]
Silk/collagen Adsorption SDF-1 tendon tissue engineering [221]

Since adsorption is based on relatively weak or moderate electrostatic, van der Waals, hydrogen and hydrophobic interactions the binding stability of the adsorbed molecules can vary depending on environmental conditions. In this way, changes in pH, ionic strength and adsorbed species concentration of the surrounding medium can result in an uncontrolled release of the immobilized species [120]. For example, bone morphogenetic proteins (BMPs) tend to diffuse away from the fracture area and high doses are required to induce the desired osteogenic response. The release profile of BMP-2 from collagen sponges shows an initial burst during the first 10 minutes, where the carrier loses around 30% of the BMP-2, followed by slow release during the next 3 to 5 days. This initial burst release can cause clinical complications, such as ectopic bone formation, soft tissue hematomas and bone resorption [129, 130].

To overcome these issues, covalent immobilization has been widely used since it has the advantage of providing stable attachment of bioactive agents to polymeric scaffolds. With proper design, covalent conjugation has proven to be a very effective strategy to control the release profile of the immobilized agent since these molecules are retained for longer time periods at the delivery site, when compared with adsorption [11]. Carbodiimide coupling is broadly used in protein chemistry to react activated surface carboxylic acid groups from protein-based scaffolds with the amines present on the peptide or protein to be immobilized [131, 132]. Carboxylic groups are activated by using 1-ethyl-3-(3-dimethylaminopropyl)-carbodiimide (EDC) mixed with either N-hydroxysuccinimide (NHS), dicyclohexyl-carboiimide (DCC) or carbonyl diimidazole (CDI) [131, 132]. This basic protein chemistry has been extensively used to immobilize molecules as it is shown in Table 2, including BMPs and RGD peptides onto silk and collagen scaffolds. A drawback of this coupling method is the difficulty in characterizing the new peptide-protein scaffolds, due to the background noise from the protein scaffold itself, making it difficult to measure the signal coming from the small amount of peptide immobilized on the scaffold surface in order to quantify how much peptide was immobilized [132]. Another drawback can be the presence of reactive amine groups aside from the N-terminal amine. These reactive side groups need to be protected, followed by deprotection after the coupling chemistry is carried out, although the use of harsh conditions can affect the biological activity of the immobilized molecules [131].

Glutaraldehyde, polyethylene glycol diacrylate and hexamethylene diisocyanate can be used to bridge the amine groups present in the peptide or protein to be immobilized and in the protein based scaffolds [132, 133]. Glutaraldehyde has been used to couple insulin [134] and lipase [135] onto silk scaffolds and to crosslink blends of collagen and silk [136]. However, the potential release of toxic residual molecules formed during the crosslinking process is a concern if these biomaterials are to be used for biomedical applications [133].

Encapsulation of bioactive molecules within protein matrices has also been explored as a method to control the release of bioactive agents. In many cases chemical modifications are required in order to have better control over the release profile of the encapsulated molecules. Crosslinked gelatin microspheres later impregnated with basic fibroblast growth fact (bFGF) and loaded into collagen sponges were used in order to have controlled release of bFGF at a defect site [137] (Table 2). Furthermore, crosslinked collagen microspheres loaded with bovine serum albumin (BSA) and nerve growth factor were prepared and release profiles assessed [44]. In both studies collagen microspheres had to be crosslinked in order to reduce the initial burst and attain better control of protein release. EDC and NHS were also used as coupling reagents to covalently bind 2,3-dihydroxybenzoic to gelatin microspheres, which were incorporated into a reconstituted collagen scaffold for a wound dressing [138]. Silk microspheres were used for the encapsulation of bioactive proteins and other molecules, exploiting the self-assembly properties of silk to control the release profile [139].

Many formulations and delivery strategies have been explored in order to achieve functionalization and sustained release of different molecules. However, in the particular case of bioactive proteins loaded into protein-based scaffolds, protein structure and topology must be considered in order to prevent protein denaturation, as a consequence of the adsorption or immobilization processes, and protein aggregation during the release period which can result in the loss of bioactivity [11, 140]. Proteins in denatured forms are often antigenic and can induce immunogenic reactions with negative clinical consequences [140].

Most of the methods being used for functionalization of polymeric structures and drug release have some disadvantages and new strategies are clearly needed. Advances in the fields of self-assembly and biotechnology, mainly via recombinant DNA approaches, can offer some important options to address the deficiencies noted above, to help in the development of the next generation of biomaterials. The importance of recombinant DNA technology for the development of new protein based biomaterials will be the focus of the next section.

4. Recombinant proteins for tissue engineering

Since mammalian tissues are the main source of materials such as collagen, gelatin, fibrin and elastin there are concerns with disease transmission and immunogenic responses in in vivo studies, as well as batch-to-batch variability [19, 141]. To overcome these limitations, peptide synthesis and recombinant DNA protein methodologies have been explored. Chemical synthesis can be a quick and efficient method to fabricate short peptides in relatively small quantities [142]. However, the synthesis of peptide sequences with more than 35–40 amino acids is not feasible due to a drop in yield and efficiency paralleled by an exponential increase in cost [143]. Recombinant DNA technology provides well established protocols for cloning, mutation and gene fusion in different host cells for the expression of peptides and proteins with a broad range of sizes [144]. Furthermore, the increased efficiency in making synthetic oligonucleotides and the use of standardized kits and protocols for cloning and protein expression make the transgenic production approach more cost-effective for large scale protein production [144]. Besides engineering biological components already present in nature as shown in Table 3, the field of synthetic biology is also focused on the design of new peptides and protein sequences. This can be achieved by establishing new artificial self-assembling peptides or by fusing together different bioactive domains or protein motifs that are not otherwise found together in nature. Table 4 gives an overview of the studies published during the past few years using this approach [144]. Since genetic engineering offers the possibility of altering the amino acid sequence of the expressed protein by adding or substituting codons, it is possible to generate alternative sequences with tunable properties that can be used as promising biomaterials for medical applications.

Table 3.

Biopolymers expressed in recombinant systems and their potential uses.

Protein Expression system Advantages/Applications References
Collagen I Transgenic corn Food and pharmaceutical industries [222]
Yeast Pichia pastoris Identical 4-hydroxyproline content to human collagen; medical applications such as corneal replacement [147, 150, 223]
Yeast Saccharomyces cerevisiae Study of collagen expression and maturation [224, 225]
Mammalian HT1080 cells Optimization of recombinant collagen expression and isolation methodology [226, 227]
Insect cells Optimization of recombinant collagen expression and isolation methodology; Structural studies [228, 229]
Mammalian, mouse milk Optimization of recombinant collagen expression [230, 231]
E. coli JM109 strain Large quantities production/Therapeutic, biomaterial, or bioengineering applications, [232]
E. col Bone tissue engineering [151]
Collagen II Yeast Pichia pastoris Identical 4-hydroxyproline content to human collagen [147]
Insect cells Optimizing recombinant collagen expression systems [233]
Collagen III Yeast Pichia pastoris Higher production level; Identical 4-hydroxyproline content to human collagen; Scientific and medical applications such as corneal replacement [147, 150, 234]
Yeast Saccharomyces cerevisiae Optimizing recombinant collagen expression systems [235]
Insect cells 4-hydroxyproline content similar to human collagen; Study of collagen chain association and folding [236, 237]
Silkworm Viable expression system for bulk protein expression [238]
Collagen V Mammalian cells Structural studies [239]
Collagen VI Mammalian cells Collagen and heparin binding studies [240]
Collagen VII Mammalian cells Study of dystrophic epidermolysis genetic disorder [241]
Collagen X Mammalian HEK293 cells Optimizing recombinant collagen expression systems [242]
Collagen XI E. coli BL21 Study the regulation of collagen fibrillogenesis [243]
Collagen-like protein Mammalian HT1080 cells Biomedical applications [152, 244]
Gelatin-like proteins Yeast Pichia pastoris Biomedical applications [245]
Elastin-like peptides Yeast Pichia pastoris Optimizing cloning and expression process [246]
E. coli strain BL21-Gold Vascular replacement; Tissue engineering, controlled drug release and cell encapsulation; Biomedical applications [167, 247253]
E. coli BLR strain Biomedical applications [254]
Spider silk major ampullate from Nephila clavipes E. coli RY-3041 Structural studies/Biomedical applications [255, 256]
E. coli SG 13009pREP4 Structural studies/Biomedical applications [155]
E. coli BL21 Structural studies/Biomedical applications [257259]
E. coli M109 strain Structural studies/Biomedical applications [260]
Yeast Pichia pastoris Structural studies/Biomedical applications [158]
Spider silk major dragline proteins ADF-3 and ADF-4 from Araneus diadematus E. coli BLR strain Structural studies/Biomedical applications [157]
Spider silk flagelliform from Nephila clavipes E. coli BL21 strain Structural studies/Biomedical applications [156]
Spider silk like proteins - NcDS, (SpI)7 and [(SpI)4/(SpII)1]4 E. coli BL21 strain Structural studies/Biomedical applications [261]
Fibrinogen Mammalian cells Fibrin sealant [262]
Yeast Pichia pastoris Fibrin sealant [263]
Fibronectin E. coli Cell adhesion [264, 265]

Table 4.

New chimeric proteins with potential application in the biomedical field.

Fusion protein Expression system Applications References
R136K (FGF-1 mutant) + collagen biding domain E. coli BL21 (pLysS) strain Selective binding to collagen and potent angiogenic, mitogenic and chemotactic activity for endothelial cells [266, 267]
VEGF + collagen biding domain E. coli BL21 strain Improve diabetic wound healing [268]
FGF + fibronectin cell binding domain E. coli JM109 strain Stimulates angiogenesis, biomedical applications/tissue engineering [269]
FGF + collagen binding domain E. coli BL21 strain Delivery systems/Biomedical applications/Tissue engineering [270]
FGF + glutathione S-transferase (GST-bFGF) E. coli Stimulate the growth of human umbilical vein endothelial cells [271]
FGF2 + Fibronectin (FGF2-FNIII9-10) E. coli TOP10 strain Delivery of bioactive molecules [179]
EGF + collagen binding domain E. coli BL21(DE3) strain Delivery systems/Biomedical applications/Tissue engineering [270]
EGF-collagen Insect cells Tissue engineering applications [181]
EGF + immunoglobulin G (IgG) Fc region (EGF-Fc) E. coli BL21 strain Cell adhesion [272]
Silk + elastin (SELP-47 K) E. coli Promote cell attachment and growth/Tissue Engineering [273]
Spider silk + dentin matrix protein E. coli RY-3041 strain Biomedical applications/Tissue engineering [177]
Spider silk + bone sialoprotein E. coli RY-3041 strain Biomedical applications/Tissue engineering [176]
Spider silk + antimicrobial domain (HNP-2, HNP-4 and hepcidin) E. coli RY-3041 strain Biomedical applications/Tissue engineering [178]
Bombyx mori silk + RGD + elastin (FES8) E. coli BL21 strain Biomedical applications [274]
RGDS + silk fibroin (RGDSx2 fibroin) Silkworm Facilitate chondrogenesis [275]
Collagen + GYIPEAPRDGQAYVRKDGEWVLLSTFL E. coli BL21 strain Stabilize the triple helix formed in the proteins/Biomedical applications [276]
BMP-2 + collagen-biding domain E. coli BL21 strain Bone repair [277279]
TGF-B1-F1 and TGF-B1-F2 + collagen binding domain E. coli Biomedical applications/Tissue engineering [280]
hbFGF-F1 and hbFGF-F2 + collagen binding domain E. coli Biomedical applications/Tissue engineering [281]
PDGF + collagen binding domain E. coli BL21 strain Tissue regeneration and wound repair [282]
Fibronectin III7–10 + cadherin 11 EC 1–2 E. coli Rosetta-gami strain Orthopaedic regeneration [283]
Fibronectin cell binding domain-EGF (C-EGF) E. coli HBIOI strain Drug delivery [180]
Fibronectin cell binding domain-EGF (FNCBD-EGF) E. coli Skin wounds, catheter-injured arteries, and hind limb muscles [284]
RGD/EGF/hydrophobic sequence E12 (ERE–EGF) E. coli Controlling cell functions [285]
NGF-β + collagen binding domain E. coli BL21 strain Delivery system for neuronal development and regeneration [286]

Below we will address some of the proteins that have been effectively cloned and expressed in different recombinant systems. The potential of genetic engineering to be used as a tool for the functionalization of biopolymers with different bioactive peptides through the synthesis of new fusion proteins will also be discussed.

Collagen has been cloned and expressed in recombinant systems (Table 3). The use of recombinant collagen has benefits since it can be a safe product with useful self-assembly features [144] and the possibility of being functionalized with bio-instructive domains [19] such as cell adhesion ligands [141]. Over the past 20 years recombinant systems for the large scale-production of recombinant collagen have been developed and optimized. Recombinant collagen has been expressed in mammalian cells, insect cells, Escherichia coli (E. coli), transgenic tobacco, mice and silkworm [145]. From these recombinant hosts only the mammalian cells expressed collagen with 4-hydroxyproline content identical to native collagen. However, since the level of protein production was low (0.6–20 mg/L) this system was not commercially viable [146]. Since the production cost in yeast and E. coli is much lower than in mammalian cell culture, a multigene expression technology was adopted in order to overcome the absence of the enzyme prolyl 4-hydroxylase, an essential element in the synthesis of fully hydroxylated collagens [146]. The absence of this enzyme leads to non-triple-helical and non-functional collagen molecules, which are unstable below physiological temperatures and thus unsuitable for medical applications. Hence, the multigene expression approach based on the co-expression of procollagen polypeptide chains and α- and β-subunits of proyl 4-hydroxylase using the yeast, Pichia pastoris, was developed [147]. Collagen types I, II and III were expressed with a 4-hydroxyproline content identical to the native human proteins and expression levels of 0.2 to 0.6 g/L in 2 L bioreactors were achieved [147]. The use of recombinant collagen as a gel has been reported for chondrocytes [148], as a microcarrier [149], as corneal substitutes [150] and for bone regeneration applications [151]. Furthermore, customized collagen-like peptides formed with tandem repeats of the D4 domain of human collagen type II, a critical sequence for supporting the migration of chondrocytes, were also reported [140]. Chondrocytes seeded on polyglycolic acid scaffolds coated with this collagen-like protein formed cartilaginous constructs with superior properties to the scaffolds coated with native type II collagen [152]. These advances highlight the importance of recombinant DNA technology in the synthesis of proteins with applications that until now have only been available from animal sources.

Recombinant DNA technology was particularly advantageous in the expression of large and repetitive proteins such as silk. As in the case of collagen, different expression hosts have been explored for the biosynthesis of spider silk (Table 3). Major ampullate silk was successfully expressed by bovine mammary epithelial cells, hamster kidney cells, insect cells and in the milk of transgenic goats, generally with low yields [153]. However, bacteria can be grown at large scales and have the advantage of being easier to handle and more cost-effective. Therefore, E. coli has been actively pursued as an expression host for spider silks. Since bacterial hosts have distinct codon usages, silk sequences from different spider species were reverse transcribed into cDNA, using the E. coli codon preferences, and double stranded oligonucleotides coding for different domains of silk proteins were prepared [154]. These double strand oligonucleotides were then assembled into synthetic genes coding for silk proteins [153]. This cloning strategy was employed with successes for the expression of Nephila clavipes consensus sequence for major ampullate silk protein 1 (MaSp1) and MaSp2 [155] and the flagelliform silk protein [156] from the same species. Cloning and expression in E. coli, of both major ampullate silks ADF-3 and ADF-4 from the species Araneus diadematus was also reported (Table 3) with yields between 140 and 360 mg/L [157]. Besides E. coli, other hosts for the cloning and expression of spider silks have also been explored. The yeast Pichia pastoris is considered an attractive host for the expression of recombinant proteins since this expression system is well developed for industrial fermentation, reaching high cell densities using low-cost media. For these reasons it was successfully used for the expression of spider silk dragline using genes of up to 3,000 codons with no evidence of truncated synthesis, a common occurrence in E. coli host [158]. Plants such as tobacco and Arabidopsis thaliana are also being explored as transgenic host systems for silk proteins, with yields of 2% in tobacco leaves, 8.5% in A. thaliana leaf apoplasts and 18% in the endoplasmic reticulum of seeds [153]. Similar approaches as above for collagens and silks have been applied to the fabrication of recombinant elastin-like proteins that mimic native elastin (Table 3) [56]. These new protein polymers have a modular structure formed with repeats of the pentapeptide (VP-Xaa-Yaa-G)n where Xaa is either G or A and Yaa can be any residue but P. These recombinant elastin-like proteins are capable of reversible temperature-dependant self assembly in aqueous medium [67]. This feature allows for the purification of protein based upon temperature-induced aggregation. Elastomeric pentapeptides with up to 251 GVGVP repeats were soluble in low ionic solution at temperatures below 25ºC [159]. Above this temperature the polymer hydrophobically folds into β-spiral structures that further aggregate due to hydrophobic associations. These aggregates can then be collected by selective centrifugation. This methodology allows for facile purification [160, 161]. Moreover, elastin-like polypeptides (ELPs) can be used as a purification tag. The fusion of ELPs with other proteins exploits the inverse temperature transition of ELPs and provides a simple method for the isolation of a recombinant ELP fusion proteins by cycling the protein solution through the soluble and insoluble phases using inverse transition cycling [162164]. ELP tags can be cleaved by a pH shift and removed by a final thermal precipitation [164].

Additionally there is the possibility of amino acid substitutions in the pentapeptide repeats [165] and previous studies have shown that the replacement of G in (VPGVG)n by A in (VPAVG)n leads to mechanical changes in the protein from elastic to plastic [67, 166]. The physical crosslinking resulting from this amino acid replacement leads to a more plastic matrix with a Young’s modulus two orders of magnitude higher than in the case of (VPGVG)n [166]. Also, physical crosslinking has advantages over chemical crosslinking since it allows for easy processing, avoids the use of chemical reagents and excludes the need of removing unreacted intermediates [167]. Synthetic amphiphilic block copolymers with distinct block polarity composed of hydrophilic and hydrophobic segments can also be generated [67]. These block copolymers exhibit tunable mechanical and amphiphilic properties dependent on the amino acid substitution. The flexibility of these block copolymer designs extends the range of applications from micelles formed by self-assembly of amphiphilic sequences for drug delivery, to temperature responsive hydrogels for cell encapsulation and coatings of medical devices to improve host responses [168170]. Genetic engineering also offers the possibility of enriching the sequences of proteins to improve their biological activity by fusing them with other protein motifs with specific bioactivities (Table 4). Initial elastin matrices for cell adhesion showed that cells did not adhere to these biomaterials [171]. RGD and REDV cell adhesion peptide sequences were inserted into the elastins leading to a dramatic increase in cell attachment [169, 170]. Silk-based block copolymers were also engineered to carry an RGD cell binding domain for intracellular gene delivery. The presence of labelled DNA inside cells was detected by confocal laser scanning microscopy and demonstrates the potential of these silk bioengineered block copolymers as highly tailored gene delivery systems [172]. The addition of a recognition site for an enzyme with proteolytic activity can also be incorporated into the sequences, favouring biomaterials degradation [173]. The fusion of the N. clavipes consensus sequence for MaSp1 with proteins such as dentin matrix protein and bone sialoprotein, involved in calcium phosphate deposition in teeth and bone [174, 175], respectively, also had positive results from a biomaterials perspective [176, 177]. In both fusion proteins the silk domain retained its self assembly properties and the dentin matrix protein and bone sialoprotein domains maintained their ability to induce the deposition of calcium phosphates. These results demonstrated the potential of chimeric proteins for applications in tissue engineering and regenerative medicine for the design of new protein-based scaffolds for bone regeneration [176, 177].

Furthermore, promising results were also obtained when the N. clavipes consensus sequence for MaSp1 was fused with antimicrobial peptides, namely neutrophil defensins 2 and 4 and hepcidin, using a step-by-step cloning methodology [178]. The cloning and expression of these new fusion proteins expanded these chimera or fusion approaches to include antimicrobial-functionalized protein-based biomaterials [178] offering a path forward in reducing the use of antibiotics to prevent infection in implants and in the design of a new generation of protein-based materials bioengineered to prevent the onset of infections.

Other proteins have also been expressed as fusion proteins with biological activity such as FGF2-FNIII9-10 formed by a fibronectin fragment FNIII9-10 connected to the carboxy terminus of fibroblast growth factor 2 (FGF-2) [179]. Previous studies reported the synergistic effect of fibronectin and FGF-2 on osteoblast adhesion. The FGF2-FNIII9-10 fusion protein showed a significant increase in cell adhesion and proliferation when compared with FNIII9-10 alone [179]. The cell-binding domain of human fibronectin was also fused with epidermal growth factor (EGF), important in tissue regeneration to accelerate wound healing and enhance cell proliferation. The new construct, designated as C-EGF, had both cell-adhesive and EGF activity and the recombinant construct may be an effective drug delivery system for EGF in therapeutic situations [180]. EGF polypeptide was fused with collagen type III and the new construct retained the triple helix of collagen and the mitogenic activity of EGF, suggesting that this protein could be used as a biocompatible, biodegradable and adhesive fibrous mitogen for tissue regeneration [181].

The examples outlined above highlight the potential of synthetic biology in the synthesis of biopolymers for tissue engineering and regenerative medicine (Figure 1).

Fig. 1.

Fig. 1

Scheme highlighting some of the features and applications of chimeric protein-based biomaterials synthesized through recombinant DNA technology.

5. Conclusions

Genetic engineering makes it possible to develop new biopolymers with a complexity and functionality resembling natural polymers formed in nature. By using synthetic DNA it is possible to combine different functional domains for a fusion protein, merging cell adhesion and migration, mechanical properties and antimicrobial factors, towards multifunctional biomaterial systems. This approach eliminates the need to use chemical methodologies for covalent binding of bioactive motifs or crosslinking, which can have drawbacks of protein denaturation and residuals with toxicity. Although there has been a significant progress in exploiting genetic engineering for tissue engineering and regenerative medicine purposes during recent years, there remains a lot to be explored in order to take full advantage of the outstanding potential of genetic engineering to be used as a tool in the development of the next generation of custom-design biomaterials.

Acknowledgments

Sílvia Gomes thanks the Portuguese Foundation for Science and Technology (FCT) for providing her a PhD grant (SFRH/BD/28603/2006). This work was carried out under the scope of the FIND & BIND project funded by the agency EU-EC (FP7 program), the FCT R&D project ProteoLight (PTDC/FIS/68517/2006) funded by the FCT agency, the Chimera project (PTDC/EBB-EBI/109093/2008) funded by the FCT agency, the NIH (P41 EB002520) Tissue Engineering Resource Center and the NIH (EB003210 and DE017207).

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Tsubouchi M, Matsui S, Banno Y, Kurokawa K, Kawakami K. Overview of the clinical application of regenerative medicine products in Japan. Health Policy. 2008;88:62–72. doi: 10.1016/j.healthpol.2008.02.011. [DOI] [PubMed] [Google Scholar]
  • 2.Laurencin CT, Ambrosio AMA, Borden MD, Cooper JA. Tissue engineering: orthopedic applications. Annu Rev Biomed Eng. 1999;1:19–46. doi: 10.1146/annurev.bioeng.1.1.19. [DOI] [PubMed] [Google Scholar]
  • 3.Salgado AJ, Coutinho OP, Reis RL. Bone tissue engineering: state of the art and future trends. Macromol Biosci. 2004;4:743–65. doi: 10.1002/mabi.200400026. [DOI] [PubMed] [Google Scholar]
  • 4.Younger EM, Chapman MW. Morbidity at bone graft donor sites. J Orthop Trauma. 1989;3:192–95. doi: 10.1097/00005131-198909000-00002. [DOI] [PubMed] [Google Scholar]
  • 5.Weir MD, Xu HH. Osteoblastic induction on calcium phosphate cement-chitosan constructs for bone tissue engineering. J Biomed Mater Res A. 2010;94A:223–33. doi: 10.1002/jbm.a.32665. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Wise DL, Trantolo DJ, Lewandrowski K-U, Gresser JD, Cattaneo MV, Yaszemski MJ. Biomaterials engineering and devices. Totowa NJ: Human Press; 2000. [Google Scholar]
  • 7.Gunatillake PA, Adhikari R. Biodegradable synthetic polymers for tissue engineering. Eur Cell Mater. 2003;5:1–16. doi: 10.22203/ecm.v005a01. [DOI] [PubMed] [Google Scholar]
  • 8.Sokolsky-Papkov M, Agashi K, Olaye A, Shakesheff K, Domb AJ. Polymer carriers for drug delivery in tissue engineering. Adv Drug Delivery Rev. 2007;59:187–206. doi: 10.1016/j.addr.2007.04.001. [DOI] [PubMed] [Google Scholar]
  • 9.Mark Kvd, Park J, Bauer S, Schmuki P. Nanoscale engineering of biomimetic surfaces: cues from the extracellular matrix. Cell Tissue Res. 2010;339:131–53. doi: 10.1007/s00441-009-0896-5. [DOI] [PubMed] [Google Scholar]
  • 10.Williams DJ, Sebastine IM. Tissue engineering and regenerative medicine: manufacturing challenges. IEE Proc Nanobiotechnol. 2005;152:207–10. doi: 10.1049/ip-nbt:20050001. [DOI] [PubMed] [Google Scholar]
  • 11.Biondi M, Ungaro F, Quaglia F, Netti PA. Controlled drug delivery in tissue engineering. Adv Drug Delivery Rev. 2008;60:229–42. doi: 10.1016/j.addr.2007.08.038. [DOI] [PubMed] [Google Scholar]
  • 12.George A, Ravindran S. Protein templates in hard tissue engineering. Nano Today. 2010;5:254–66. doi: 10.1016/j.nantod.2010.05.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Kohn J, Welsh WJ, Knight D. A new approach to the rationale discovery of polymeric biomaterials. Biomaterials. 2007;28:4171–77. doi: 10.1016/j.biomaterials.2007.06.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Bie CD. Genzyme: 15 years of cell and gene therapy research. Regen Med. 2007;2:95–97. doi: 10.2217/17460751.2.1.95. [DOI] [PubMed] [Google Scholar]
  • 15.Basad E, Ishaque B, Bachmann G, Stürz H, Steinmeyer J. Matrix-induced autologous chondrocyte implantation versus microfracture in the treatment of cartilage defects of the knee: a 2-year randomised study. Knee Surg Sports Traumatol Arthrosc. 2010;18:519–27. doi: 10.1007/s00167-009-1028-1. [DOI] [PubMed] [Google Scholar]
  • 16.Zaulyanov L, Kirsner RS. A review of a bi-layered living cell treatment (Apligraf) in the treatment of venous leg ulcers and diabetic foot ulcers. Clin Interv Aging. 2007;2:93–98. doi: 10.2147/ciia.2007.2.1.93. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Degim IT, Çelebi N. Controlled delivery of peptides and proteins. Curr Pharm Des. 2007;13:99–117. doi: 10.2174/138161207779313795. [DOI] [PubMed] [Google Scholar]
  • 18.Nagaoka M, Jiang HL, Hoshiba T, Akaike T, Cho CS. Application of recombinant fusion proteins for tissue engineering. Ann Biomed Eng. 2010;38:683–93. doi: 10.1007/s10439-010-9935-3. [DOI] [PubMed] [Google Scholar]
  • 19.Romano NH, Sengupta D, Chung C, Heilshorn SC. Protein-engineered biomaterials: nanoscale mimics of the extracellular matrix. Biochim Biophys Acta. 2010;1810:339–49. doi: 10.1016/j.bbagen.2010.07.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Zelzer M, Ulijn RV. Next-generation peptide nanomaterials: molecular networks, interfaces and supramolecular functionality. Chem Soc Rev. 2010;39:3351–57. doi: 10.1039/c0cs00035c. [DOI] [PubMed] [Google Scholar]
  • 21.Patterson J, Martino MM, Hubbell JA. Biomimetic materials in tissue engineering. Mater Today. 2010;13:14–22. [Google Scholar]
  • 22.Krishna OD, Kiick KL. Protein- and peptide-modified synthetic polymeric biomaterials. Biopolymers. 2010;94:32–48. doi: 10.1002/bip.21333. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Mano JF, Silva GA, Azevedo HS, Malafaya PB, Sousa RA, Silva SS, Boesel LF, Oliveira JM, Santos TC, Marques AP, Neves NM, Reis RL. Natural origin biodegradable systems in tissue engineering and regenerative medicine: present status and some moving trends. J R Soc Interface. 2007;4:999–1030. doi: 10.1098/rsif.2007.0220. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Silva SS, Mano JF, Reis RL. Potential applications of natural origin polymer-based systems in soft tissue regeneration. Crit Rev Biotechnol. 2010;30:200–21. doi: 10.3109/07388551.2010.505561. [DOI] [PubMed] [Google Scholar]
  • 25.Chung HJ, Park TG. Surface engineered and drug releasing prefabricated scaffolds for tissue engineering. Adv Drug Delivery Rev. 2007;59:249–62. doi: 10.1016/j.addr.2007.03.015. [DOI] [PubMed] [Google Scholar]
  • 26.Heim M, Römer L, Scheibel T. Hierarchical structures made of proteins. The complex architecture of spider webs and their constituent silk proteins. Chem Soc Rev. 2010;39:156–64. doi: 10.1039/b813273a. [DOI] [PubMed] [Google Scholar]
  • 27.Madsen K, Mark Kvd, Menxel Mv, Friberg U. Analysis of collagen types synthesized by rabbit ear cartilage chondrocytes in vivo and in vitro. Biochem J. 1984;221:189–96. doi: 10.1042/bj2210189. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Mansell JP, Yarram SJ, Brown NL, Sandy JR. Type 1 collagen synthesis by human osteoblasts in response to placental lactogen and chaperonin 10, a homolog of early-pregnancy factor. In Vitro Cell Dev Biol Anim. 2002;38:518–22. doi: 10.1290/1071-2690(2002)038<0518:ticsbh>2.0.co;2. [DOI] [PubMed] [Google Scholar]
  • 29.Harkness RD. Biological functions of collagen. Biol Rev. 1961;36:399–463. doi: 10.1111/j.1469-185x.1961.tb01596.x. [DOI] [PubMed] [Google Scholar]
  • 30.Lee CH, Singla A, Lee Y. Biomedical applications of collagen. Int J Pharm. 2001;221:1–22. doi: 10.1016/s0378-5173(01)00691-3. [DOI] [PubMed] [Google Scholar]
  • 31.Kadler KE, Baldock C, Bella J, Boot-Handford RP. Collagens at a glance. J Cell Sci. 2007;120:1955–58. doi: 10.1242/jcs.03453. [DOI] [PubMed] [Google Scholar]
  • 32.Veit G, Kobbe B, Keene DR, Paulsson M, Koch M, Wagener R. Collagen XXVIII, a novel von Willebrand factor A domain containing protein with many imperfections in the collagenous domain. J Biol Chem. 2006;81:3494–504. doi: 10.1074/jbc.M509333200. [DOI] [PubMed] [Google Scholar]
  • 33.Heino J, Huhtala M, Käpylä J, Johnson MS. Evolution of collagen-based adhesion systems. Int J Biochem Cell Biol. 2009;41:341–48. doi: 10.1016/j.biocel.2008.08.021. [DOI] [PubMed] [Google Scholar]
  • 34.Shoulders MD, Raines RT. Collagen structure and stability. Annu Rev Biochem. 2009;78:929–58. doi: 10.1146/annurev.biochem.77.032207.120833. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Hubbell JA. Materials as morphogenetic guides in tissue engineering. Curr Opin Biotechnol. 2003;14:551–58. doi: 10.1016/j.copbio.2003.09.004. [DOI] [PubMed] [Google Scholar]
  • 36.Pabbruwe MB, Kafienah W, Tarlton JF, Mistry S, Fox DJ, Hollander AP. Repair next term of meniscal previous termcartilagenext term white zone tears using a stem cell/previous termcollagennext term-scaffold implant. Biomaterials. 2010;31:2583–91. doi: 10.1016/j.biomaterials.2009.12.023. [DOI] [PubMed] [Google Scholar]
  • 37.Park IS, Kim SH, Kim YH, Kim IH, Kim SH. A collagen/smooth muscle cell-incorporated elastic scaffold for tissue-engineered vascular grafts. J Biomater Sci Polym Ed. 2009;20:1645–60. doi: 10.1163/156856208X386237. [DOI] [PubMed] [Google Scholar]
  • 38.Helary C, Bataille I, Abed A, Illoul C, Anglo A, Louedec L, Letourneur D, Meddahi-Pellé A, Giraud-Guille MM. Concentrated collagen hydrogels as dermal substitutes. Biomaterials. 2010;31:481–90. doi: 10.1016/j.biomaterials.2009.09.073. [DOI] [PubMed] [Google Scholar]
  • 39.Wehrhan F, Nkenke E, Melnychenko I, Amann K, Schlegel KA, Goerlach C, Zimmermann WH, Schultze-Mosgau S. Skin repair using a porcine collagen I/III membrane--vascularization and epithelization properties. Dermatol Surg. 2010;36:919–30. doi: 10.1111/j.1524-4725.2010.01569.x. [DOI] [PubMed] [Google Scholar]
  • 40.Lyons FG, Al-Munajjed AA, Kieran SM, Toner ME, Murphy CM, Duffy GP, O’Brien FJ. The healing of bony defects by cell-free collagen-based scaffolds compared to stem cell-seeded tissue engineered constructs. Biomaterials. 2010;31:9232–43. doi: 10.1016/j.biomaterials.2010.08.056. [DOI] [PubMed] [Google Scholar]
  • 41.Kleinmann G, Larson S, Hunter B, Stevens S, Mamalis N, Olson RJ. Collagen shields as a drug delivery system for the fourth-generation fluoroquinolones. Ophthalmologica. 2007;221:51–56. doi: 10.1159/000096523. [DOI] [PubMed] [Google Scholar]
  • 42.Maeda M, Kadota K, Kajihara M, Suno A, Fujioka K. Sustained release of human growth hormone (hGH) from collagen film and evaluation of effect on wound healing in db/db mice. J Controlled Release. 2001;77:261–72. doi: 10.1016/s0168-3659(01)00512-0. [DOI] [PubMed] [Google Scholar]
  • 43.Chan BP, Hui TY, Wong MY, Yip KHK, Chan GCF. Mesenchymal stem cell-encapsulated collagen microspheres for bone tissue engineering. Tissue Eng Part C. 2010;16:225–35. doi: 10.1089/ten.tec.2008.0709. [DOI] [PubMed] [Google Scholar]
  • 44.Chan OCM, So K-F, Chan BP. Fabrication of nano-fibrous collagen microspheres for protein delivery and effects of photochemical crosslinking on release kinetics. J Controlled Release. 2008;129:135–43. doi: 10.1016/j.jconrel.2008.04.011. [DOI] [PubMed] [Google Scholar]
  • 45.Kawazoe N, Inoue C, Tateishi T, Chen G. A cell leakproof PLGA-collagen hybrid scaffold for cartilage tissue engineering. Biotechnol Prog. 2010;26:819–26. doi: 10.1002/btpr.375. [DOI] [PubMed] [Google Scholar]
  • 46.Tierney CM, Haugh MG, Liedl J, Mulcahy F, Hayes B, O’Brien FJ. The effects of collagen concentration and crosslink density on the biological, structural and mechanical properties of collagen-GAG scaffolds for bone tissue engineering. J Mech Behav Biomed Mater. 2009;2:202–09. doi: 10.1016/j.jmbbm.2008.08.007. [DOI] [PubMed] [Google Scholar]
  • 47.Faralli JA, Schwinn MK, Gonzalez JM, Filla MS, Peters DM. Functional properties of fibronectin in the trabecular meshwork. Exp Eye Res. 2009;88:689–93. doi: 10.1016/j.exer.2008.08.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Mao Y, Schwarzbauer JE. Fibronectin fibrillogenesis, a cell-mediated matrix assembly process. Matrix Biology. 2005;24:389–99. doi: 10.1016/j.matbio.2005.06.008. [DOI] [PubMed] [Google Scholar]
  • 49.Pankov R, Yamada KM. Fibronectin at a glance. J Cell Sci. 2002;115:3861–63. doi: 10.1242/jcs.00059. [DOI] [PubMed] [Google Scholar]
  • 50.Rexeisen EL, Fan W, Pangburn TO, Taribagil RR, Bates FS, Lodge TP, Tsapatsis M, Kokkoli E. Self-Assembly of fibronectin mimetic peptide-amphiphile nanofibers. Langmuir. 2009;26:1953–59. doi: 10.1021/la902571q. [DOI] [PubMed] [Google Scholar]
  • 51.Amaral IF, Unger RE, Fuchs S, Mendonça AM, Sousa SR, Barbosa MA, Pêgo AP, Kirkpatrick CJ. Fibronectin-mediated endothelialisation of chitosan porous matrices. Biomaterials. 2009;30:5465–75. doi: 10.1016/j.biomaterials.2009.06.056. [DOI] [PubMed] [Google Scholar]
  • 52.Custódio CA, Alves CM, Reis RL, Mano JF. Immobilization of fibronectin in chitosan substrates improves cell adhesion and proliferation. J Tissue Eng Regen Med. 2010;4:316–23. doi: 10.1002/term.248. [DOI] [PubMed] [Google Scholar]
  • 53.Bush KA, Pins GD. Carbodiimide conjugation of fibronectin on collagen basal lamina analogs enhances cellular binding domains and epithelialization. Tissue Eng Part A. 2010;16:829–38. doi: 10.1089/ten.tea.2009.0514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Barbucci R, Magnani A, Chiumiento A, Pasqui D, Cangioli I, Lamponi S. Fibroblast cell behavior on bound and adsorbed fibronectin onto hyaluronan and sulfated hyaluronan substrates. Biomacromolecules. 2005;6:638–45. doi: 10.1021/bm049642v. [DOI] [PubMed] [Google Scholar]
  • 55.Wittmer CR, Phelps JA, Saltzman WM, Tassel PRV. Fibronectin terminated multilayer films: protein adsorption and cell attachment studies. Biomaterials. 2007;28:851–60. doi: 10.1016/j.biomaterials.2006.09.037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Almine JF, Bax DV, Mithieux SM, Nivison-Smith L, Rnjak J, Waterhouse A, Wise SG, Weiss AS. Elastin-based materials. Chem Soc Rev. 2010;39:3371–79. doi: 10.1039/b919452p. [DOI] [PubMed] [Google Scholar]
  • 57.Debelle L, Tamburro AM. Elastin: molecular description and function. Int J Biochem Cell Biol. 1999;31:261–72. doi: 10.1016/s1357-2725(98)00098-3. [DOI] [PubMed] [Google Scholar]
  • 58.Muiznieks LD, Weiss AS, Keeley FW. Structural disorder and dynamics of elastin. Biochem Cell Biol. 2010;88:239–50. doi: 10.1139/o09-161. [DOI] [PubMed] [Google Scholar]
  • 59.Rodgers URS, Weiss A. Integrin αvβ3 binds a unique non-RGD site near the C-terminus of human tropoelastin. Biochimie. 2004;86:173–78. doi: 10.1016/j.biochi.2004.03.002. [DOI] [PubMed] [Google Scholar]
  • 60.Sionkowska A, Skopinska-Wisniewska J, Gawron MJ, Kozlowska AP. Chemical and thermal cross-linking of collagen and elastin hydrolysates. Int J Biol Macromol. 2010;47:570–77. doi: 10.1016/j.ijbiomac.2010.08.004. [DOI] [PubMed] [Google Scholar]
  • 61.McClure MJ, Sell SA, Simpson DG, Walpoth BH, Bowlin GL. A three-layered electrospun matrix to mimic native arterial architecture using polycaprolactone, elastin, and collagen: a preliminary study. Acta Biomater. 2010;6:2422–33. doi: 10.1016/j.actbio.2009.12.029. [DOI] [PubMed] [Google Scholar]
  • 62.Hu X, Wang X, Rnjak J, Weiss AS, Kaplan DL. Biomaterials derived from silk-tropoelastin protein systems. Biomaterials. 2010;31:8121–31. doi: 10.1016/j.biomaterials.2010.07.044. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Wise SG, Byrom MJ, Waterhouse A, Bannon PG, Ng MK, Weiss AS. A multilayered synthetic human elastin/polycaprolactone hybrid vascular graft with tailored mechanical properties. Acta Biomater. 2010;7:295–303. doi: 10.1016/j.actbio.2010.07.022. [DOI] [PubMed] [Google Scholar]
  • 64.Dinerman AA, Cappello J, El-Sayed M, Hoag SW, Ghandehari H. Influence of Solute Charge and Hydrophobicity on Partitioning and Diffusion in a Genetically Engineered Silk-Elastin-Like Protein Polymer Hydrogel. Macromol Biosci. 2010;10:1235–47. doi: 10.1002/mabi.201000061. [DOI] [PubMed] [Google Scholar]
  • 65.Rocha LB, Adam RL, Leite NJ, Metze K, Rossi MA. Biomineralization of polyanionic collagen–elastin matrices during cavarial bone repair. J Biomed Mater Res A. 2006;79:237–45. doi: 10.1002/jbm.a.30782. [DOI] [PubMed] [Google Scholar]
  • 66.Bessa PC, Machado R, Nürnberger S, Dopler D, Banerjee A, Cunha AM, Rodríguez-Cabello JC, Redl H, Griensven Mv, Reis RL, Casal M. Thermoresponsive self-assembled elastin-based nanoparticles for delivery of BMPs. J Controlled Release. 2010;142:312–18. doi: 10.1016/j.jconrel.2009.11.003. [DOI] [PubMed] [Google Scholar]
  • 67.Kim W, Chaikof EL. Recombinant elastin-mimetic biomaterials: emerging applications in medicine. Adv Drug Delivery Rev. 2010;62:1468–78. doi: 10.1016/j.addr.2010.04.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Laura Martín MA, Alessandra Girotti F, Arias Javier, Carlos Rodríguez-Cabello J. Synthesis and characterization of macroporous thermosensitive hydrogels from recombinant elastin-like polymers. Biomacromolecules. 2009;10:3015–22. doi: 10.1021/bm900560a. [DOI] [PubMed] [Google Scholar]
  • 69.Maskarinec SA, Tirrell DA. Protein engineering approaches to biomaterials design. Curr Opin Biotechnol. 2005;16:422–26. doi: 10.1016/j.copbio.2005.06.009. [DOI] [PubMed] [Google Scholar]
  • 70.Smith GF. Fibrinogen–fibrin conversion. The mechanism of fibrin-polymer formation in solution. Biochem J. 1980;185:1–11. doi: 10.1042/bj1850001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Blomäck B, Hessel B, Hogg D. Disulfide bridges in NH2-terminal part of human fibrinogen. Thromb Res. 1976;8:639–58. doi: 10.1016/0049-3848(76)90245-0. [DOI] [PubMed] [Google Scholar]
  • 72.Mosesson MW, Siebenlist KR, Meh DA. The structure and biological features of fibrinogen and fibrin. Ann NY Acad Sci. 2001;936:11–30. doi: 10.1111/j.1749-6632.2001.tb03491.x. [DOI] [PubMed] [Google Scholar]
  • 73.Siebenlist KR, Diorio JP, Budzynski AZ, Mosesson MW. The polymerization and thrombin-binding properties of des-(Bβ1–42) fibrin. J Biol Chem. 1990;265:18650–55. [PubMed] [Google Scholar]
  • 74.Ferry JD, Morrison PR. Preparation and properties of serum and plasma proteins. VIII. The conversion of human fibrinogen to fibrin under various conditions. J Am Chem Soc. 1947;69:388–400. doi: 10.1021/ja01194a066. [DOI] [PubMed] [Google Scholar]
  • 75.Lorand L, Graham RM. Transglutaminases: crosslinking enzymes with pleiotropic functions. Nat Rev Mol Cell Biol. 2003;4:140–56. doi: 10.1038/nrm1014. [DOI] [PubMed] [Google Scholar]
  • 76.Spotnitz WD. Fibrin sealant: past, present, and future: a brief review. World J Surg. 2010;34:632–34. doi: 10.1007/s00268-009-0252-7. [DOI] [PubMed] [Google Scholar]
  • 77.Ahmann KA, Weinbaum JS, Johnson SL, Tranquillo RT. Fibrin degradation enhances vascular smooth muscle cell proliferation and matrix deposition in fibrin-based tissue constructs fabricated in vitro. Tissue Eng Part A. 2010;16:3261–70. doi: 10.1089/ten.tea.2009.0708. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Lee Y-B, Polio S, Lee W, Dai G, Menon L, Carroll RS, Yoo S-S. Bio-printing of collagen and VEGF-releasing fibrin gel scaffolds for neural stem cell culture. Exp Neurol. 2010;223:645–52. doi: 10.1016/j.expneurol.2010.02.014. [DOI] [PubMed] [Google Scholar]
  • 79.Kang S-W, Kim i-S, Park K-S, Cha B-H, Shim J-H, Kim JY, Cho D-W, Rhie J-W, Lee S-H. Surface modification with fibrin/hyaluronic acid hydrogel on solid-free form-based scaffolds followed by BMP-2 loading to enhance bone regeneration. Bone. 2010;48:298–306. doi: 10.1016/j.bone.2010.09.029. [DOI] [PubMed] [Google Scholar]
  • 80.Durbeej M. Laminins. Cell Tissue Res. 2010;339:259–68. doi: 10.1007/s00441-009-0838-2. [DOI] [PubMed] [Google Scholar]
  • 81.Kikkawa Y, Takahashi N, Matsuda Y, Miwa T, Akizuki T, Kataoka A, Nomizu M. The influence of synthetic peptides derived from the laminin α1 chain on hepatocyte adhesion and gene expression. Biomaterials. 2009;30:6888–95. doi: 10.1016/j.biomaterials.2009.09.011. [DOI] [PubMed] [Google Scholar]
  • 82.Hozumi K, Akizuki T, Yamada Y, Hara T, Urushibata S, Katagiri F, Kikkawa Y, Nomizu M. Cell adhesive peptide screening of the mouse laminin α1 chain G domain. Arch Biochem Biophys. 2010;503:213–22. doi: 10.1016/j.abb.2010.08.012. [DOI] [PubMed] [Google Scholar]
  • 83.Sarfati G, Dvir T, Elkabets M, Apte RN, Cohen S. Targeting of polymeric nanoparticles to lung metastases by surface-attachment of YIGSR peptide from laminin. Biomaterials. 2010;32:152–61. doi: 10.1016/j.biomaterials.2010.09.014. [DOI] [PubMed] [Google Scholar]
  • 84.Schvartz I, Seger D, Shaltiel S. Vitronectin. Int J Biochem Cell Biol. 1999;31:539–44. doi: 10.1016/s1357-2725(99)00005-9. [DOI] [PubMed] [Google Scholar]
  • 85.Schleicher I, Parker A, Leavesley D, Crawford R, Upton Z, Xiao Y. Surface modification by complexes of vitronectin and growth factors for serum-free culture of human osteoblasts. Tissue Eng. 2005;11:1688–98. doi: 10.1089/ten.2005.11.1688. [DOI] [PubMed] [Google Scholar]
  • 86.Steele JG, Johnson G, Underwood PA. Role of serum vitronectin and fibronectin in adhesion of fibroblasts following seeding onto tissue culture polystyrene. J Biomed Mater Res A. 1992;26:861–84. doi: 10.1002/jbm.820260704. [DOI] [PubMed] [Google Scholar]
  • 87.Reichl S. Films based on human hair keratin as substrates for cell culture and tissue engineering. Biomaterials. 2009;30:6854–66. doi: 10.1016/j.biomaterials.2009.08.051. [DOI] [PubMed] [Google Scholar]
  • 88.Hill P, Brantley H, Dyke MV. Some properties of keratin biomaterials: kerateines. Biomaterials. 2010;31:585–93. doi: 10.1016/j.biomaterials.2009.09.076. [DOI] [PubMed] [Google Scholar]
  • 89.Yamauchi K, Maniwa M, Mori T. Cultivation of fibroblast cells on keratin-coated substrata. J Biomater Sci Polym Ed. 1998;9:259–70. doi: 10.1163/156856298x00640. [DOI] [PubMed] [Google Scholar]
  • 90.Tachibana A, Kaneko S, Tanabe T, Yamauchi K. Rapid fabrication of keratin–hydroxyapatite hybrid sponges toward osteoblast cultivation and differentiation. Biomaterials. 2005;26:297–302. doi: 10.1016/j.biomaterials.2004.02.032. [DOI] [PubMed] [Google Scholar]
  • 91.Fujii T, Murai S, Ohkawa K, Hirai T. Effects of human hair and nail proteins and their films on rat mast cells. J Mater Sci: Mater Med. 2008;19:2335–42. doi: 10.1007/s10856-007-3341-x. [DOI] [PubMed] [Google Scholar]
  • 92.Sutherland TD, Young JH, Weisman S, Hayashi CY, Merritt DJ. Insect silk: one name, many materials. Annu Rev Entomol. 2010;55:171–88. doi: 10.1146/annurev-ento-112408-085401. [DOI] [PubMed] [Google Scholar]
  • 93.Omenetto FG, Kaplan DL. New opportunities for an ancient material. Science. 2010;329:528–31. doi: 10.1126/science.1188936. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Sehnal F, Žurovec M. Construction of silk fiber core in Lepidoptera. Biomacromolecules. 2004;5:666–74. doi: 10.1021/bm0344046. [DOI] [PubMed] [Google Scholar]
  • 95.Mahoney DV, Vezie DL, Eby RK, Adams WW, Kaplan D. Aspects of the morphology of dragline silk of Nephila clavipes. In: Kaplan D, Adams WW, Farmer B, Viney C, editors. Silk Polymers ACS Symposium Series. Vol. 544. Washington DC: American Chemical Society; 1997. pp. 196–210. [Google Scholar]
  • 96.Sponner A, Vater W, Monajembashi S, Unger E, Grosse F, Weisshart K. Composition and hierarchical organisation of a spider silk. PLoS ONE. 2007;2:e998/1–8. doi: 10.1371/journal.pone.0000998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Beek JDv, Hess S, Vollrath F, Meier BH. The molecular structure of spider dragline silk: folding and orientation of the protein backbone. Proc Natl Acad Sci USA. 2002;99:10266–71. doi: 10.1073/pnas.152162299. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Hardy JG, Scheibel TR. Silk-inspired polymers and proteins. Biochem Soc Trans. 2009;37:677–81. doi: 10.1042/BST0370677. [DOI] [PubMed] [Google Scholar]
  • 99.Gil ES, Mandal BB, Park S-H, Marchant JK, Omenetto FG, Kaplan DL. Helicoidal multi-lamellar features of RGD-functionalized silk biomaterials for corneal tissue engineering. Biomaterials. 2010;31:8953–63. doi: 10.1016/j.biomaterials.2010.08.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Lawrence BD, Marchant JK, Pindrus MA, Omenetto FG, Kaplan DL. Silk film biomaterials for cornea tissue engineering. Biomaterials. 2009;30:1299–308. doi: 10.1016/j.biomaterials.2008.11.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Chao P-HG, Yodmuang S, Wang X, Sun L, Kaplan DL, Vunjak-Novakovic G. Silk hydrogel for cartilage tissue engineering. J Biomed Mater Res B. 2010;95:84–90. doi: 10.1002/jbm.b.31686. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Wang Y, Bella E, Lee CSD, Migliaresi C, Pelcastre L, Schwartz Z, Boyan BD, Motta A. The synergistic effects of 3-D porous silk fibroin matrix scaffold properties and hydrodynamic environment in cartilage tissue regeneration. Biomaterials. 2010;31:4672–81. doi: 10.1016/j.biomaterials.2010.02.006. [DOI] [PubMed] [Google Scholar]
  • 103.Soffer L, Wang X, Zhang X, Kluge J, Dorfmann L, Kaplan DL, Leisk G. Silk-based electrospun tubular scaffolds for tissue-engineered vascular grafts. J Biomater Sci Polym Ed. 2008;19:653–64. doi: 10.1163/156856208784089607. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Zhou J, Cao C, Xilan Ma JL. Electrospinning of silk fibroin and collagen for vascular tissue engineering. Int J Biol Macromol. 2010;47:514–19. doi: 10.1016/j.ijbiomac.2010.07.010. [DOI] [PubMed] [Google Scholar]
  • 105.Kim HJ, Kim U-J, Kim HS, Li C, Wada M, Leisk GG, Kaplan DL. Bone tissue engineering with premineralized silk scaffolds. Bone. 2008;42:1226–34. doi: 10.1016/j.bone.2008.02.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Sofia S, McCarthy MB, Gronowicz G, Kaplan DL. Functionalized silk-based biomaterials for bone formation. J Biomed Mater Res. 2001;54:139–48. doi: 10.1002/1097-4636(200101)54:1<139::aid-jbm17>3.0.co;2-7. [DOI] [PubMed] [Google Scholar]
  • 107.Lammel AS, Hu X, Park S-H, Kaplan DL, Scheibel TR. Controlling silk fibroin particle features for drug delivery. Biomaterials. 2010;31:4583–91. doi: 10.1016/j.biomaterials.2010.02.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Uebersax L, Merkle HP, Meinel L. Insulin-like growth factor I releasing silk fibroin scaffolds induce chondrogenic differentiation of human mesenchymal stem cells. J Controlled Release. 2008;127:12–21. doi: 10.1016/j.jconrel.2007.11.006. [DOI] [PubMed] [Google Scholar]
  • 109.Spiess K, Lammel A, Scheibel T. Recombinant spider silk proteins for applications in biomaterial. Macromol Biosci. 2010;10:998–1007. doi: 10.1002/mabi.201000071. [DOI] [PubMed] [Google Scholar]
  • 110.Kluge JA, Rabotyagova O, Leisk GG, Kaplan DL. Spider silks and their applications. Trends Biotechnol. 2008;26:244–51. doi: 10.1016/j.tibtech.2008.02.006. [DOI] [PubMed] [Google Scholar]
  • 111.Widhe M, Bysell H, Nystedt S, Schenning I, Malmsten M, Johansson J, Rising A, Hedhammar M. Recombinant spider silk as matrices for cell culture. Biomaterials. 2010;31:9575–85. doi: 10.1016/j.biomaterials.2010.08.061. [DOI] [PubMed] [Google Scholar]
  • 112.Allmeling C, Jokuszies A, Reimers K, Kall S, Vogt PM. Use of spider silk fibres as an innovative material in a biocompatible artificial nerve conduit. J Cell Mol Med. 2006;10:770–77. doi: 10.1111/j.1582-4934.2006.tb00436.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Allmeling C, Jokuszies A, Reimers K, Kall S, Choi CY, Brandes G, Kasper C, Scheper T, Guggenheim M, Vogt PM. Spider silk fibres in artificial nerve constructs promote peripheral nerve regeneration. Cell Prolif. 2008;41:408–20. doi: 10.1111/j.1365-2184.2008.00534.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Baoyong L, Jian Z, Denglong C, Min L. Evaluation of a new type of wound dressing made from recombinant spider silk protein using rat models. Burns. 2010;36:891–96. doi: 10.1016/j.burns.2009.12.001. [DOI] [PubMed] [Google Scholar]
  • 115.Waite JH. The DOPA ephemera: a recurrent motif in invertebrates. Biol Bull. 1992;183:178–84. doi: 10.2307/1542421. [DOI] [PubMed] [Google Scholar]
  • 116.Deming TJ. Mussel byssus and biomolecular materials. Curr Opin Chem Biol. 1999;3:100–05. doi: 10.1016/s1367-5931(99)80018-0. [DOI] [PubMed] [Google Scholar]
  • 117.Hwang DS, Sim SB, Cha HJ. Cell adhesion biomaterial based on mussel adhesive protein fused with RGD peptide. Biomaterials. 2007;28:4039–46. doi: 10.1016/j.biomaterials.2007.05.028. [DOI] [PubMed] [Google Scholar]
  • 118.Lim S, Choi YS, Kang DG, Song YH, Cha HJ. The adhesive properties of coacervated recombinant hybrid mussel adhesive proteins. Biomaterials. 2010;31:3715–22. doi: 10.1016/j.biomaterials.2010.01.063. [DOI] [PubMed] [Google Scholar]
  • 119.Murphy JL, Vollenweider L, Xu F, Lee BP. Adhesive performance of biomimetic adhesive-coated biologic scaffolds. Biomacromolecules. 2010;11:2976–84. doi: 10.1021/bm1007794. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Beutner R, Michael J, Schwenzer B, Scharnweber D. Biological nano-functionalization of titanium-based biomaterial surfaces: a flexible toolbox. J R Soc Interface. 2010;7:S93–S105. doi: 10.1098/rsif.2009.0418.focus. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Bačáková L, Filová E, Rypáček F, Švorčík V, Starý V. Cell adhesion on artificial materials for tissue engineering. Physiol Res. 2004;53:S35–S45. [PubMed] [Google Scholar]
  • 122.Friess W, Uludag H, Foskett S, Biron R, Sargeant C. Characterization of absorbable collagen sponges as recombinant human bone morphogenetic protein-2 carriers. Int J Pharm. 1999;185:51–60. doi: 10.1016/s0378-5173(99)00128-3. [DOI] [PubMed] [Google Scholar]
  • 123.Karageorgiou V, Tomkins M, Fajardo R, Meinel L, Snyder B, Wade K, Chen J, Vunjak-Novakovic G, Kaplan DL. Porous silk fibroin 3-D scaffolds for delivery of bone morphogenetic protein-2 in vitro and in vivo. J Biomed Mater Res A. 2006;78:324–34. doi: 10.1002/jbm.a.30728. [DOI] [PubMed] [Google Scholar]
  • 124.Wongpanit P, Ueda H, Tabata Y, Rujiravanit R. In vitro and in vivo release of basic fibroblast growth factor using a silk fibroin scaffold as delivery carrier. J Biomater Sci Polym Ed. 2010;21:1403–19. doi: 10.1163/092050609X12517858243706. [DOI] [PubMed] [Google Scholar]
  • 125.Kleinheinz J, Jung S, Wermker K, Fischer C, Joos U. Release kinetics of VEGF165 from a collagen matrix and structural matrix changes in a circulation model. Head Face Med. 2010;6:1–7. doi: 10.1186/1746-160X-6-17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Kilian O, Hossain H, Flesch I, Sommer U, Nolting H, Chakraborty T, Schnettler R. Elution kinetics, antimicrobial efficacy, and degradation and microvasculature of a new gentamicin-loaded collagen fleece. J Biomed Mater Res B. 2008;90:210–22. doi: 10.1002/jbm.b.31275. [DOI] [PubMed] [Google Scholar]
  • 127.Wang X, Zhang X, Castellot J, Herman I, Iafrati M, Kaplan DL. Controlled release from multilayer silk biomaterial coatings to modulate vascular cell responses. Biomaterials. 2008;29:894–903. doi: 10.1016/j.biomaterials.2007.10.055. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Zhao J, Shinkai M, Takezawa T, Ohba S, Chung U-i, Nagamune T. Bone regeneration using collagen type I vitrigel with bone morphogenetic protein-2. J Biosci Bioeng. 2009;107:318–23. doi: 10.1016/j.jbiosc.2008.10.007. [DOI] [PubMed] [Google Scholar]
  • 129.Han D, Liu W, Ao Q, Wang G. Optimal delivery systems for bone morphogenetic proteins in orthopedic applications should model initial tissue repair structures by using a heparin-incorporated fibrin–fibronectin matrix. Med Hypotheses. 2008;71:374–78. doi: 10.1016/j.mehy.2008.01.035. [DOI] [PubMed] [Google Scholar]
  • 130.Shields LBE, Raque GHD, Glassman S, Campbell M, Vitaz T, Harpring J, Shields CB. Adverse effects associated with high-dose recombinant human bone morphogenetic protein-2 use in anterior cervical spine fusion. Spine. 2006;31:542–47. doi: 10.1097/01.brs.0000201424.27509.72. [DOI] [PubMed] [Google Scholar]
  • 131.Hersel U, Dahmen C, Kessler H. RGD modified polymers: biomaterials for stimulated cell adhesion and beyond. Biomaterials. 2003;24:4385–415. doi: 10.1016/s0142-9612(03)00343-0. [DOI] [PubMed] [Google Scholar]
  • 132.Murphy AR, Kaplan DL. Biomedical applications of chemically-modified silk fibroin. J Mater Chem. 2009;19:6443–50. doi: 10.1039/b905802h. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Rafat M, Li F, Fagerholm P, Lagali NS, Watsky MA, Munger R, Matsuura T, Griffith M. PEG-stabilized carbodiimide crosslinked collagen–chitosan hydrogels for corneal tissue engineering. Biomaterials. 2008;29:3960–72. doi: 10.1016/j.biomaterials.2008.06.017. [DOI] [PubMed] [Google Scholar]
  • 134.Zhang Y-Q, Ma Y, Xia Y-Y, Shen W-D, Mao J-P, Zha X-M, Shirai K, Kiguchi K. Synthesis of silk fibroin-insulin bioconjugates and their characterization and activities in vivo. J Biomed Mater Res B. 2006;79:275–83. doi: 10.1002/jbm.b.30539. [DOI] [PubMed] [Google Scholar]
  • 135.Chatterjee S, Barbora L, Cameotra SS, Mahanta P, Goswami P. Silk-fiber immobilized lipase-catalyzed hydrolysis of emulsified sunflower oil. Appl Biochem Biotechnol. 2009;157:593–600. doi: 10.1007/s12010-008-8405-y. [DOI] [PubMed] [Google Scholar]
  • 136.Yeo S, Oh J-E, Jeong L, Lee TS, Lee SJ, Park WH, Min B-M. Collagen-based biomimetic nanofibrous scaffolds: preparation and characterization of collagen/silk fibroin bicomponent nanofibrous structures. Biomacromolecules. 2008;9:1106–16. doi: 10.1021/bm700875a. [DOI] [PubMed] [Google Scholar]
  • 137.Kimura Y, Tsuji W, Yamashiro H, Toi M, Inamoto T, Tabata Y. In situ adipogenesis in fat tissue augmented by collagen scaffold with gelatin microspheres containing basic fibroblast growth factor. J Tissue Eng Regen Med. 2010;4:55–61. doi: 10.1002/term.218. [DOI] [PubMed] [Google Scholar]
  • 138.Adhirajan N, Shanmugasundaram N, Shanmuganathan S, Babu M. Functionally modified gelatin microspheres impregnated collagen scaffold as novel wound dressing to attenuate the proteases and bacterial growth. Eur J Pharm Sci. 2009;36:235–45. doi: 10.1016/j.ejps.2008.09.010. [DOI] [PubMed] [Google Scholar]
  • 139.Wang X, Wenk E, Matsumoto A, Meinel L, Li C, Kaplan DL. Silk microspheres for encapsulation and controlled release. J Controlled Release. 2007;117:360–70. doi: 10.1016/j.jconrel.2006.11.021. [DOI] [PubMed] [Google Scholar]
  • 140.Wu F, Jin T. Polymer-based sustained-release dosage forms for protein drugs, challenges, and recent advances. AAPS PharmSciTech. 2009;9:1218–29. doi: 10.1208/s12249-008-9148-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Langer R, Tirrell DA. Designing materials for biology and medicine. Nature. 2004;428:487–92. doi: 10.1038/nature02388. [DOI] [PubMed] [Google Scholar]
  • 142.Sato AK, Viswanathan M, Kent RB, Wood CR. Therapeutic peptides: technological advances driving peptides into development. Curr Opin Biotechnol. 2006;17:638–42. doi: 10.1016/j.copbio.2006.10.002. [DOI] [PubMed] [Google Scholar]
  • 143.Kastin AJ. Handbook of biologivally avtive peptides. Burlington: Elsevier; 2006. [Google Scholar]
  • 144.Kyle S, Aggeli A, Ingham E, McPherson MJ. Production of self-assembling biomaterials for tissue engineering. Trends Biotechnol. 2009;27:423–33. doi: 10.1016/j.tibtech.2009.04.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Báez J, Olsen D, Polarek JW. Recombinant microbial systems for the production of human collagen and gelatin. Appl Microbiol Biotechnol. 2005;69:245–52. doi: 10.1007/s00253-005-0180-x. [DOI] [PubMed] [Google Scholar]
  • 146.Olsen D, Yang C, Bodo M, Chang R, Leigh S, Baez J, Carmichael D, Perälä M, Eija-Riitt, Hämäläinen, Jarvinen M, Polarek J. Recombinant collagen and gelatin for drug delivery. Adv Drug Delivery Rev. 2003;55:1547–67. doi: 10.1016/j.addr.2003.08.008. [DOI] [PubMed] [Google Scholar]
  • 147.Myllyharju J, Nokelainen M, Vuorela A, Kivirikko KI. Expression of recombinant human type I-III collagens in the yeast pichia pastoris. Biochem Soc Trans. 2000;28:353–57. [PubMed] [Google Scholar]
  • 148.Pulkkinen HJ, Tiitu V, Valonen P, Jurvelin JS, Lammi MJ, Kiviranta I. Engineering of cartilage in recombinant human type II collagen gel in nude mouse model in vivo. Osteoarthritis Cartilage. 2010;18:1077–87. doi: 10.1016/j.joca.2010.05.004. [DOI] [PubMed] [Google Scholar]
  • 149.Dame MK, Varani J. Recombinant collagen for animal product-free dextran microcarriers. In Vitro Cell Dev Biol Anim. 2008;44:407–14. doi: 10.1007/s11626-008-9139-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Liu W, Merrett K, Griffith M, Fagerholm P, Dravida S, Heyne B, Scaiano JC, Watsky MA, Shinozaki N, Lagali N, Munger R, Li F. Recombinant human collagen for tissue engineered corneal substitutes. Biomaterials. 2008;29:1147–58. doi: 10.1016/j.biomaterials.2007.11.011. [DOI] [PubMed] [Google Scholar]
  • 151.Wang Y, Cui FZ, Hu K, Zhu XD, Fan DD. Bone regeneration by using scaffold based on mineralized recombinant collagen. J Biomed Mater Res B. 2008;86:29–35. doi: 10.1002/jbm.b.30984. [DOI] [PubMed] [Google Scholar]
  • 152.Ito H, Steplewski A, Alabyeva T, Fertala A. Testing the utility of rationally engineered recombinant collagen-like proteins for applications in tissue engineering. J Biomed Mater Res A. 2006;76A:551–60. doi: 10.1002/jbm.a.30551. [DOI] [PubMed] [Google Scholar]
  • 153.Vendrely C, Scheibel T. Biotechnological production of spider-silk proteins enables new applications. Macromol Biosci. 2007;7:401–09. doi: 10.1002/mabi.200600255. [DOI] [PubMed] [Google Scholar]
  • 154.Scheibel T. Spider silks: recombinant synthesis, assembly, spinning, and engineering of synthetic proteins. Microb Cell Fact. 2004;3:1–10. doi: 10.1186/1475-2859-3-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Prince JT, McGrath KP, DiGirolamo CM, Kaplan DL. Construction, cloning, and expression of synthetic genes encoding spider dragline silk. Biochemistry. 1995;34:10879–85. doi: 10.1021/bi00034a022. [DOI] [PubMed] [Google Scholar]
  • 156.Zhou Y, Wu S, Conticello VP. Genetically directed synthesis and spectroscopic analysis of a protein polymer derived from a flagelliform silk sequence. Biomacromolecules. 2001;2:111–25. doi: 10.1021/bm005598h. [DOI] [PubMed] [Google Scholar]
  • 157.Huemmerich D, Helsen CW, Quedzuweit S, Oschmann J, Rudolph R, Scheibel T. Primary structure elements of spider dragline silks and their contribution to protein solubility. Biochemistry. 2004;43:13604–12. doi: 10.1021/bi048983q. [DOI] [PubMed] [Google Scholar]
  • 158.Fahnestock SR, Bedzyk LA. Production of synthetic spider dragline silk protein in Pichia pastoris. Appl Microbiol Biotechnol. 1997;47:33–39. doi: 10.1007/s002530050884. [DOI] [PubMed] [Google Scholar]
  • 159.McPherson DT, Xu J, Urry DW. Product purification by reversible phase transition following Escherichia coli expression of genes encoding up to 251 repeats of the elastomeric pentapeptide GVGVP. Protein Expression Purif. 1996;7:51–77. doi: 10.1006/prep.1996.0008. [DOI] [PubMed] [Google Scholar]
  • 160.Urry DW. Thermally riven self-assembly, molecular structuring and entropic mechanisms in elastomeric polypeptides. In: Balaram P, Ramaseshan S, editors. Molecular conformation and biological interactions. Bangalore, India: Indian Acad. of Sci; 1991. pp. 555–83. [Google Scholar]
  • 161.Urry DW. Free energy transduction in polypeptides and proteins based on inverse temperature transitions. Prog Biophys Mol Biol. 1992;57:23–57. doi: 10.1016/0079-6107(92)90003-o. [DOI] [PubMed] [Google Scholar]
  • 162.Hassouneh W, Christensen T, Chilkoti A. Elastin-like polypeptides as a purification tag for recombinant proteins. In: Coligan JE, Dunn BM, Speicher DW, Wingfield PT, editors. Current Protocols in Protein Science. New Jersey: John Wiley & Sons, Inc; 2010. pp. 1–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Meyer DE, Chilkoti A. Purification of recombinant proteins by fusion with thermally-responsive polypeptides. Nature. 1999;17:1112–15. doi: 10.1038/15100. [DOI] [PubMed] [Google Scholar]
  • 164.Wu W-Y, Fong BA, Gilles AG, Wood DW. Recombinant protein purification by self-cleaving elastin-like polypeptide fusion tag. In: Coligan JE, Dunn BM, Speicher DW, Wingfield PT, editors. Current Protocols in Protein Science. New Jersey: John Wiley & Sons, Inc; 2009. pp. 1–18. [DOI] [PubMed] [Google Scholar]
  • 165.Urry DW, Luan CH, Parker TM, Gowda DC, Prasad KU, Reid MC, Safavy A. Temperature of polypeptide inverse temperature transition depends on mean residue hydrophobicity. J Am Chem Soc. 1991;113:4346–48. [Google Scholar]
  • 166.Luan CH, Urry DW. Elastic, plastic, and hydrogel protein-based polymers. In: Mark JE, editor. Polymer Data Handbook. New York: Oxford University Press; 1999. pp. 78–89. [Google Scholar]
  • 167.Wu X, Sallach R, Haller CA, Caves JA, Nagapudi K, Conticello VP, Levenston ME, Chaikof EL. Alterations in physical cross-linking modulate mechanical properties of two-phase protein polymer networks. Biomacromolecules. 2005;6:3037–44. doi: 10.1021/bm0503468. [DOI] [PubMed] [Google Scholar]
  • 168.Barbosa JS, Ribeiro A, Testera AM, Alonso M, Arias FJ, Rodríguez-Cabello JC, Mano JF. Development of biomimetic chitosan-based hydrogels using an elastin-like polymer. Adv Eng Mater. 2010;12:B37–B44. [Google Scholar]
  • 169.Costa RR, Custódio CA, Testera AM, Arias FJ, Rodríguez-Cabello JC, Alves NM, Mano JF. Stimuli-responsive thin coatings using elastin-like polymers for biomedical applications. Adv Funct Mater. 2009;19:3210–18. [Google Scholar]
  • 170.Nicol A, Gowda DC, Parker TM, Urry DW. Cell adhesive properties of bioelastic materials containing cell attachment sequences. In: Gebelein C, Carraher C, editors. Biotechnology and bioactive polymers. New York: Plenum Press; 1994. pp. 95–114. [Google Scholar]
  • 171.Urry DW, Nicol A, Gowda DC, Hoban LD, McKee A, Williams T, Olsen DB, Cox BA. Medical applications of bioelastic materials. In: Gebelein CG, editor. Biotechnological polymers: medical, pharmaceutical and industrial applications. Atlanta, Georgia: Technomic Publishing Co. Inc; 1993. pp. 82–103. [Google Scholar]
  • 172.Numata K, Hamasaki J, Subramanian B, Kaplan DL. Gene delivery mediated by recombinant silk proteins containing cationic and cell binding motifs. J Controlled Release. 2010;146:136–43. doi: 10.1016/j.jconrel.2010.05.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Arias FJ, Reboto V, Martín S, López I, Rodríguez-Cabello JC. Tailored recombinant elastin-like polymers for advanced biomedical and nano(bio)technological applications. Biotechnol Lett. 2006;28:687–95. doi: 10.1007/s10529-006-9045-3. [DOI] [PubMed] [Google Scholar]
  • 174.Benesch J, Mano JF, Reis RL. Proteins and their peptide motifs in acellular apatite mineralization of scaffolds for tissue engineering. Tissue Eng Part B. 2008;14:433–45. doi: 10.1089/ten.teb.2008.0121. [DOI] [PubMed] [Google Scholar]
  • 175.Ganss B, Kim RH, Sodek J. Bone sialoprotein. Crit Rev Oral Biol Med. 1999;10:79–98. doi: 10.1177/10454411990100010401. [DOI] [PubMed] [Google Scholar]
  • 176.Gomes S, Leonor IB, Mano JF, Reis RL, Kaplan DL. Spider silk-bone sialoprotein as a novel fusion protein for bone tissue engineering. Soft Matter. 2010 doi: 10.1039/C1SM05024AArticle. in press. [DOI] [Google Scholar]
  • 177.Huang J, Wong C, George A, Kaplan DL. The effect of genetically engineered spider silk-dentin matrix protein 1 chimeric protein on hydroxyapatite nucleation. Biomaterials. 2007;28:2358–67. doi: 10.1016/j.biomaterials.2006.11.021. [DOI] [PubMed] [Google Scholar]
  • 178.Gomes S, Leonor IB, Mano JF, Reis RL, Kaplan DL. Antimicrobial functionalized genetically engineered spider silk. Biomaterials. 2010 doi: 10.1016/j.biomaterials.2011.02.040. Article in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Jang J-H, Chung C-P. Engineering and expression of a recombinant fusion protein possessing fibroblast growth factor-2 and fibronectin fragment. Biotechnol Lett. 2004;26:1837–40. doi: 10.1007/s10529-004-5278-1. [DOI] [PubMed] [Google Scholar]
  • 180.Kawase Y, Ohdate Y, Shimojo T, Taguchi Y, Kimizuka F, Kato I. Construction and characterization of a fusion protein with epidermal growth factor and the cell-binding domain of fibronectin. FEBS Lett. 1992;298:126–28. doi: 10.1016/0014-5793(92)80037-h. [DOI] [PubMed] [Google Scholar]
  • 181.Hayashi M, Tomita M, Yoshizato K. Production of EGF–collagen chimeric protein which shows the mitogenic activity. Biochim Biophys Acta. 2001;1528:187–95. doi: 10.1016/s0304-4165(01)00187-8. [DOI] [PubMed] [Google Scholar]
  • 182.Nettles DL, Chilkoti A, Setton LA. Applications of elastin-like polypeptides in tissue engineering. Adv Drug Delivery Rev. 2010;62:1479–85. doi: 10.1016/j.addr.2010.04.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Vrhovski B, Weiss AS. Biochemistry of tropoelastin. Eur J Biochem. 1998;258:1–18. doi: 10.1046/j.1432-1327.1998.2580001.x. [DOI] [PubMed] [Google Scholar]
  • 184.Janmey PA, Winer JP, Weisel JW. Fibrin gels and their clinical and bioengineering applications. J R Soc Interface. 2009;6:1–10. doi: 10.1098/rsif.2008.0327. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Lane EB, McLean WHI. Keratins and skin disorders. J Pathol. 2004;204:355–66. doi: 10.1002/path.1643. [DOI] [PubMed] [Google Scholar]
  • 186.Miyagi Y, Chiu LLY, Cimini M, Weisel RD, Radisic M, Li R-K. Biodegradable collagen patch with covalently immobilized VEGF for myocardial repair. Biomaterials. 2011;32:1280–90. doi: 10.1016/j.biomaterials.2010.10.007. [DOI] [PubMed] [Google Scholar]
  • 187.Shen YH, Shoichet MS, Radisi M. Vascular endothelial growth factor immobilized in collagen scaffold promotes penetration and proliferation of endothelial cells. Acta Biomater. 2008;4:477–89. doi: 10.1016/j.actbio.2007.12.011. [DOI] [PubMed] [Google Scholar]
  • 188.Nillesen STM, Geutjes PJ, Wismans R, Schalkwijk J, Daamen WF, Kuppevelt THv. Increased angiogenesis and blood vessel maturation in acellular collagen–heparin scaffolds containing both FGF2 and VEGF. Biomaterials. 2007;28:1123–31. doi: 10.1016/j.biomaterials.2006.10.029. [DOI] [PubMed] [Google Scholar]
  • 189.Sun B, Chen B, Zhao Y, Sun W, Chen K, Zhang J, Wei Z, Xiao Z, Dai J. Crosslinking heparin to collagen scaffolds for the delivery of human platelet-derived growth factor. J Biomed Mater Res B. 2009;91:366–72. doi: 10.1002/jbm.b.31411. [DOI] [PubMed] [Google Scholar]
  • 190.Teixeira S, Yang L, Dijkstra PJ, Ferraz MP, Monteiro FJ. Heparinized hydroxyapatite/collagen three-dimensional scaffolds for tissue engineering. J Mater Sci Mater Med. 2010;21:2385–92. doi: 10.1007/s10856-010-4097-2. [DOI] [PubMed] [Google Scholar]
  • 191.Wu JM, Xu YY, Li ZH, Yuan XY, Wang PF, Zhang XZ, Liu YQ, Guan J, Guo Y, Li RX, Zhang H. Heparin-functionalized collagen matrices with controlled release of basic fibroblast growth factor. J Mater Sci Mater Med. 2011;22:107–14. doi: 10.1007/s10856-010-4176-4. [DOI] [PubMed] [Google Scholar]
  • 192.Zhao Y, Zhang J, Wang X, Chen B, Xiao Z, Shi C, Wei Z, Hou X, Wang Q, Dai J. The osteogenic effect of bone morphogenetic protein-2 on the collagen scaffold conjugated with antibodies. J Controlled Release. 2010;141:30–37. doi: 10.1016/j.jconrel.2009.06.032. [DOI] [PubMed] [Google Scholar]
  • 193.He Q, Zhao Y, Chen B, Xiao Z, Zhang J, Chen L, Chen W, Deng F, Dai J. Improved cellularization and angiogenesis using collagen scaffolds chemically conjugated with vascular endothelial growth factor. Acta Biomater. 2010;7:1084–1093. doi: 10.1016/j.actbio.2010.10.022. [DOI] [PubMed] [Google Scholar]
  • 194.Takeda Y, Tsujigiwa H, Nagatsuka H, Nagai N, Yoshinobu J, Okano M, Fukushima K, Takeuchi A, Yoshino T, Nishizaki K. Regeneration of rat auditory ossicles using recombinant human BMP-2/collagen composites. J Biomed Mater Res A. 2005;73A:133–41. doi: 10.1002/jbm.a.30257. [DOI] [PubMed] [Google Scholar]
  • 195.Côté M-F, Laroche G, Gagnon E, Chevallier P, Doillon CJ. Denatured collagen as support for a FGF-2 delivery system: physicochemical characterizations and in vitro release kinetics and bioactivity. Biomaterials. 2004;25:3761–72. doi: 10.1016/j.biomaterials.2003.10.026. [DOI] [PubMed] [Google Scholar]
  • 196.Maehara H, Sotome S, Yoshii T, Torigoe I, Kawasaki Y, Sugata Y, Yuasa M, Hirano M, Mochizuki N, Kikuchi M, Shinomiya K, Okawa A. Repair of large osteochondral defects in rabbits using porous hydroxyapatite/collagen (HAp/Col) and fibroblast growth factor-2 (FGF-2) J Orthop Res. 2010;28:677–86. doi: 10.1002/jor.21032. [DOI] [PubMed] [Google Scholar]
  • 197.Gavénis K, Klee D, Pereira-Paz RM, Walter Mv, Mollenhauer J, Schneider U, Schmidt-Rohlfing B. BMP-7 loaded microspheres as a new delivery system for the cultivation of human chondrocytes in a collagen type-I gel. J Biomed Mater Res B. 2007;82:275–83. doi: 10.1002/jbm.b.30731. [DOI] [PubMed] [Google Scholar]
  • 198.Borselli C, Ungaro F, Oliviero O, d’Angelo I, Quaglia F, Rotonda MIL, Netti PA. Bioactivation of collagen matrices through sustained VEGF release from PLGA microspheres. J Biomed Mater Res A. 2010;92:94–102. doi: 10.1002/jbm.a.32332. [DOI] [PubMed] [Google Scholar]
  • 199.Chou C-H, Cheng WTK, Lin C-C, Chang C-H, Tsai C-C, Lin F-H. TGF-beta1 immobilized tri-co-polymer for articular cartilage tissue engineering. J Biomed Mater Res B. 2006;77:338–48. doi: 10.1002/jbm.b.30432. [DOI] [PubMed] [Google Scholar]
  • 200.Srouji S, Rachmiel A, Blumenfeld I, Livne E. Mandibular defect repair by TGF-beta and IGF-1 released from a biodegradable osteoconductive hydrogel. J Craniomaxillofac Surg. 2005;33:79–84. doi: 10.1016/j.jcms.2004.09.003. [DOI] [PubMed] [Google Scholar]
  • 201.Kuo Y-C, Ku I-N. Effects of gel concentration, human fibronectin, and cation supplement on the tissue-engineered cartilage. Biotechnol Prog. 2007;23:238–45. doi: 10.1021/bp060253h. [DOI] [PubMed] [Google Scholar]
  • 202.Fan H, Hu Y, Li X, Wu H, Lv R, Bai J, Wang J, Qin L. Ectopic cartilage formation induced by mesenchymal stem cells on porous gelatin-chondroitin-hyaluronate scaffold containing microspheres loaded with TGF-beta1. Int J Artif Organs. 2006;29:602–11. doi: 10.1177/039139880602900610. [DOI] [PubMed] [Google Scholar]
  • 203.Ogawa T, Akazawa T, Tabata Y. In vitro proliferation and chondrogenic differentiation of rat bone marrow stem cells cultured with gelatin hydrogel microspheres for TGF-beta1 release. J Biomater Sci Polym Ed. 2010;21:609–21. doi: 10.1163/156856209X434638. [DOI] [PubMed] [Google Scholar]
  • 204.Kempen DHR, Lu L, Heijink A, Hefferan TE, Creemers LB, Maran A, Yaszemski MJ, Dhert WJA. Effect of local sequential VEGF and BMP-2 delivery on ectopic and orthotopic bone regeneration. Biomaterials. 2009;30:2816–25. doi: 10.1016/j.biomaterials.2009.01.031. [DOI] [PubMed] [Google Scholar]
  • 205.Kempen DHR, Lu L, Hefferan TE, Creemers LB, Maran A, Classic KL, Dhert WJA, Yaszemski MJ. Retention of in vitro and in vivo BMP-2 bioactivities in sustained delivery vehicles for bone tissue engineering. Biomaterials. 2008;29:3245–52. doi: 10.1016/j.biomaterials.2008.04.031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Royce SM, Askari M, Marra KG. Incorporation of polymer microspheres within fibrin scaffolds for the controlled delivery of FGF-1. J Biomater Sci Polym Ed. 2004;15:1327–36. doi: 10.1163/1568562041960016. [DOI] [PubMed] [Google Scholar]
  • 207.Chung Y-I, Ahn K-M, Jeon S-H, Lee S-Y, Lee J-H, Tae G. Enhanced bone regeneration with BMP-2 loaded functional nanoparticle-hydrogel complex. J Controlled Release. 2007;121:91–99. doi: 10.1016/j.jconrel.2007.05.029. [DOI] [PubMed] [Google Scholar]
  • 208.Park K-H, Kim H, Moon S, Na K. Bone morphogenic protein-2 (BMP-2) loaded nanoparticles mixed with human mesenchymal stem cell in fibrin hydrogel for bone tissue engineering. J Biosci Bioeng. 2009;108:530–37. doi: 10.1016/j.jbiosc.2009.05.021. [DOI] [PubMed] [Google Scholar]
  • 209.Campbell PG, Miller ED, Fisher GW, Walker LM, Weiss LE. Engineered spatial patterns of FGF-2 immobilized on fibrin direct cell organization. Biomaterials. 2005;26:6762–70. doi: 10.1016/j.biomaterials.2005.04.032. [DOI] [PubMed] [Google Scholar]
  • 210.Lee J, Choi WI, Tae G, Kim YH, Kang SS, Kim SE, Kim S-H, Jung Y, Kim SH. Enhanced regeneration of the ligament-bone interface using a poly(L-lactide-co-ε-caprolactone) scaffold with local delivery of cells/BMP-2 using a heparin-based hydrogel. Acta Biomater. 2011;7:244–57. doi: 10.1016/j.actbio.2010.08.017. [DOI] [PubMed] [Google Scholar]
  • 211.Kim M, Lee JY, Jones CN, Revzin A, Tae G. Heparin-based hydrogel as a matrix for encapsulation and cultivation of primary hepatocytes. Biomaterials. 2010;31:3596–603. doi: 10.1016/j.biomaterials.2010.01.068. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Gotoh Y, Niimi S, Hayakawa T, Miyashita T. Preparation of lactose–silk fibroin conjugates and their application as a scaffold for hepatocyte attachment. J Mater Chem. 2004;25:1131–40. doi: 10.1016/s0142-9612(03)00633-1. [DOI] [PubMed] [Google Scholar]
  • 213.Fan H, Liu H, Wang Y, Toh SL, Goh JC. Development of a silk cable-reinforced gelatin/silk fibroin hybrid scaffold for ligament tissue engineering. Cell Transplant. 2008;17:1389–401. doi: 10.3727/096368908787648047. [DOI] [PubMed] [Google Scholar]
  • 214.Lu Qiang, Zhang Xiaohui, Hu Xiao, Kaplan DL. Green process to prepare silk fibroin/gelatin biomaterial scaffolds. Macromol Biosci. 2010;10:289–98. doi: 10.1002/mabi.200900258. [DOI] [PubMed] [Google Scholar]
  • 215.Seo Y-K, Yoon H-H, Song K-Y, Kwon S-Y, Lee H-S, Park Y-S, Park J-K. Increase in cell migration and angiogenesis in a composite silk scaffold for tissue-engineered ligaments. J Orthop Res. 2009;27:495–503. doi: 10.1002/jor.20752. [DOI] [PubMed] [Google Scholar]
  • 216.Chen X, Qia Y-Y, Wang L-L, Yin Z, Yin G-L, Zou X-H, Ouyang H-W. Ligament regeneration using a knitted silk scaffold combined with collagen matrix. Biomaterials. 2008;29:3683–92. doi: 10.1016/j.biomaterials.2008.05.017. [DOI] [PubMed] [Google Scholar]
  • 217.Takezawa T, Ozaki K, Takabayashi C. Reconstruction of a hard connective tissue utilizing a pressed silk sheet and type-I collagen as the scaffold for fibroblasts. Tissue Eng. 2007;13:1357–66. doi: 10.1089/ten.2006.0248. [DOI] [PubMed] [Google Scholar]
  • 218.Kirker-Head C, Karageorgiou V, Hofmann S, Fajardo R, Betz O, Merkle HP, Hilbe M, Rechenberg Bv, McCool J, Abrahamsen L, Nazarian A, Cory E, Curtis M, Kaplan DL, Meinel L. BMP-silk composite matrices heal critically sized femoral defects. Bone. 2007;41:247–455. doi: 10.1016/j.bone.2007.04.186. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219.Wenk E, Meinel AJ, Wildy S, Merkle HP, Meinel L. Microporous silk fibroin scaffolds embedding PLGA microparticles for controlled growth factor delivery in tissue engineering. Biomaterials. 2009;30:2571–81. doi: 10.1016/j.biomaterials.2008.12.073. [DOI] [PubMed] [Google Scholar]
  • 220.Gil ES, Frankowski DJ, Bowman MK, Gozen AO, Hudson SM, Spontak RJ. Mixed protein blends composed of gelatin and Bombyx mori silk fibroin: effects of solvent-induced crystallization and composition. Biomacromolecules. 2006;7:728–35. doi: 10.1021/bm050622i. [DOI] [PubMed] [Google Scholar]
  • 221.Shen W, Chen X, Chen J, Yin Z, Heng BC, Chen W, Ouyang H-W. The effect of incorporation of exogenous stromal cell-derived factor-1 alpha within a knitted silk-collagen sponge scaffold on tendon regeneration. Biomaterials. 2010;31:7239–49. doi: 10.1016/j.biomaterials.2010.05.040. [DOI] [PubMed] [Google Scholar]
  • 222.Zhang C, Baez J, Pappu KM, Glatz CE. Purification and characterization of a transgenic corn grain-derived recombinant collagen type I alpha 1. Biotechnol Prog. 2009;25:1660–68. doi: 10.1002/btpr.257. [DOI] [PubMed] [Google Scholar]
  • 223.Nokelainen M, Tu H, Vuorela A, Notbohm H, Kivirikko KI, Myllyharju J. High-level production of human type I collagen in the yeast Pichia pastoris. Yeast. 2001;18:797–806. doi: 10.1002/yea.730. [DOI] [PubMed] [Google Scholar]
  • 224.Olsen DR, Leigh SD, Chang R, McMullin H, Ong W, Tai E, Chisholm G, Birk DE, Berg RA, Hitzeman RA, Toma PD. Production of human type I collagen in yeast reveals unexpected new insights into the molecular assembly of collagen trimers. J Biol Chem. 2001;276:24038–43. doi: 10.1074/jbc.M101613200. [DOI] [PubMed] [Google Scholar]
  • 225.Toman PD, Chisholm G, McMullin H, Giere LM, Olsen DR, Kovach RJ, Leigh SD, Fong BE, Chang R, Daniels GA, Berg RA, Hitzeman RA. Production of recombinant human type I procollagen trimers using a four-gene expression system in the yeast Saccharomyces cerevisiae. J Biol Chem. 2000;275:23303–09. doi: 10.1074/jbc.M002284200. [DOI] [PubMed] [Google Scholar]
  • 226.Fertala A, Sieron AL, Ganguly A, Li SW, Ala-Kokko L, Anumula KR, Prockop DJ. Synthesis of recombinant human procollagen II in a stably transfected tumour cell line (HT1080) Biochem J. 1994;298:31–37. doi: 10.1042/bj2980031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Geddis AE, Prockop DJ. Expression of human COL1A1 gene in stably transfected HT1080 cells: the production of a thermostable homotrimer of type I collagen in a recombinant system. Matrix. 1993;13:399–405. doi: 10.1016/s0934-8832(11)80045-4. [DOI] [PubMed] [Google Scholar]
  • 228.Myllyharju J, Lamberg A, Notbohm H, Fietzek PP, Pihlajaniemi T, Kivirikko KI. Expression of wild-type and modified proα chains of human type I procollagen in insect cells leads to the formation of stable [α1(I)]2α2(I) collagen heterotrimers and [α1(I)]3 homotrimers but not [α2(I)]3 homotrimers. J Biol Chem. 1997;272:21824–30. doi: 10.1074/jbc.272.35.21824. [DOI] [PubMed] [Google Scholar]
  • 229.Tomita M, Kitajima T, Yoshizato K. Formation of recombinant human procollagen I heterotrimers in a baculovirus expression system. J Biochem. 1997;121:1061–69. doi: 10.1093/oxfordjournals.jbchem.a021695. [DOI] [PubMed] [Google Scholar]
  • 230.John DCA, Watson R, Kind AJ, Scott AR, Kadler KE, Bulleid NJ. Expression of an engineered form of recombinant procollagen in mouse milk. Nat Biotechnol. 1999;17:385–89. doi: 10.1038/7945. [DOI] [PubMed] [Google Scholar]
  • 231.Toman PD, Pieper F, Sakai N, Karatzas C, Platenburg E, Wit Id, Samuel C, Dekker A, Daniels GA, Berg RA, Platenburg GJ. Production of recombinant human type I procollagen homotrimer in the mammary gland of transgenic mice. Transgenic Res. 1999;8:415–27. doi: 10.1023/a:1008959924856. [DOI] [PubMed] [Google Scholar]
  • 232.Buechter DD, Paolella DN, Leslie BS, Brown MS, Mehos KA, Gruskin EA. Co-translational Incorporation of trans-4-hydroxyproline into recombinant proteins in bacteria. J Biol Chem. 2003;278:645–50. doi: 10.1074/jbc.M209364200. [DOI] [PubMed] [Google Scholar]
  • 233.Nokelainen M, Helaakoski T, Myllyharju J, Notbohm H, Pihlajaniemi T, Fietzek PP, Kivirikko KI. Expression and characterization of recombinant human type II collagens with low and high contents of hydroxylysine and its glycosylated forms. Matrix Biol. 1998;16:329–38. doi: 10.1016/s0945-053x(98)90004-x. [DOI] [PubMed] [Google Scholar]
  • 234.Vuorela A, Myllyharju J, Nissi R, Pihlajaniemi T, Kivirikko KI. Assembly of human prolyl 4-hydroxylase and type III collagen in the yeast Pichia pastoris: formation of a stable enzyme tetramer requires coexpression with collagen and assembly of a stable collagen requires coexpression with prolyl 4-hydroxylase. EMBO J. 1997;16:6702–12. doi: 10.1093/emboj/16.22.6702. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235.Vaughan PR, Galanis M, Richards KM, Tebb TA, Ramshaw JAM, Werkmaister JA. Production of recombinant hydroxylated human type III collagen fragment in Saccharomyces cerevisiae. DNA Cell Biol. 1998;17:511–18. doi: 10.1089/dna.1998.17.511. [DOI] [PubMed] [Google Scholar]
  • 236.Lamberg A, Helaakoski T, Myllyharju J, Peltonen S, Notbohm H, Pihlajaniemi T, Kivirikko KI. Characterization of human type III collagen expressed in a baculovirus system. Production of a protein with a stable triple helix requires coexpression with the two types of recombinant prolyl 4-hydroxylase subunit. J Biol Chem. 1996;271:11988–95. doi: 10.1074/jbc.271.20.11988. [DOI] [PubMed] [Google Scholar]
  • 237.Tomita M, Ohkura N, Ito M, Kato T, Royce PM, Kitajima T. Biosynthesis of recombinant human pro-alpha 1(III) chains in a baculovirus expression system: production of disulphide-bonded and non-disulphide-bonded species containing full-length triple helices. Biochem J. 1995;312:847–53. doi: 10.1042/bj3120847. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 238.Tomita M, Munetsuna H, Sato T, Adachi T, Hino R, Hayashi M, Shimizu K, Nakamura N, Tamura T, Yoshizato K. Transgenic silkworms produce recombinant human type III procollagen in cocoons. Nat Biotechnol. 2003;21:52–56. doi: 10.1038/nbt771. [DOI] [PubMed] [Google Scholar]
  • 239.Fichard A, Tillet E, Delacoux F, Garrone R, Ruggiero F. Human recombinant α1(V) collagen chain. J Biol Chem. 1997;272:30083–87. doi: 10.1074/jbc.272.48.30083. [DOI] [PubMed] [Google Scholar]
  • 240.Tillet E, Wiedemann H, Golbik R, Pan T-C, Zhang R-Z, Mann K, Chu M-L, Timpl R. Recombinant expression and structural and binding properties of α1 (VI) and α2(VI) chains of human collagen type VI. Eur J Biochem. 1994;221:177–87. doi: 10.1111/j.1432-1033.1994.tb18727.x. [DOI] [PubMed] [Google Scholar]
  • 241.Chen M, Costa FK, Lindvay CR, Han Y-P, Woodley DT. The Recombinant Expression of Full-length Type VII Collagen and Characterization of Molecular Mechanisms Underlying Dystrophic Epidermolysis Bullosa. J Biol Chem. 2002;277:2118–24. doi: 10.1074/jbc.M108779200. [DOI] [PubMed] [Google Scholar]
  • 242.Frischholz S, Beier F, Girkontaite I, Wagner K, Pöschl E, Turnay J, Mayer U, Mark Kvd. Characterization of human type X procollagen and its NC-1 domain expressed as recombinant proteins in HEK293 cells. J Biol Chem. 1998;273:4547–55. doi: 10.1074/jbc.273.8.4547. [DOI] [PubMed] [Google Scholar]
  • 243.Warner LR, Blasick CM, Brown RJ, Oxford JT. Expression, purification, and refolding of recombinant collagen alpha1(XI) amino terminal domain splice variants. Protein Expression Purif. 2007;52:403–09. doi: 10.1016/j.pep.2006.10.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Steplewski A, Majsterek I, McAdams E, Rucker E, Brittingham RJ, Ito H, Hirai K, Adachi E, Jimenez SA, Fertala A. Thermostability gradient in the collagen triple helix reveals its multi-domain structure. J Mol Biol. 2004;338:989–98. doi: 10.1016/j.jmb.2004.03.037. [DOI] [PubMed] [Google Scholar]
  • 245.Werten MWT, Teles H, Moers APHA, Wolbert EJH, Sprakel J, Eggink G, Wolf FA. Precision gels from collagen-inspired triblock copolymers. Biomacromolecules. 2009;10:1106–13. doi: 10.1021/bm801299u. [DOI] [PubMed] [Google Scholar]
  • 246.Sallach RE, Conticello VP, Chaikof EL. Expression of a recombinant elastin-like protein in Pichia pastoris. Biotechnol Prog. 2009;25:1810–18. doi: 10.1002/btpr.208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 247.Jordan SW, Haller CA, Sallach RE, Apkarian RP, Hanson SR, Chaikof EL. The effect of a recombinant elastin-mimetic coating of an ePTFE prosthesis on acute thrombogenicity in a baboon arteriovenous shunt. Biomaterials. 2007;28:1191–97. doi: 10.1016/j.biomaterials.2006.09.048. [DOI] [PubMed] [Google Scholar]
  • 248.Nagapudi K, Brinkman WT, Leisen J, Thomas BS, Wright ER, Haller C, Wu X, Apkarian RP, Conticello VP, Chaikof EL. Protein-based thermoplastic elastomers. Macromolecules. 2005;38:345–54. [Google Scholar]
  • 249.Panitch A, Yamaoka T, Fournier MJ, Mason TL, Tirrell DA. Design and biosynthesis of elastin-like artificial extracellular matrix proteins containing periodically spacedfibronectin CS5 domains. Macromolecules. 1999;32:1701–03. [Google Scholar]
  • 250.Sallacha RE, Cui W, Balderrama F, Martinez AW, Wen J, Haller CA, Taylor JV, Wright ERRCL, Jr, Chaikof EL. Long-term biostability of self-assembling protein polymers in the absence of covalent crosslinking. Biomaterials. 2010;31:779–91. doi: 10.1016/j.biomaterials.2009.09.082. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 251.Woodhouse KA, Klement P, Chen V, Gorbet MB, Keeley FW, Stahl R, Fromstein JD, Bellingham CM. Investigation of recombinant human elastin polypeptides as non-thrombogenic coatings. Biomaterials. 2004;25:4543–53. doi: 10.1016/j.biomaterials.2003.11.043. [DOI] [PubMed] [Google Scholar]
  • 252.Wright ER, McMillan RA, Cooper A, Apkarian RP, Conticello VP. Thermoplastic elastomer hydrogels via self-assembly of an elastin-mimetic triblock polypeptide. Adv Funct Mater. 2002;12:149–54. [Google Scholar]
  • 253.Yamaoka T, Tamura T, Seto Y, Tada T, Kunugi S, Tirrell DA. Mechanism for the phase transition of a genetically engineered elastin model peptide (VPGIG)40 in aqueous solution. Biomacromolecules. 2003;4:1680–85. doi: 10.1021/bm034120l. [DOI] [PubMed] [Google Scholar]
  • 254.Nagapudi K, Brinkman WT, Leisen JE, Huang L, McMillan RA, Apkarian RP, Conticello VP, Chaikof EL. Photomediated solid-state cross-linking of an elastin mimetic recombinant protein polymer. Macromolecules. 2002;35:1730–37. [Google Scholar]
  • 255.Rabotyagova O, Cebe P, Kaplan DL. Self-assembly of genetically engineered spider silk block copolymers. Biomacromolecules. 2009;10:229–36. doi: 10.1021/bm800930x. [DOI] [PubMed] [Google Scholar]
  • 256.Rabotyagova OS, Cebe P, Kaplan DL. Role of polyalanine domains in β-sheet formation in spider silk block copolymers. Macromol Biosci. 2010;10:49–59. doi: 10.1002/mabi.200900203. [DOI] [PubMed] [Google Scholar]
  • 257.Arcidiacono S, Mello C, Kaplan DL, Cheley S, Bayley H. Purification and characterization of recombinant spider silk expressed in Escherichia coli. Appl Microbiol Biotechnol. 1998;49:31–38. doi: 10.1007/s002530051133. [DOI] [PubMed] [Google Scholar]
  • 258.Fahnestock SR, Irwin SL. Synthetic spider dragline silk proteins and their production in Escherichia coli. Appl Microbiol Biotechnol. 1997;47:23–32. doi: 10.1007/s002530050883. [DOI] [PubMed] [Google Scholar]
  • 259.Lewis RV, Hinman M, Kothakota S, Fournier MJ. Expression and purification of a spider silk protein: a new strategy for producing repetitive proteins. Protein Expr Purif. 1996;7:400–06. doi: 10.1006/prep.1996.0060. [DOI] [PubMed] [Google Scholar]
  • 260.Fukushima Y. Genetically engineered syntheses of tandem repetitive polypeptides consisting of glycine-rich sequence of spider dragline silk. Biopolymers. 1998;45:269–79. doi: 10.1002/(SICI)1097-0282(19980405)45:4<269::AID-BIP1>3.0.CO;2-J. [DOI] [PubMed] [Google Scholar]
  • 261.Mello CM, Soares JW, Arcidiacono S, Butle MM. Acid extraction and purification of recombinant spider silk proteins. Biomacromolecules. 2004;5:1849–52. doi: 10.1021/bm049815g. [DOI] [PubMed] [Google Scholar]
  • 262.Butler SP, O’Sickey TK, Lord ST, Lubon H, Gwazdauskas FC, Velande WH. Secretion of recombinant human fibrinogen by the murine mammary gland. Transgenic Res. 2004;13:437–50. doi: 10.1007/s11248-004-9589-8. [DOI] [PubMed] [Google Scholar]
  • 263.Tojo N, Miyagia I, Miuraa M, Ohi H. Recombinant human fibrinogen expressed in the yeast Pichia pastoris was assembled and biologically active. Protein Expression Purif. 2008;59:289–96. doi: 10.1016/j.pep.2008.02.010. [DOI] [PubMed] [Google Scholar]
  • 264.Cutler SM, Garca AJ. Engineering cell adhesive surfaces that direct integrin α5β1 binding using a recombinant fragment of fibronectin. Biomaterials. 2003;24:1759–70. doi: 10.1016/s0142-9612(02)00570-7. [DOI] [PubMed] [Google Scholar]
  • 265.Rico P, González-García C, Petrie TA, García AJ, Salmerón-Sánchez M. Molecular assembly and biological activity of a recombinant fragment of fibronectin (FNIII(7-10)) on poly(ethyl acrylate) Colloids Surf B. 2010;78:310–16. doi: 10.1016/j.colsurfb.2010.03.019. [DOI] [PubMed] [Google Scholar]
  • 266.Brewster LP, Washington C, Brey EM, Gassman A, Subramanian A, Calceterra J, Wolf W, Hall CL, Velander WH, Burgess WH, Greisle HP. Construction and characterization of a thrombin-resistant designer FGF-based collagen binding domain angiogen. Biomaterials. 2008;29:327–36. doi: 10.1016/j.biomaterials.2007.09.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 267.Pang Y, Wang X, Ucuzian AA, Brey EM, Burgess WH, Jones KJ, Alexander TD, Greisler HP. Local delivery of a collagen-binding FGF-1 chimera to smooth muscle cells in collagen scaffolds for vascular tissue engineering. Biomaterials. 2010;31:878–85. doi: 10.1016/j.biomaterials.2009.10.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 268.Yan X, Chen B, Lin Y, Li Y, Xiao Z, Hou X, Tan Q, Dai J. Acceleration of diabetic wound healing by collagen-binding vascular endothelial growth factor in diabetic rat model. Diabetes Res Clin Pract. 2010;90:66–72. doi: 10.1016/j.diabres.2010.07.001. [DOI] [PubMed] [Google Scholar]
  • 269.Hashi H, Hatai M, Kimizuka F, Kato I, Yaoi Y. Angiogenic activity of a fusion protein of the cell-binding domain of fibronectin and basic fibroblast growth factor. Cell Struct Funct. 1994;19:37–47. doi: 10.1247/csf.19.37. [DOI] [PubMed] [Google Scholar]
  • 270.Nishi N, Matsushita O, Yuube K, Miyanaka H, Okabe A, Wada F. Collagen-binding growth factors: production and characterization of functional fusion proteins having a collagen-binding domain. Proc Natl Acad Sci USA. 1998;95:7018–23. doi: 10.1073/pnas.95.12.7018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 271.Sheng Z, Chang S-B, Chirico WJ. Expression and purification of a biologically active basic fibroblast growth factor fusion protein. Protein Expression Purif. 2003;27:267–71. doi: 10.1016/s1046-5928(02)00601-0. [DOI] [PubMed] [Google Scholar]
  • 272.Ogiwara K, Nagaoka M, Cho C-S, Akaike T. Construction of a novel extracellular matrix using a new genetically engineered epidermal growth factor fused to IgG-Fc. Biotechnol Lett. 2005;27:1633–37. doi: 10.1007/s10529-005-2605-0. [DOI] [PubMed] [Google Scholar]
  • 273.Qiu W, Huang Y, Teng W, Cohn CM, Cappello J, Wu X. Complete recombinant silk-elastinlike protein-based tissue scaffold. Biomacromolecules. 2010;11:3219–27. doi: 10.1021/bm100469w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 274.Yang M, Tanaka C, Yamauchi K, Ohgo K, Kurokawa M, Asakura T. Silk-like materials constructed from sequences of Bombyx mori silk fibroin, fibronectin, and elastin. J Biomed Mater Res A. 2008;84:353–63. doi: 10.1002/jbm.a.31348. [DOI] [PubMed] [Google Scholar]
  • 275.Kambe Y, Yamamoto K, Kojima K, Tamada Y, Tomita N. Effects of RGDS sequence genetically interfused in the silk fibroin light chain protein on chondrocyte adhesion and cartilage synthesis. Biomaterials. 2010;31:7503–11. doi: 10.1016/j.biomaterials.2010.06.045. [DOI] [PubMed] [Google Scholar]
  • 276.Du C, Wang M, Liu J, Pan M, Cai Y, Yao J. Improvement of thermostability of recombinant collagen-like protein by incorporating a foldon sequence. Appl Microbiol Biotechnol. 2008;79:195–202. doi: 10.1007/s00253-008-1427-0. [DOI] [PubMed] [Google Scholar]
  • 277.Chen B, Lin H, Wang J, Zhao Y, Wang B, Zhao W, Sun W, Dai J. Homogeneous osteogenesis and bone regeneration by demineralized bone matrix loading with collagen-targeting bone morphogenetic protein-2. Biomaterials. 2007;28:1027–35. doi: 10.1016/j.biomaterials.2006.10.013. [DOI] [PubMed] [Google Scholar]
  • 278.Chen B, Lin H, Zhao Y, Wang B, Zhao Y, Liu Y, Liu Z, Dai J. Activation of demineralized bone matrix by genetically engineered human bone morphogenetic protein-2 with a collagen binding domain derived from von Willebrand factor propolypeptide. J Biomed Mater Res A. 2007;80:428–34. doi: 10.1002/jbm.a.30900. [DOI] [PubMed] [Google Scholar]
  • 279.Zhao Y, Chen B, Lin H, Sun W, Zhao W, Zhang J, Dai J. The bone-derived collagen containing mineralized matrix for the loading of collagen-binding bone morphogenetic protein-2. J Biomed Mater Res A. 2009;88:725–34. doi: 10.1002/jbm.a.31928. [DOI] [PubMed] [Google Scholar]
  • 280.Tuan TL, Cheung DT, Wu LT, Yee A, Gabriel S, Han B, Morton L, Nimni ME, Hall FL. Engineering, expression and renaturation of targeted TGF-beta fusion proteins. Connect Tissue Res. 1996;34:1–9. doi: 10.3109/03008209609028888. [DOI] [PubMed] [Google Scholar]
  • 281.Andrades JA, Wu LT, Hall FL, Nimni ME, Becerra J. Engineering, expression, and renaturation of a collagen-targeted human bFGF fusion protein. Growth Factors. 2001;18:261–75. doi: 10.3109/08977190109029115. [DOI] [PubMed] [Google Scholar]
  • 282.Lin H, Chen B, Sun W, Zhao W, Zhao Y, Dai J. The effect of collagen-targeting platelet-derived growth factor on cellularization and vascularization of collagen scaffolds. Biomaterials. 2006;27:5708–14. doi: 10.1016/j.biomaterials.2006.07.023. [DOI] [PubMed] [Google Scholar]
  • 283.Zhang Y, Xiang Q, Dong S, Li C, Zhou Y. Fabrication and characterization of a recombinant fibronectin/cadherin bio-inspired ceramic surface and its influence on adhesion and ossification in vitro. Acta Biomater. 2010;6:776–85. doi: 10.1016/j.actbio.2009.08.025. [DOI] [PubMed] [Google Scholar]
  • 284.Ishikawa T, Terai H, Yamamoto T, Harada K, Kitajima T. Delivery of a growth factor fusion protein having collagen-binding activity to wound tissues. Artif Organs. 2003;27:147–54. doi: 10.1046/j.1525-1594.2003.07009.x. [DOI] [PubMed] [Google Scholar]
  • 285.Elloumi I, Kobayashi R, Funabashi H, Mie M, Kobatake E. Construction of epidermal growth factor fusion protein with cell adhesive activity. Biomaterials. 2006;27:3451–58. doi: 10.1016/j.biomaterials.2006.02.003. [DOI] [PubMed] [Google Scholar]
  • 286.Sun W, Lin H, Chen B, Zhao W, Zhao Y, Dai J. Promotion of peripheral nerve growth by collagen scaffolds loaded with collagen-targeting human nerve growth factor-beta. J Biomed Mater Res A. 2007;84:1054–61. doi: 10.1002/jbm.a.31417. [DOI] [PubMed] [Google Scholar]

RESOURCES