Abstract
Effective drug delivery is essential for the success of a medical treatment. Polymeric drug delivery systems (DDSs) are preferred over systemic administration of drugs due to their protection capacity, directed release and reduced side effects. Among the numerous polymer sources, silks and recombinant silks have drawn significant attention over the past decade as DDSs. Native silk is produced from a variety of organisms, which are then used as sources or guides of genetic material for heterologous expression or engineered designs. Recombinant silks bear the outstanding properties of natural silk, such as processability in aqueous solution, self-assembly, drug loading capacity, drug stabilization/protection and degradability, while incorporating specific properties beneficial for their success as DDS, such as monodispersity and tailored physicochemical properties. Moreover, the on-demand inclusion of sequences that customize the DDS for the specific application enhances efficiency. Often, inclusion of a drug into a DDS is achieved by simple mixing or diffusion and stabilized by non-specific molecular interactions; however, these interactions can be improved by the incorporation of drug-binding peptide sequences. In this review we provide an overview of native sources for silks and silk sequences, as well as the design and formulation of recombinant silk biomaterials as drug delivery systems in a variety of formats, such as films, hydrogels, porous sponges, or particles.
Keywords: Biomaterials, Silk-like proteins, Wound healing, Cancer therapy, Tissue regeneration
Graphical Abstract

1. Introduction
Drugs are traditionally delivered via systemic administration for small molecule drugs, most commonly through the enteral route of oral administration, where the drug is primarily absorbed in the small intestine before ultimately dispersing throughout the body. This delivery route is not suitable for every drug, which is subject to the harsh gastrointestinal (GI) environment, a long transit time, and the first-pass effect, whereby the initial concentration of drug is reduced by metabolism in the liver [1]. Intravenous injection is the most common parenteral route of systemic drug delivery, which overcomes some of the challenges associated with oral administration by allowing a high concentration of drug to circumvent physiological barriers with the shortest transit time. These delivery routes can sometimes be ineffective at treating disease, as it is challenging to reach therapeutic concentrations at the target disease site while balancing the use of higher doses that can be harmful to healthy organs and tissues. For larger, more complex drugs such as peptides, proteins or nucleic acid, systemic delivery of the naked macromolecules can lead to premature degradation through native proteases and nucleases and may lead to immune responses. These larger molecules are also usually unable to cross the cell membrane for therapeutic action. To circumvent these limitations, drug delivery systems (DDSs) which can provide more targeted outcomes are necessary. DDSs can be formulated as locally implanted reservoirs where therapeutically-relevant concentrations can reach the disease site while minimizing side-effects [2]. Alternatively, DDSs can encapsulate drugs into nano- or micro-sized carriers that are programmed to target a diseased site and release drugs upon exposure to specific environmental signals or via cellular uptake. The drug release profile from DDSs can be adjusted to the specific need, while minimizing drug-related toxic side effects. In addition, DDSs should also provide protection of the drug to maximize stability and bioactivity [3, 4].
Polymers are suitable candidates for DDSs due to their physicochemical stability and tunable properties. However, synthetic polymers present challenges for DDS that can limit applicability. For example, they can exhibit a broad molecular weight distribution, sometimes can present issues of biocompatibility, are often processed in organic solvents and some degradation products can be inflammatory (e.g., some polyesters). The use of protein-based polymers avoids these limitations as they are usually monodisperse (if generated via bioengineering approaches), biocompatible, biodegradable, degrade enzymatically into amino acids, and can be processed under mild, aqueous conditions. Biopolymers also contain many reactive chemical handles which can be used for chemical conjugation to enhance targeting, stability and overall function. These characteristics make them attractive candidates for DDS. Within this context, silk protein has gained a lot of attention for its low immunogenicity and ability to be formulated into many material formats including hydrogels, films, sponges and particles. Silk-based DDS have demonstrated favorable bioactivities in delivery of chemotherapy agents for cancer treatments, promoting wound healing, sustained release of cytokines and morphogens, delivering antibiotics, and gene therapy, among others.
Silks have a unique combination of properties [5], and can be sourced from natural sources (silkworms) or by expression in heterological systems; mainly bacteria and yeast (although many other systems are available currently) [6]. The development of recombinant DNA technologies provides more consistent control of polymer features, and this approach has supported the combination of silk-like motifs with other therapeutically relevant protein sequences, providing tailor-made protein-based biomaterials for DDSs. In addition, recombinant polymers are homogeneous, with nearly monodisperse molecular weight and composition, and good batch-to-batch consistency. These materials are also biodegradable, often with no toxic side products, and the degradation rate can be tailored on demand, for example, via inclusion of protease-sensitive domains. Another advantage of these recombinant materials is that they allow the inclusion of cell-and/or tissue-targeting domains, increasing the specificity and selectivity of the DDS.
In this review we discuss natural sources of silk, the structure and properties related to primary sequence and how these features have been utilized to guide the design of recombinant analogues. We then provide an overview of how these recombinant silks have been used as local implant DDSs in various formats, such as hydrogels, films, and sponges. We will also discuss silk particle-based DDSs for systemic delivery and how these have been tailored to accommodate drugs ranging from small molecules to large biomacromolecules, as well as inorganic nanoparticles. Lastly, we provide a perspective on the future outlook for silk and recombinant silk DDSs.
2. Natural Silk Sources and their Recombinant Analogues
Silks are protein-based fibers spun from various arthropods including silkworms, spiders, ants, bees, wasps, and lacewing insects [7]. These silks exhibit different mechanical properties depending on their required function. For example, orb weaving spiders produce seven types of silk from different glands for building web frames and draglines (major ampullate), supporting web frames (minor ampullate), catching prey (flagelliform), coating webs with sticky substrates (aggregate), wrapping prey (aciniform), attachment of silk to substrates (pyriform), and constructing egg sacs (tubuliform) [8]. The major ampullate dragline silks exhibit superior toughness, even to Kevlar [9]. The flagelliform silks, while three orders of magnitude lower in stiffness than dragline, also have high toughness due to the high extensibility for catching prey. The tubuliform silk in egg cases exhibit high strength but low elasticity, and unlike other silks can bend before breaking [10]. Aciniform silks are strong and also highly extensible (86%) for wrapping prey, which makes them 50% tougher than dragline [11]. Silkworm silks contain both fibroin, which makes up the core fiber and mechanics, and sericin, which acts as a sticky glue holding the fibers together [12]. The variety of mechanical properties of these natural materials provides an extensive tool box for the creation of various DDS. The mechanical stiffness of particulate DDS can influence circulation half-life, biodistribution, infiltration, and cell uptake [13]. Stiffness of implantable DDS has been shown to play a role in macrophage activation phenotypes [2].
The variety of mechanical properties is a result of the silk composition, protein structure, and primary sequence. The toughness of silkworm silks and major ampullate dragline silk is a result of repeats of alternating hydrophobic and hydrophilic domains, where the hydrophobic domains form antiparallel β-sheet crystals for strength, while the hydrophilic domains form amorphous regions for extensibility [14, 15]. The domestic mulberry silkworm fibrion from Bombyx mori, contains GAGAGS motifs which make up the hydrophobic domains [15], while wild, non-mulberry silkworm fibroins and major ampullate dragline spidroins contain stronger interacting polyalanine motifs [14, 16]. Silks from the order Hymenoptera, including bee, ant, and hornet silks, contain four small, highly alpha-helical, nonrepetitive silk proteins rich in alanine which assemble through coiled-coil motifs [17–20]. Egg stalk silks from lacewings have a unique cross-beta structure, which provides high stiffness but also high extensibility upon wetting or high humidity [21, 22]. The mechanical properties of a silk-based DDS can be controlled via crystalline beta-sheet content, which in turn controls enzymatic degradation in vivo and drug release kinetics [23].
In addition to mechanical properties, silks exhibit excellent biological properties that make them exceptional candidates for drug delivery applications. Spider silks as well as silkworm fibroin and sericin are individually well tolerated in in vivo applications; however, sericin must be removed from fibroin through a degumming process, as its use in conjunction with fibroin activates an adaptive immune response [24]. Fibroin alone is non-toxic, biocompatible and biodegradable [25], and degradation can be tuned based on the material morphology and processing conditions related to crystallinity [26]. Additionally, the use of silk fibroin and sericin individually have been shown to support cell proliferation and migration [27, 28]. Non-mulberry silkworm fibroins contain integrin-binding fibronectin-like RGD motifs which further enhance proliferation and migration as they provide handles for cell attachment [29], where Antheraea mylitta and Antheraea yamamai contain 10 and 12 of these motifs, respectively [16]. Sericin’s hydrophilic composition provides interesting antibacterial [30], antioxidant [31], anti-inflammatory [32], anticancer [33], and UV resistant properties [34].
In engineering materials for drug delivery, recombinant DNA technologies are utilized to modify silks with mutations to alter charge and drug transport kinetics [35, 36], or to generate chemical conjugation handles [37, 38]. The fusion of bioactive peptides can be performed to load certain drugs [39], improve drug retention [40], target particular cell types [41], tune material degradation [42], and enhance cellular infiltration [38, 43–45]. B. mori silkworm silk has been harvested for textiles since the 4th millennium BC [46], thus silkworm farming, or sericulture, can produce large quantities of the material at a relatively low cost [47]. Due to this availability, recombinant engineering of silks have been pursued for more selective goals, such as transgenic silkworms to modify silk fibroin with other components directly [48], or bioactive fusions to native sericin though TALEN-mediated genome editing [49].
Unlike B. mori silkworms, spiders are territorial and cannot be farmed, thus unlike silkworm silk, large quantities of the protein are not readily available. Thus, recombinant DNA strategies have emerged to express major ampullate spidroins in other hosts such as E. coli [43], Pichia pastoris [50], goats [51], plants [52], and B. mori silkworms [53]. Recombinant analogues of dragline spidroin proteins from Nephila clavipes (MaSp1/1F9 and MaSp2/2E12) and Euprosthenops australis (MaSp1/4RepCT) have been generated by cloning sequences derived from the internal repetitive sequences of the respective spidroin (Table 1) [50, 54]. Additionally, shorter length consensus repeats have been derived from N. clavipes (MaSp1 and MaSp2) and A. diadematus (ADF-3 and ADF-4) [43, 55, 56]. The number of repeats that can be expressed in E. coli is limited by protein solubility and stability of the repetitive, glycine rich gene composition [57]. Metabolically engineered E. coli with elevated glycyl-tRNA produced up to 96 repeats, a 284.8 kDa MaSp1, with yields around 0.5–0.75 g/L in high cell density cultivation. This recombinant silk was formed into fibers with comparable properties to native dragline silk fibers, where the toughness was twice that in the native fiber.
Table 1:
Silk sources and the repetitive motifs and their sequence alignments used in recombinant silk production.
| Source | Protein | Recombinant Protein/Motif | Repetitive Silk-like Sequence Alignments | Number of Repeats | Refs |
|---|---|---|---|---|---|
| Bombyx mori | Fibroin | Hydrophobic motif | GAGAGS | 1, 2, 4 2 4 8 |
[68, 69] [70] [71] [42] |
| Sericin | Ser1 Consensus | SSTGSSSNTDSNSNSVGSSTSGGSSTYGYSSNSRDGSV | 2 4, 8, 12 |
[61, 62] [59] |
|
| Sericin | -----------STDLAGSSTSGGSSTYGYSSNSRDGSV SSTGSSSNTDASTDLTGSSTSGGSSTYGYSSSNRDGSV LATGSSSNTDASTTEE STTSAGSSTEGYS-------- |
1 | [60, 63, 64] | ||
| Ser4mer | SSTGSTSNTDSSSKSAGSRTSGGSSTYGYSSSHRGGSV SSTGSSSNTDSSTKNAGSSTSGGSSTYGYSSSHRGGSV SSTGSSSNTDSSTKSAGSSTSGGSSTYGYSSRHRGGRV SSTGSSSTTDASSNSVGSSTSGGSSTYGYSSNSRDGS- |
1 | [65] | ||
| Antheraea pernyi | Fibroin | EAEFN | EEAAAAAAAAAAAAEETGRGDSPAAS | 5, 10 | [67] |
| Nephila clavipes | MaSp1 | 1F9 | AGQGGYGGLGSQG AGRGGLGGQGAGAAAAAAAGGAGQ GGLGGQG A GQGAGASAAAA GGAGQGGYGGLGSQG AGRGGLGGQGAGAVAAAAAGGAGQGGYGGLGSQG AGR GGQGAGA AAAAAGGAGQRGYGGLGNQG |
9 | [50] |
| MaSp1 Consensus | SGRGGLGGQGAGA AAAAGGAGQGGYGGLGSQGT | 6 15 32, 48, 64, 80, 96 |
[44, 72–74] [41, 43] [57] |
||
| A block | GAGA AAAAGGAGTS | BA, BA2, BA3, BA6 | [75] | ||
| B block | SQGGYGGLGSQGSGRGGLGGQTS | ||||
| MaSp2 | 2E12 | --G PGGYGPGQQGP GAAAAASA GRGPGGYGPGQQGPGGSGAAAAAA SGPGGYGPGQQGPGGPGAAAAAAA GRGPGGYGPGQQGPGGSGAAAAAA GRGPGGYGPGQQGPGGPGAAAAAA |
12 | [50] | |
| MaSp2 Consensus | GPGGYGPGQQGPSGPGSAAAAAAAA GPGGYGPGQQTS------- |
9 15 |
[40, 55] [40] |
||
| EMS2 | EGPGGYGPGQQGPSGPGSAAAAAAAA GPGGYGPGQQTS------- |
15 | [35] | ||
| Araneus diadematus | ADF-3 | A block | GPYGPGASAAAAAAGGYGPGSGQQ | (AQ)12, (QAQ)8 (AQ)12, (AQ)24 |
[56] [76] |
| Q block | GPGQQGPGQQGPGQQGPGQQ | ||||
| ADF-4 | C block/eADF4(C16) | GSSAAAAAAAASGPGGYGPENQGPSGPGGYGPGGP | 16 | [37, 38, 56, 76–81] | |
| eADF4(κ16) | GSSAAAAAAAASGPGGYGPKNQGPSGPGGYGPGGP | [36, 37, 80] | |||
| eADF4(Ω16) | GSSAAAAAAAASGPGGYGPQNQGPSGPGGYGPGGP | [80] | |||
| Euprosthenops australis | MaSp1 | 4RepCT | QGGYGGLGQGGYGQGAGSS AAAAAAAAAAAAGGQGGQ GQGGYGQGSGGS AAAAAAAAAAAAAAAGR GQGGYGQGSGGN AAAAAAAAAAAAAAAGQGGQGGYGRQSQG A GS AAAAAAAAAAAAAAGSG QGGYGGQGQGGYGQ SS |
1 | [45, 54, 82–88] |
| Latrodectus mactans | TuSp1 | eTuSp1 | FSSASSASAVGQVGYQIGLNAAQTLGISNAPAFADAVSQ AVRTVGVGASPFQYANAVSNAFGQLLGGQGILTQENAAG LASSVSSAISSAASSVAAQAASAAQSSAFAQSQAAAQAF SQAASRSASQSAAQAGSSSTSTTTTTSQAASQAASQSAS SSXSA |
1 | [89] |
| Mallada signata | MalXB2 | AS module | GSAGASSNGSSATASK GSAGATSNGSTAVASK GSAGASSGNSTASATK |
8 | [90] |
Repetitive, high molecular weight silks are difficult to produce heterologously in microbial hosts, while the expression of full length native silk proteins from Hymenopterans can be achieved as these are small nonrepetitive proteins [17–20]. For example, honey bee silk proteins (Amel1-4) were expressed in E. coli and recovered from inclusion bodies (0.2–2.5g/L) [22]. However, research is limited in applying these materials to drug delivery.
Recombinant B. mori silk-like proteins have also been expressed heterologously. The predominant GAGAGS motif in conjunction with stimuli-responsive elastin-like polypeptides led to the development of recombinant silk-elastin-like polypeptides (SELPs) [58]. Recombinant sericin based on the 38 amino acid consensus repeat in the Ser1 gene (Table 1) of B. mori was expressed in E. coli [59], Pichia pastoris and cell free systems [60]. These recombinant sericins exhibited cell proliferative [61], cryoprotective [62], bactericidal [63], and antioxidant properties [64]. The 4mer repetitive sequence from the native sericin gene showed higher solubility than the synthetic consensus repeats [65]. For non-mulberry fibroin, the increased stability of polyalanine rich domains makes the cocoon more difficult to process using LiBr extraction methods [66]. Inspired by Antheraea pernyi, recombinant silk EAEFNn (n = 5 or 10) was expressed in in E. coli to generate coatings to promote osteogenic adhesion and growth (Table 1) [67].
The specific application of these silks and their recombinant analogs in drug delivery will be discussed in more detail in the following sections.
3. Implantable devices
Silk in bulk format can be used as a depot for the delivery of drugs, biomolecules, growth factors, or DNA at a fast or slow sustained rate [91, 92]. Silk provides for controllable DDS, while also providing stabilization and protection to the embedded or sequestered (bio)molecules, preserving bioactivity [93]. The first step when developing a DDS using silk is loading the bioactive molecules, which can be achieved by bulk mixing, diffusion, or chemical conjugation. Water-soluble drugs or biomolecules can be dissolved with the silk, and the mixture processed into the desired material format. Drug capture and stabilization in the silk matrix depends on the number and strength of the interactions, such as hydrophilic or hydrophobic interactions, electrostatic interactions, hydrogen bonding, pi-pi interactions, or Van der Walls forces [80, 94–96]. When the drug-silk mixture is placed in an aqueous environment, drug is slowly released. The release of hydrophilic drugs is usually attributed to diffusion mechanisms through the amorphous domains of the silk carrier [97–99], whereas the crystalline domains act as physical barriers to diffusion which results in a more sustained release profile in some instances [96, 100, 101]. The influence of higher-order silk structures on drug binding and release has been extensively studied [94]. Hydrophobic drugs exhibit longer release profiles, usually explained by hydrophobic interactions between the drug and the silk matrix, along with the lower solubility in water.
Due to the inherent ability to self-assemble or to be chemically crosslinked, the silk proteins can be processed into numerous material formats [102], such as films [103], 3D-foams [104], hydrogels [105–107], or coatings [108] (among many others). An additional feature of recombinant silk materials, which exhibit many of the properties of natural silk [109] (tunable control of structure, chemical modifications, self-assembly, controllable mechanical properties, biocompatibility, biodegradability) is the possibility to alter the composition and introduce peptide ligands, epitopes, drug interacting sequences, and other protein-derived structures that extend their properties [110, 111]. These features help recombinant silks to stand out over other natural or synthetic materials.
The ability to expand and tune the properties of silks, including enhancing interactions with drugs, directed delivery of the drug towards a specific cell type, inclusion of unnatural amino acids for chemical modification, or the tailored degradation of the silk drug carriers, among other features, makes recombinant silks a promising platform for the development of DDSs. The controlled customization of recombinant silk offers significant improvement over native silk, allowing the introduction of unique and multifunctional domains into the silk sequence. These modifications can refine the already outstanding properties of native silk, not only overcoming its limitations but also providing for new applications.
3.1. Hydrogels
Silk hydrogels are a useful format for drug delivery because they can mimic the extracellular matrix (ECM) in structure, to positively support 3D cell cultures [112]. Hydrogels can be formed out of silk solutions by physical [113–115] or chemical crosslinking [116–118], or by a combination of both approaches [119]. Physical crosslinking can be induced by a local increase in protein concentration, through the use of organic solvents [120], acidification, changes in the ionic concentration of the solution [121], or by mechanical forces, such as shear or ultrasound [122–124]. These inputs favor the formation of nano- and microcrystals composed of hydrogen bonded beta sheets among silk molecules. Chemical crosslinking takes advantage of the chemically reactive groups present in silk, such as the hydroxyl groups on tyrosines. Examples of chemicals for crosslinking include ammonium peroxodisulfate (APS) [125] and tris(2,20-bipyridyl) dichlororuthenium(II) [117], or the oxidizing enzyme horseradish peroxidase (HRP) in combination with H2O2 [126, 127]. With this last methodology, tyrosine residues are oxidized with formation of di-tyrosine bonds. The advantage of using HRP to crosslink silk is that the reaction can be carried out under physiological conditions [128]. Moreover, chemical crosslinking modulates the mechanical properties [127]. Due to the mild conditions for HRP crosslinking, the inclusion of different drugs, bioactive molecules, and even living cells in the formulation can be achieved, the latter option leading to the development of live-cell DDS therapies using, for example, pluripotent stem cells or engineered T cells [129].
Hydrophilic drugs can be loaded into the recombinant silk hydrogels by different methods, such as direct loading, where the bioactive compound is mixed with the precursor solution prior to gelation, or introduced via diffusion post formation of the hydrogel. These strategies are not exclusive and can be combined to tailor the release profile of the drug. Hydrogels prepared from the recombinant spider silk eADF4(C16) were used to encapsulate biologically active compounds by direct loading and/or via diffusion in aqueous solvents [130]. Factors such as pore size influenced the release kinetics of model compounds. In the same work, recombinant silk hydrogel particles loaded with bioactive compounds in macroscopic silk hydrogels were used to influence the release profile of the bioactive compounds, extending them from days to weeks without affecting the bioactivity [130].
One of the problems with hydrophobic drugs relies on loading of the DDS. A novel gelation route was developed using eADF4(κ16) and eADF4(Ω16) recombinant spider silks and aqueous-organic solvent mixtures for 3D printing [80]. The co-solubilization of poorly water-soluble 6-mercaptopurine with organic solvents increased the amount of drug loaded in the gels, while the use of potassium phosphate induced rapid gelation of the system, for 3D printable drug-loaded hydrogel inks. In another example, co-precipitation using organic solvents and oil was used to fabricate recombinant eTuSp1 tubuliform spidroin silk from black widow spider microgels for a DDS with enzymatically-driven drug release [131]. The use of organic solvents increased loading capacity of doxorubicin while reducing beta-sheet formation when compared to other methods. The higher alpha-helix and lower beta-sheet content resulted in a lower resistance to proteolytic degradation and thus, proteinase-triggered release of drugs carried in the system.
Composite materials are useful in drug delivery systems via synergistic properties to address limitations and complementarity with the individual components, such as to increase the drug-loading capacity of the DDS. Antheraea assama silk fibroin (AaSF) was functionalized with a recombinant spider silk fusion protein (GN-4RepCT) bearing a fibronectin cell-binding domain to develop wound healing scaffolds [88]. The presence of the recombinant spider silk enhanced wound healing in an in vivo murine burn model, with increased vascularization and re-epithelialization when compared to the control. Moreover, the composite material enhanced the expression of collagen type I and III [88]. In another example, the processability of recombinant spider silk eADF4(C16) and the loading and release capacity of mesoporous silica nanoparticles were combined to develop antimicrobial-loaded composite hydrogels with antimicrobial properties, with drug release extending over 15 days [132].
The use of a recombinant approach for these materials allows expansion of the inherent properties by genetic sequence-level inclusion of genetic fusions, to generate (bio)active peptides or other protein sequences with the silk protein chains upon expression. Bioengineered RGD-modified variants of eADF3 and eADF4 recombinant spidroins were prepared and tested for antimicrobial properties [133] to demonstrate microbes were now unable to form biofilms on recombinant silk films, hydrogel surfaces, or within hydrogels. The bacterial and fungal repellent performance of these recombinant spider silk materials was related to higher-order structural features (i.e., secondary and tertiary) responsible for the formation of hydrophobic patches. The RGD-modified spider silk repelled microbes but allowed mammalian cell adhesion and proliferation [133]. This unique combination of properties (antibacterial, antifungal and cell-adhesive) in a single recombinant silk opens the possibility for new wound-healing and DDS, where the load of bioactive substances is focused on the improvement of the healing process and not on the prevention of infections. Silk has also been combined with thermoresponsive proteins for DDS. The temperature-driven aggregation of these proteins favors the formation of beta-sheets among the silk domains via physical crosslinking to form stable hydrogels. A chimeric protein was prepared which contained repeat units of resilin with the carboxy-terminal domain of the major ampullate spidroin 1 (MaSp1) of Nephila clavipes [134]. The resulting copolymer was thermoresponsive in water, forming reversible hydrogels at low temperatures and irreversible hydrogels in response to an increase in solution temperature. The mechanical properties of the hydrogel were dependent on the ratio of resilin to silk and on the presence of different salts. Using Rhodamine B as a model compound for drug release, the release profile was pH-dependent, with a slower release rate correlated with lower pH [134]. The recombinant fusion of silk and elastin-like protein (SELP) has also been used as a DDS. A SELP bearing lysine residues for controlled delivery of GM-0111, a modified glycosaminoglycan product of hyaluronic acid digestion and chemical sulfation, was developed and used to treat mucosal inflammatory diseases [71]. GM-0111 release from SELP was much slower compared with other thermally responsive polymers, which resulted in an increased bioaccumulation in the rectum. In a murine model, the SELP-GM0111 composite provided localized protection against radiation induced damage, reducing pain and animal mass loss.
3.2. Films
Film formats are a solid formulation for DDS. Where drug release is regulated by thickness, crystallinity, or layer-by-layer composites with other material [135–137]. Films can be applied directly as a device or used as a drug-reservoir coating with other devices [138]. The flexibility of films can be useful, such as for conformal coatings on irregular tissue surfaces to improve the consistent release of drug to the tissue, and this can be achieved by using plasticizers or the inherent properties of the silk [139, 140]. Plasticizers can impact the mechanical properties of the film/coating as well as the release profile of the drug [12].
A water-based method was developed to fabricate transparent thin films of recombinant spider silk protein eADF4(C16), with and without plasticizer (glycerol or 2-pyrrolidone) [141]. The soft processing conditions allowed direct loading of the films with small bioactive drugs (paracetamol and tetracaine HCl), complex sugars (dextran), or model proteins (lysozyme and bovine serum albumin). Drug release was dependent on the strength of interactions with the recombinant silk and molecular weight, with a slower rate of release for the higher the molecular weight of the compound. By combining different layers and coatings of silk, the release profiles of the different compounds could be tuned or controlled. A recombinant silk based on the C-terminal domain of the E. australis (4RepCT) self-assembled into fibrillar nanometer-thick membranes with elasticity and strong mechanical properties, with the films permeable to small molecules and proteins [142]. Moreover, these nanomembranes contained a fibronectin-derived motif, which supported human keratinocyte attachment and proliferation. The combination of biocompatibility, cell attachment, and drug/protein permeability provided biomolecular communication between the cell layer and the drug reservoir, with potential applications in drug delivery systems, drug studies, surgical transplants, and wound healing.
Covalently linking drugs to the recombinant silks through different bonds further expands the properties of these materials, particularly with labile chemical bonds towards stimuli-responsive DDS. pH- and redox-sensitive systems were developed based on the polycationic eADF4(C16) and the polyanionic eADF4(κ16) recombinant silk capable of triggering the release of drugs controlled by environmental conditions [37]. This mechanism was tested in different formats with 6-mercaptopurine, doxorubicin, 5,5′-dithiobis-2-nitrobenzoic acid (DTNB) and para-dimethylaminobenzaldehyde (DMAB), to support the efficiency and versatility of the methodology. Recombinant 4RepCT spider silk was also engineered to incorporate the non-natural methionine analogue L-azidohomoalanine for functionalization using azide-alkyne cycloaddition [143]. The recombinant silk was simultaneously chemically functionalized with different fluorophores, demonstrating multiple and simultaneous modifications of the recombinant silks. When decorated with a broad spectrum antibiotic via a labile linker, the recombinant silk demonstrated prolonged antibiotic activity [143].
Another interesting strategy is to develop recombinant silk-based systems that are capable of recruiting specific substances in situ that tailor or direct biological behavior. For example, films were prepared from recombinant eADF4(C16) spidroin containing tag sequences derived from non-collagenous proteins in bone known to bind to collagen and to initiate mineralization [144]. Calcium phosphate deposition and mouse preosteoblast adhesion were directed by the presence of these peptides, demonstrating that these films could initiate mineralization and tailor cell-adhesion. Films were also prepared using a 15mer derived from the consensus repeat of the Nephila clavipes drag line silk protein engineered with a hydroxyapatite binding domain at either the N- or C-termini [145]. Functionalization of the recombinant spider silk did not affect the assembly or mechanical properties, or the growth and proliferation of human mesenchymal stem cells. The presence of the hydroxyapatite binding domain increased the content of crystalline hydroxyapatite and the osteoinductive properties up to 3-fold when compared to the control. A recombinant silver-binding peptide was engineered into 6 or 15 repetitions of the consensus repeat of N. clavipes dragline silk as a template for nucleation and growth of silver crystals was prepared and inhibited the growth of Gram-positive and Gram-negative bacteria [73].
3.3. Sponges
There are several methods to fabricate 3D porous sponge structures from silk. The most widely used methods are freeze-thaw treatments and salt leaching, followed by a beta-sheet forming process, such as autoclaving [146, 147], immersion in ethanol/methanol [148, 149], or water vapor annealing [150]. Silks in the form of sponges are attractive formats for drug delivery, as they offer a high surface-to-volume ratio [151], which makes them very efficient in the delivery of hydrophobic drugs. When using silk, the release profile can be tuned by varying the crystallinity. Another important characteristic that makes sponges attractive is their capillary action, which facilitates the permeation of the release media into the sponge, and thus drug delivery [102].
Self-assembling mucoadhesive sponges using recombinant spider silk based on the 4RepCT protein were prepared for gastrointestinal drug delivery [152]. The recombinant 4RepCT silk protein was fused to either a human galectin carbohydrate recognition domain (hGal3) that specifically binds the Galβ1–3GlcNAc and Galβ1–4GlcNAc mucin glycans or to a polylysine tail. The incorporation of hGal3 resulted in stronger mucoadhesive properties when compared to the polylysine functionalized version due to its more specific interactions with mucin.
4. Systemic Delivery
Injectable silk-based DDSs have been used as a vehicle for the delivery of a variety of drugs, and most often for systemic delivery of chemotherapeutic agents. The majority of this research has been focused on the use of native silkworm material [153], however, recombinant silk proteins offer sequence-level control over material properties in a variety of formats, such as particles, spheres, and capsules.
DDSs, including nano-sized or micro-sized particles, spheres, and capsules, have been investigated due to their surface area, ease of systemic administration, and their ability to permeate cells and tumors [154]. Nanoparticles are generally defined by the range from 1 to 100 nm, while microparticles are 100 nm to 100 μm diameter [155]. Size is an important characteristic that affects drug delivery, with sub 100 nm particles showing increased uptake by mammalian cells depending on the particle shape and material composition [154]. While the size of a particle is important, several other properties govern the effectiveness of particle-based DDS, including size distribution, surface charge, biocompatibility, stability, and controllable drug release [154–156].
In addition to general physicochemical characteristics, particle-based DDSs must be tailored to the target disease to achieve therapeutic benefits. For example, passive targeting mechanisms can sometimes be adequate; numerous cancer treatment strategies have taken advantage of the enhanced permeability and retention (EPR) effect, a process by which macromolecules or nanoparticles enter the tumor microenvironment through intercellular gaps in blood vessels and are retained due to pressure [157]. Active targeting mechanisms are being increasingly investigated to improve drug specificity and limit off-target effects, with a large number of biological ligands capable of binding to cell receptors and facilitating uptake [158]. These advances are summarized in a detailed review of nanoparticle delivery to tumors, where it was shown that only 0.7% of the administered nanoparticle dose was successfully delivered to the target tumor [159]. DDSs must also be tailored to the drug itself, where properties like solubility, charge, toxicity, and degradation need to be considered. Recombinant silks offer a highly customizable DDS framework that can be tuned to meet these demands, as summarized in Figure.
4.1. Small Molecule Delivery
Small molecules make up the largest portion of current therapeutic drugs and are generally defined as organic molecules with a molecular weight below 1,000 daltons or smaller for oral delivery routes [160, 161]. The most widely studied small molecule drugs with respect to recombinant silk-based DDSs are chemotherapeutic agents [162]. These drugs make excellent candidates for encapsulation within a delivery system to reduce severe systemic side effects, which commonly include cardiovascular toxicity, pulmonary toxicity, nephrotoxicity, hepatotoxicity, and more recently severe cutaneous adverse reactions (SCARs) [163, 164]. Recombinant silk, particularly spidroin-based materials, have been studied as DDSs for doxorubicin, etoposide, mitoxantrone, and 6-mercaptopurine [35, 37, 153]. Due to the severity of off-target effects from these drugs, a primary focus has been the control of drug loading and release [155].
The majority of recent studies involving recombinant silk as a chemotherapy platform have focused on the delivery of doxorubicin. Doxorubicin is a hydrophilic anthracycline drug with a positively charged primary amine at physiological pH. Doxorubicin binds to and inhibits topoisomerase I and II preventing the separation of the DNA double helix and inhibiting replication, and also intercalates in DNA base pairs which further inhibits the process of replication and DNA synthesis [165]. However, like many chemotherapeutic agents, doxorubicin has severe off target effects. The most concerning side effect is cardiotoxicity, with a substantial risk of cardiomyopathy correlated with a lifetime accumulation of doxorubicin [165]. To reduce this risk, researchers have focused on encapsulation and targeted delivery using recombinant silk-based systems.
Recombinant spider silk has been studied extensively as a DDS for the delivery of doxorubicin in cancer therapy, with a focus on recombinant MaSp1 and MaSp2 variants. The targeted delivery of doxorubicin to human epidermal growth factor receptor 2 (HER2)-positive breast cancer cell lines using recombinant MaSp1 (MS1) microspheres was demonstrated [41]. Fusion of a Her2-binding peptide, denoted either H2.1 or H2.2, to the N-terminus of the MS1 sequence increased particle binding to Her2-positive cells in vitro when compared to the control MS1. When loaded with doxorubicin, H2.1- and H2.2-functionalized silk reduced cancer cell viability by up to 40% over the non-functionalized silk. Follow on studies investigated the role of the silk component in this DDS, comparing the particle size, morphology, zeta-potential, drug loading efficiency, and efficacy when the MS1 sequence was exchanged for a sequence derived from MaSp2 (MS2) [55, 166]. Particles produced with the MS2 variants (fused with H2.1, H2.2, or non-functionalized) were smaller, more spherical, and less prone to aggregation than the MS1 counterparts, but displayed lower drug-loading efficiency. Blending of the two functionalized silks resulted in particles with superior intermediate properties and a decrease in non-specific toxicity, highlighting the importance of silk identity to physicochemical and therapeutic properties [166].
The efficacy of H2.1-functionalized MS1 silk was tested in vivo in a murine Her2-positive breast cancer model [167]. The findings were similar to the in vitro studies performed with human Her2-positive cell lines; the H2.1MS1 silk particles loaded with doxorubicin had a selective cytotoxic effect on Her2 positive mouse cancer cells, and significantly reduced tumor volume over 20 days in a dose-dependent manner. Additionally, there appeared to be significantly lower off-target cytotoxicity with this DDS compared to unencapsulated doxorubicin, as the weight of mice was unaffected by the loaded MS1 sphere treatment. In a separate study, the unloaded MS1 spheres showed no toxicity when administered in a healthy mouse model [168]. Although particle biodistribution was affected by the presence of the H2.1 targeting peptide, with significantly more particles observed in organs compared to the non-functionalized MS1 spheres, histopathological examination of the organs indicated no morphological changes. The combination of these two studies is a significant advance in the application of recombinant silks for cancer treatment and could pave the way for early clinical trials in the future.
A consistent challenge for many DDSs, including those formulated with recombinant silks, is to control drug release until it has reached its target. To address this issue, an MS2 silk was bioengineered to contain a doxorubicin-binding peptide (DOXMS2) to enhance drug encapsulation and reduce leakage in the circulatory system [40]. By blending the DOXMS2 with the H2.1MS1 silk described previously [166], selective delivery of the particles to Her2-positive cancer cells was demonstrated in vitro. Furthermore, the DOXMS2 improved initial loading of doxorubicin by 12% and reduced release by 23%, 28%, and 30% (pH 4.5, pH 6, and pH 7.4, respectively). Another strategy to control drug binding and release is to manipulate electrostatic interactions between the carrier and the drug [169]. This strategy was recently employed to study the binding and release of positively charged doxorubicin [35]. The researchers modified the sequence of MS2, introducing one glutamic acid residue to each repeating unit to obtain negatively charged EMS2 silk with a lower isoelectric point (pI = 3.15 vs 5.27). The loading efficiency of doxorubicin into EMS2 particles was unchanged compared to MS2, however the EMS2 particles released doxorubicin at a faster rate with a dependence on pH. Using silk spheres consisting of recombinant tubuliform spidroin (eTuSp1) from the Latrodectus mactans (black widow) spider, the positive charge of doxorubicin was again utilized to control drug release [89]. Histidine residues were introduced into the silk sequence to achieve a pI of 4.8, resulting in a positively charged molecule at lysosomal pH (~4.5). The spheres displayed tight control over release in a pH-dependent manner, with 15% and 20% cumulative release at pH 7.4 and 6.5, respectively, and greater than 70% cumulative release at pH 4.5.
In addition to electrostatic interactions, another attractive strategy to control drug binding and release is the direct covalent attachment of the drug to the carrier. There are several examples of this strategy being applied to elastin-like proteins [170, 171], however only one instance of this for silk-based systemic delivery of small molecule drugs [37]. In this study, doxorubicin was chemically conjugated to the surface of silk particles from eADF4(C16), an engineered polyanionic spidroin from Araneus diadematus (European garden spider), using a pH-labile hydrazine linker. The conjugated particles showed little drug release for 48 hours at pH 7.4; when the pH was reduced to 4 to trigger drug release, 100% cumulative release was achieved over the next 24 hours. In vitro experiments using hydrazine-linked fluorescent dyes indicated that release was triggered within 16 hours of internalization.
One group of alternative materials that has been studied for small molecule delivery are SELPs. There are several examples of nanoparticles developed using SELPs without drug encapsulation; varied ratios of B. mori silk fibroin motifs to elastin motifs to achieve spherical micellar-like particles [68], and the effects of various charged and polar guest residues within elastin blocks to form mucoadhesive nanoparticles [70]. The encapsulation of doxorubicin within SELP nanoparticles was also reported [69]. The SELPs delivered a cytotoxic dose of doxorubicin to HeLa cells in vitro, with one formulation (S4E8Y, with a 1:2 ratio of silk blocks to elastin blocks) performing better than free doxorubicin alone. To improve doxorubicin release from self-assembling SELP nanoparticles, a matrix metalloproteinase-responsive peptide (MMP) was encoded within the SELP sequence [42]. Insertion of the hydrophilic MMP sequence into the elastin region did not impair particle assembly and showed a correlation between extent of assembly and distance of the insertion site from the hydrophobic silk. Similarly, the release of doxorubicin from the particles was faster when the MMP sequence was further from the hydrophobic silk blocks. However, incubation with matrix metalloproteinase-9 (MMP-9) had no effect on doxorubicin release in any conditions tested, with the conclusion that doxorubicin release was driven entirely by water penetration and dissolution of doxorubicin from SELP assemblies. Although matrix metalloproteinase-dependent release studies have shown promising results in other DDSs [172], it seems further study is required to translate this efficacy to SELP formulations.
Several studies have demonstrated the flexibility of recombinant silks as a drug carrier by testing other small molecule formulations. Of note, recombinant spider silks have been studied for their efficacy in the encapsulation and delivery of three other chemotherapeutics, etoposide, mitoxantrone, and 6-mercaptopurine. Similar to doxorubicin and other chemotherapeutic agents, a major function of these drugs is to inhibit DNA synthesis and replication to prevent proliferation of cancer cells, with additional mechanisms of action relating to cell metabolism, cytokine secretion, and apoptosis [173–175]. Due to their systemic toxicity, these therapeutics can similarly benefit from encapsulation and controlled release. Microspheres composed of either MS1 or MS2 recombinant spider silk were analyzed for their efficacy in loading and release of mitoxantrone and etoposide [176]. The MS2 spheres were superior for the loading and release of both mitoxantrone and etoposide, correlated with a lower hydrophobicity and greater negative zeta-potential. To further investigate this relationship, a microsphere DDS was developed using the recombinant silk EMS2, a negatively-charged variant of MS2 containing glutamic acid residues with an even greater negative zeta-potential [35]. Loading efficiency of both neutral etoposide and positively charged mitoxantrone was higher in the EMS2 spheres than the unmodified MS2. Drug release was pH-dependent for both drugs regardless of silk composition; however, mitoxantrone exhibited a fast burst release from EMS2 that limited its potential application, while etoposide exhibited sustained release from EMS2 spheres up to 100% of the encapsulated drug (at pH 4.5). In the case of 6-mercaptopurine, the triggered release of the drug when coupled to cysteine-tagged eADF4(κ16) particles was demonstrated using a redox-sensitive disulfide bond [37]. Release of 6-mercaptopurine was rapidly induced over 30 minutes in 5 mM glutathione (GSH), mimicking the intracellular redox environment. However, in vitro release of a similarly coupled fluorophore was more persistent, only observed after 8 hours within HeLa cells.
Another class of small molecule drugs for drug delivery are antibiotics. With the rise of multidrug-resistant bacterial infections, new strategies have become necessary to combat these pathogens. Nanomaterials, such as core-shell nanoparticles, micelles, and liposomes, have been studied extensively as a potential strategy to deliver antibiotics systemically and with targeted efficacy to overwhelm drug-resistant bacteria [177–181]. Recombinant silk DDSs are an intriguing platform for the targeted delivery of antibiotics due to the ease with which peptides can be incorporated, either for targeting or for direct antimicrobial activity; however, there has only been limited study on this approach to date. One example is the smart delivery of vancomycin using infection-responsive MS2 nanospheres [182]. Vancomycin is currently used in the treatment of serious infections that have been unresponsive to other antibiotics, such as methicillin-resistant Staphylococcus aureus [183]. To combat this pathogen, functionalized MS2 nanospheres with a thrombin-sensitive peptide (TSP) were prepared and studied, leveraging studies that indicate an increase in thrombin-like activity during S. aureus infections due to its secretion of staphylocoagulase [184, 185]. In vitro analysis of vancomycin-loaded TSP-conjugated particles incubated with S. aureus exudate showed a significant increase in drug release compared to the unconjugated MS2 particles (85% vs 15% cumulative release, respectively). In vivo analysis of the drug-loaded, TSP-conjugated particles in a murine septic arthritis model demonstrated improved bacterial clearing. Although this application was restricted to the chemical conjugation of TSP, these infection-responsive peptides could be included directly in the genetic sequence of recombinant silk proteins in future studies.
4.2. Biomacromolecule Delivery
While small molecule drugs have historically played a large role in disease treatment, there are many so-called “undruggable” diseases that have not responded to these treatments [186, 187]. Biomacromolecules such as nucleic acids and proteins are an alternative class of therapeutics that have increasingly become the focus of studies in the treatment of these diseases, due in large part to rapid advances in genetic engineering. Nucleic acids can be utilized in a variety of ways; as plasmid DNA (pDNA) and messenger RNA (mRNA) to express genes locally as part of therapeutic protein replacement strategies [188, 189], small interfering RNA (siRNA) and micro RNA (miRNA) to knock down harmful protein expression [190–192], as well as aptamers that can provide highly specific targeting capabilities [193, 194]. Peptide- and protein-based therapeutics can interact directly with target cells, providing specific cellular functions or regulating biological processes, and have already grown into a large global market [195–197]. However, like small molecule drugs, biomacromolecular therapeutics come with their own challenges for drug delivery. Factors such as high molecular weight and charge can prevent these molecules from permeating the cell membrane, and they are highly susceptible to enzymatic cleavage by endonucleases or proteases [23, 198]. As such, DDSs can be used to mitigate these challenges by providing a protective effect and facilitating transport across the membrane, and recombinant silks offer some attractive properties in this regard.
One way that recombinant silk can be used to deliver nucleic acids is by using polycationic blocks to interact with and mask the negatively charged DNA. Poly(l-lysine) blocks can be encoded in the recombinant sequence to provide a positive charge at physiological pH, as demonstrated [199]. Six repeats of the MaSp1 consensus repeat sequence were fused to either 15, 30, or 45 consecutive lysine residues on the C-terminus. When complexed with pDNA by coincubation, the lysine-containing silks formed particles in the range of 250–500 nm. These particles were noncytotoxic (with the exception of silks containing the 45-lysine block, which reduced cell viability by ~15%) to human embryonic kidney (HEK) cells and were capable of transfecting them with green fluorescent protein (GFP)-expressing pDNA with limited efficiency when the particles were deposited onto a silk film. Inclusion of multiple cell-binding RGD motifs at the N-terminus further increased transfection efficiency while simultaneously reducing the average particle diameter to 186 nm in solution [200]. Replacement of the RGD motifs with a dimeric ppTG1 peptide, a lysine-rich cell penetrating peptide (CPP) that enhances translocation, improved transfection efficiency further to a level comparable to the commercial transfection reagent Lipofectamine 2000 [201]. To increase the therapeutic relevance of silk-pDNA complexes, tumor-homing peptides (THPs) can be fused to the sequence as well. Incorporation of the F3 or CGKRK THPs enabled the targeted transfection of MDA-MB-435 and MDA-MB-231 cells both in vitro and in vivo in a murine xenograft model [202]. Building on this foundational work, the functionality of MaSp1-polylysine nanoparticles was expanded by investigating their efficacy for delivering pDNA to human mesenchymal stem cells (hMSCs) [203]. By fusing the nuclear localization sequence of the large tumor antigen of the Simian virus 40 (SV40) and a dimerized hMSC high affinity binding peptide (HAB) to MaSp1 silk, transfection of hMSCs with GFP-expressing pDNA at 50% of the efficiency of Lipofectamine 2000 was demonstrated.
Therapeutic protein and peptide delivery has largely been focused on native silk fibroin strategies. However, there have been some notable advances in the use of recombinant silks, which benefit from the ease with which peptide sequences can be added directly to the silk sequence. The protective benefits of encapsulating an active protein or peptide are clear, as demonstrated utilizing eADF4(C16) spider silk to encapsulate β-galactosidase [204]. Using an aqueous emulsion in silicon oil, capsules ranging from 1 to 30 μm in diameter were formed that contained the active enzyme and resisted protease degradation. Although this study did not focus on therapeutic drug delivery, demonstrating the proof of concept was an important step in biomacromolecule delivery. eADF4(C16) silk was utilized as a DDS to protect the ovalbumin epitope from degradation and deliver it to antigen-presenting dendritic cells as a part of a novel vaccination strategy [78]. The OVA257–264 peptide (SIINFEKL), which is routinely used in vaccine studies, was fused to eADF4(C16) via a cleavable linker (cathepsin B or cathepsin S), and these recombinant proteins were processed into microparticles. The particles were preferentially taken up by dendritic cells in vitro and did not affect their viability or induce a proinflammatory response. When fused to the core silk sequence via cathepsin S, the particles induced strong cytotoxic T-cell proliferation both in vitro and in vivo using a murine model. When coupled with the stability of the particle suspensions and the simplicity of manufacturing, these findings present a promising new DDS for peptide vaccination using recombinant spider silks.
In addition to spider silks, recombinant silk sericin has been used effectively for the delivery of protein-based therapeutics. Ulcerative colitis was treated using an oral DDS containing recombinant human lactoferrin (rhLF) [205]. Lactoferrin is an iron-binding glycoprotein with antibacterial properties that plays an essential role in the human immune system, and has been studied as a potential therapeutic for the prevention and treatment of inflammatory bowel diseases [206]. To generate rhLF-functionalized silk, researchers hybridized previously developed recombinant silkworm strain (D9L-rhLF) [207] with another silkworm strain (cottony cocoon) that produces larger sericin-rich cocoons with a greater yield of rhLF. Nanospheres prepared from this recombinant sericin provided a protective environment for the rhLF, with 75% remaining after 24 hours of incubation in simulated gastric fluid and 42% remaining after a similar incubation with artificial intestinal fluids. Uptake and bioavailability of the rhLF nanospheres was significantly higher than free rhLF both in vitro and in vivo (murine model), and oral administration of the nanospheres reduced dextran sulfate sodium-induced colitis in the mice.
4.3. Inorganic Therapeutics
The earliest scientific description of inorganic nanoparticles was by Michael Faraday in 1857 as the result of a reduction of gold chloride, making these one of the longest-studied nanomaterials [208]. As of 2019, there were 25 inorganic nanomedicines on the market for the treatment of diseases like cancer, diabetes, bacterial infections, anemia, dental caries, and many more [209]. In addition, inorganic nanoparticles are uniquely beneficial for diagnostic and theranostic applications due to their optical, electronic, and thermal properties, in addition to the benefits shared by organic nanoparticle systems [210, 211]. To facilitate the binding of targeting moieties, increase biocompatibility, and to prevent aggregation, inorganic nanoparticles can be coated by organic materials such as silk. Although the benefits of recombinant DNA technology are clear with regard to the facile fusion of metal binding domains and targeting peptides, few studies have been conducted using recombinant silks as the field continues to develop.
Magnetic iron oxide nanoparticles (IONPs) are one such DDS that have been investigated in combination with recombinant silk materials. IONPs are one of the most studied inorganic nanomedicines, with applications as contrast agents in techniques such as magnetic resonance imaging (MRI) and magnetic particle imaging (MPI) [210, 212–214], and as therapeutic agents themselves for hyperthermia-based treatments or magnetically guided drug delivery [215, 216]. Bioengineered spider silks have been used in conjunction with IONPs; the recombinant silks MS1, MS2, and EMS2 were investigated for their ability to form microspheres in the presence of IONPs [217]. The IONPs did not interfere with sphere formation, and although IONPs were associated with each of the three silks investigated the EMS2 variants were able to bind significantly more of the material. The IONPs were localized to the surface of the silk spheres and exhibited good magnetic properties, although the magnetic properties were not investigated for therapeutic applications such as hyperthermia. Instead, the authors investigated the composite spheres for the loading of doxorubicin and found that the presence of IONPs in the EMS2/IONP composites enhanced drug loading efficiency more than two-fold while also slowing the drug release kinetics at a pH above 4.5. Building on this foundational work, bioengineered spider silks with metal-binding and targeting moieties were pursued [218]. MS1 silks were functionalized with an Fe1 adhesion peptide, which enhanced the binding of negatively charged IONPs, and this MS1Fe1 silk was blended with the previously described H2.1MS1 silk to provide targeting to Her2-positive cancer cells. Doxorubicin loading in the composite particles was enhanced by the presence of IONPs (1.3-fold increase) with a similar pH-dependent release profile as observed for the EMS2/IONP composites. In vitro cytotoxicity analysis with Her2-positive SKBR3 cells showed that the IONP blended spheres containing doxorubicin reduced cell viability slightly more than those without IONPs. However, the most striking results were seen with the use of a magnetic field to stimulate hyperthermia, which resulted in an almost 3-fold increase in cells in the late phase of apoptosis.
Several other inorganic nanoparticles have been investigated in conjunction with recombinant silks, including selenium, gold, and manganese carbonate, although their therapeutic applications are less developed than the previous IONP examples. Selenium nanoparticles (SeNPs) are of interest due to their antibacterial properties coupled with a low cytotoxicity for human cells, and have been explored as a new type of antibiotic to combat MDR pathogens [4–6]. SeNPs were coated with positively charged eADF4(κ16) recombinant spider silk and investigated for antibacterial properties [219]. The positively charged composite particles showed strong antibacterial effects by reduction in colony forming units on agar plates, with a minimum bactericidal concentration (MBC) against E. coli of 8 ± 1 μg/mL. The cytotoxicity of these particles was evaluated using Balb/3T3 mouse embryo fibroblasts and HaCaT human skin keratinocytes and were found to be noncytotoxic up to concentrations of 31 μg/mL. Although these findings are promising, further study is required to investigate the therapeutic use of these particles as a DDS. Gold nanoparticles (AuNPs) share many properties with IONPs, and have a unique property known as localized surface plasmon resonance (LSPR) that can generate heat when exposed to resonant frequency photons and be used as a form of photothermal therapy [211]. Coating of Ni-NTA functionalized AuNPs with recombinant SELPs with an N-terminal His-tag was investigated [220]. These composite nanoparticles displayed thermally reversible aggregation when heated to 60°C, which also shifted the plasmon resonance, however the goal of this study was not therapeutic in nature. Manganese carbonate nanoparticles have also been explored as potential photothermal therapy agents and MRI contrast agents [221, 222]. Although recombinant eADF4 silk has recently been shown to enhance manganese carbonate mineralization, there has been no direct investigation of these composites for therapeutic drug delivery [223].
5. Engineering Silkworms for Recombinant Protein Expression
Currently, there is an increasing demand for therapeutic proteins with reduced costs. With the development of advanced tools for genetic manipulation, the silkworm silk gland can be engineered as a cost-effective bioreactor for the expression of recombinant proteins along with the synthesis of the native silk proteins [224]. Bombyx mori is one of the most attractive choices for the production of recombinant proteins due to the high protein production, low feed costs, short generation time, rich genetic resource and insight, and the feasibility of post-translational modifications. The first use of B. mori as a bioreactor used a baculovirus as the transfection vector for the expression of α-interferon [225].
Recombinant proteins can be co-expressed in the silk gland [207, 226, 227], or fused to a single chimeric silk molecule [228], with piggyBac as one of the most common vectors used for transfection [229, 230]. Fertilized B. mori silkworm eggs are transfected with modified versions of the piggyBac vector containing the recombinant protein gene together with a helper plasmid carrying the piggyBac transposase gene. Due to the stochastic insertion of the recombinant gene, successful insertion in the transgenic silkworm is confirmed in the next generation using reporter genes, such as the DsRed, which induces the expression of a red fluorescent protein in silkwormś eyes. After successful transfection a recombinant protein can be continuously produced after a transgenic silkworm strain is established [224]. Recombinant protein can either be co-expressed or fused to the silk backbone sequence. When co-expressed, the bioactive compound remains embedded in the silk matrix and is released due to diffusion, degradation, or proteolysis of the matrix, a characteristic also controllable via fusion of a matrix metalloprotease with the fibroin structure [231, 232]. When expressed as fusion proteins, they are inherently a structural part of the matrix so long as the fused domains do not interfere with the hierarchical assembly of silk fibers (Error! Reference source not found.).
A recombinant protein comprising the basic fibroblast growth factor (bFGF)-binding P7 peptide was introduced at the C-terminus of the fibroin light chain [233]. A transgenic variant expressed the recombinant version in 26% of the light chains, exhibiting high bFGF binding affinity. Human epidermal growth factor was fused to a truncated version of the B. mori fibroin heavy chain [234]. The transgenic silkworm cocoon silk significantly increased human fibroblast cell proliferation with no apparent cytotoxicity, demonstrating potential as wound dressing material. The use of recombinant proteins permits the incorporation of other proteins, to expand the bioactivity of these chimeric silks. To increase the biocompatibility of silk, and using transgenic silkworms as the recombinant factory, partial sequences of collagen or fibronectin were inserted into the fibroin light chain [47]. The recombinant silks possessed higher cell-adhesive activity when compared to the unmodified silk.
Sericin is the other major component of B. mori silk and is also amenable to being engineered using recombinant techniques. B. mori silkworms were engineered to co-express acidic fibroblast growth factor (FGF1) together with sericin [235]. Degumming of the silk fiber resulted in FGF1-loaded sericin which formed porous hydrogels upon cooling. The FGF1-loaded sericin was injectable and exhibited relevant mechanical properties while helping retain the bioactivity of the FGF1. The sustained release of FGF1 promoted cell proliferation in the sericin hydrogels, and based on the absence of inflammatory responses, demonstrated biocompatibility and no immunogenicity. Following a similar approach, a sericin hydrogel loaded with recombinant human platelet-derived growth factor (PDGF-BB) was developed [228]. Upon degumming the silk fibers, the PDGF-BB-loaded sericin formed hydrogels with similar mechanical properties to chemically crosslinked sericin, with a 13.1% content of the growth factor. Sericin hydrogels slowly released the PDGF-BB while retaining its bioactivity for more than 42 days, and was biocompatible, exhibiting no immunogenicity (in vitro and in vivo) and supporting the differentiation of osteoblastic and human mesenchymal stem cells. A strategy to deliver recombinant human lactoferrin (rhLF) using transgenic silkworms was also developed [207, 236]. The genetically engineered silkworms co-expressed rhLF that was secreted together with sericin and incorporated into the silk fibers. Hydrogels prepared from this rhLF-silk sericin blend were biocompatible and biodegradable, and this DDS demonstrated a protective effect on the rhLF in the gastrointestinal tract for efficient delivery in mice [236].
Future Outlook
Polymer-based DDSs have been attracting research interest due to their controllable properties. However, the drawbacks of synthetic polymers (broad molecular weight distribution, issues of biocompatibility, use of organic solvents, and inflammatory degradation products) have directed attention towards other types of materials. As discussed in this review, natural polymers such as silk and recombinant silk-like proteins have beneficial properties that can address these shortcomings. In addition to the examples documented here, there are several areas that warrant further investigation. Lesser studied silks such as those from the order Hymenoptera (ants, bees, and wasps) offer some interesting properties, such as a high mechanical strength (sometimes surpassing B. mori silk fibroin) and a non-repetitive alpha-helical structure [17–20]. Several of these silks have also been demonstrated to be highly expressed at full length in E. coli at much greater titers than many recombinant silkworm or spider silks, which are often challenging to express due to their repetitive sequence. Exploration of these alternative silks can expand the portfolio of potential materials for DDS, and may help to address challenging designs. The incorporation of multiple functional domains into one DDS is another exciting candidate for further research. Several recent studies have demonstrated the benefits of this multifunctionality, and with the continual advancements seen in genetic engineering and modeling we can better predict complex interactions and optimize them for drug delivery.
Although the use of recombinant silk materials for DDSs shows great promise, challenges remain that hinder translation to clinical trials and beyond. No recombinant silk DDS has yet been clinically approved. As is often the case with recombinant proteins, the scalability of silk materials is currently limited and often exacerbated by complications with heterologous expression of repetitive sequences. For B. mori silks, the direct engineering of silkworms may be a viable solution to scale-up, however this comes with its own challenges such as low genetic stability and low expression of transgenic products.
Genetic engineering advances are primed to improve therapeutic outcomes of delivered drugs while minimizing off-target effects. The different silk types, each with specific characteristics, in combination with on-demand tunable properties of recombinant silks, offers a wide pool of materials and properties for DDS designs. The inclusion of specific disease-targeted peptides, as well as on demand drug release from the DDS, can help to customize the DDS to increase efficiency. With emerging personalized therapies, recombinant systems are expected to become a key to success. In total, with silk proteins or domains from these proteins as core building blocks, the versatility and impact of genetically modified silks for DDS offers immense potential in the future as additional fundamental and translational studies are pursued. In addition, with growing awareness towards animal welfare, the opportunity to generate recombinant silks towards DDS goals should also increase in importance.
Figure 1.

Silk-based drug delivery systems. Drugs such as small molecules, nucleic acids, protein biologics, and inorganic nanoparticles can be loaded into silk through diffusion, mixing, or chemical conjugation. The silk, which can be processed into various formats (hydrogels, films, sponges and particles) stabilize loaded drugs through hydrogen bonds, electrostatic interactions, pi-pi stacking, and van der Waals forces. The structural content of the silk can be tuned to control drug release through both diffusion and proteolytic degradation mechanisms, where higher beta sheet content creates resistance to both diffusion of the drug as well as protease activity.
Figure 2.

Overview of sequence modifications used for particle-based systemic drug delivery. Modifications are grouped as: 1) cargo carrying domains which interact specifically with a therapeutic agent, 2) side chain modifications through incorporation of additional amino acid residues, 3) cell binding domains to interact with various cell types in a non-specific or targeted manner, and 4) the core backbone sequence made up of one or more types of silk or elastin motifs. Abbreviations: RGD (tripeptide Arg-Gly-Asp), CPP (cell penetrating peptide), NLS (nuclear localization sequence), THP (tumor homing peptide), NP (nanoparticle). Created in part with BioRender.com.
Figure 3.

Schematic of the silkworm-based strategy for recombinant protein coexpression. A) Foreign genes are inserted into fertilized eggs of B. mori silkworms using recombinant vectors, the most common of which is the piggyBac vector, or different recombinant techniques such as Zinc Fingers, TALENs or CRIPSR. B) Transgenic worms are selected in the following generation using reporter gene expression, such as fluorescent proteins expressed in the eyes. C) Recombinant proteins can be either co-expressed (grey octagons) or fused to the backbone of both sericin and fibroin (grey stars). When co-expressed, recombinant protein remains embedded in the silk matrix, while if expressed as a fusion protein they are part of the matrix itself. In both cases, sericin and fibroin can be then processed into the desired DDS.
Table 2:
Examples of recombinant silks fused to small peptides/factors, recombinant silk copolymers, recombinant silk composites and chemically functionalized recombinant silks used as DDSs.
| Refs | ||||
|---|---|---|---|---|
| N-terminal Histidine tag | 10xHis | SELP | SELP binding to Ni-NTA AuNP for nanoparticle photothermal therapy or MRI contrast agents | [220] |
| Histidine residues | 10xHis | Recombinant tubuliform spidroin (eTuSp1) | Controlled, pH-dependent release of doxorubicin | [89] |
| Positively or negatively charged C-terminal tag | 8xLys or 8xGlu | eADF4(c16) silk from A. diadematus fibroin | Enhanced carbonate nanoparticle mineralization for DDS | [223] |
Iron binding peptides
|
KSLSRHDHIHHH SVSVGMKPSPRP |
MaSp1 protein of N. clavipes | Enhanced inorganic nanoparticle adhesion | [218] |
OVA257–264 peptide Cleavable linkers:
|
SIINFEKL GFLG PMGLP |
eADF4(C16) from A. diadematus fibroin | Microparticle encapsulation of bioactive enzymes for protection, with antigen-specific delivery fused via a cleavable linker | [78] |
| Poly-lysine + integrin binding domain | 30xLys + RGD | MaSp1 protein of N. clavipes | Nanoparticle DNA Polyplexes Human embryonic kidney (HEK) cell transfection with GFP |
[200] |
| Poly-lysine + Dimeric ppTG1 cell penetrating peptide | 30xLys + (GLFKALLKLLKSLWKLLLKATS)2 | MaSp1 protein of N. clavipes | Nanoparticle DNA polyplexes for HEK cell transfection with GFP, with efficiency similar to Lipofectamine 2000 | [201] |
| Poly-lysine + Tumor-homing peptides | 30xLys + KDEPQRRSARLSAKPAPPKPEPKP KKAPAKK or (CGKRK)n, n=1 or 2 |
MaSp1 protein of N. clavipes | Nanoparticle DNA polyplexes for transfection of MDA-MB-435 and MDA-MB-231 cells (in vitro and in vivo-murine model) | [202] |
| Poly-lysine + SV40 nuclear localization sequence + hMSC high affinity binding peptide + Cell-penetrating peptide | LysX15 + PKKKRKV + SGHQLLLNKMPNGGGS + PLSSIFSRIGDP | MaSp1 protein of N. clavipes | Nanoparticle DNA polyplexes with 50% transfection efficiency compared to Lipofectamine 2000 | [203] |
| Her2-binding peptide | MYWGDSHWLQYWYETS | MaSp1 protein of N. clavipes | Microsphere targeted delivery of doxorubicin to (HER2)-positive breast cancer cell In vivo model:
|
[41] [168] [167] |
| Her2-binding peptide | MYWGDSHWLQYWYETS | MaSp1 and MaSp2 protein of N. clavipes | Microsphere targeted delivery of doxorubicin to (HER2)-positive breast cancer cell. Blend of both recombinant silks results in superior intermediate properties and decreased non-specific interactions. | [55, 166] |
| Her2-binding peptide + Doxorubicin-binding peptide (DOXMS2) | MYWGDSHWLQYWYETS + MASLWSPWYGGSWTS | MaSp2 protein of N. clavipes | Microsphere targeted delivery of doxorubicin to (HER2)-positive breast cancer cell; increased doxorubicin loading and sustained release | [40] |
| Matrix metalloproteinase-responsive peptide (MMP) | GPQFIGQ | SELPs: 815K, 815K-RS1, 815KRS2 and 815K-RS5 |
Improved release of doxorubicin
|
[42] [69] |
| Fibronectin cell binding motif | TGRGDSPA | Recombinant silk based on the C-terminal domain of E. australis: 4RepCT | Blended with A. assama silk fibroin for wound healing in third-degree burn (rat model) | [88] |
| Fibronectin cell binding motif | TGRGDSPA | Recombinant A. diadematus silk fibroin: eADF4(C16), eADF4(C16)-RGD, eADF4(C32NR4), eADF3(AQ)12 and eADF4(Ω16) variant. Recombinant B. mori fibroin |
Preparation of hydrogels, flat and patterned films, with antimicrobial properties and improved mammalian cell binding properties | [133] |
| Fibronectin cell binding motif | TGRGDSPA | Recombinant silk based on the C-terminal domain of E. australis: 4RepCT | Nanometer-thick membranes with outstanding elasticity Permeable to small molecules (Rhodamine B, Texas Red-Dextran) and proteins (FITC-BSA) Allows biocomunication between cell layer and drug reservoir |
[142] |
| Non-collagenous proteins from bone known to bind to collagen and to initiate mineralization | Osteo-tag: GLRSKSKKFRRPDIQYPDATDEDI TSHM SialoN-tag: DSSEENGNGDSSEEEEEEEETS SialoC-tag: EDESDEEEEEEEEEE |
Recombinant eADF4(C16) from A. diadematus fibroin | Enhanced calcium phosphate deposition | [144] |
| Hydroxyapatite binding domain (VTK peptide) | VTKHLNQISQSY | 15mer derived from the consensus repeat of N. clavipes dragline silk protein | N’ and C’ terminal fusion of the VTK peptide Enhanced biomineralization in vitro | [145] |
| Silver binding peptides | NPSSLFRYLPSD WSWRSPTPHVVT |
6 or 15 repetitions of the consensus repeat of N, clavipes dragline silk | Inhibited growth of Gram positive and Gram negative bacteria | [73] |
|
PLIVPYNLPLPGGVVPRMLITILG TVKPNANRIALDFQRGNDVAFHFN PRFNENNRRVIVCNTKLDNNWGRE ERQSVFPFESGKPFKIQVLVEPDH FKVAVNDAHLLQYNHRVKKLNEIS KLGISGDIDLTSASYTMI |
Recombinant spider silk 4RepCT from E. australis | Increased muchoadhesive properties | [152] |
|
|
Composite of B. mori fibroin inspired recombinant silk and fibroin extracted from B. mori | Implantable hydrogels for accelerated cell infiltration and Increased blood vessel formation in an in vivo rat model | |
|
PLLQATLGGGS LEVLFQGP |
Transgenic B. mori silkworm fibroin fusion protein | Recombinant B. mori silk fibroin with the ability to specifically find bFGF growth factor bFGF fused at the C-terminus of fibroin light chain | [233] |
| Human epidermal growth factor (hEGF1) | NSDSECPLSHDGYCLHDGVCMYIE ALDKYACNCVVGYIGERCQYRDLK WWELR |
Transgenic B. mori silkworm fibroin fusion protein | hEGF-functionalized truncated cocoon fibroin heavy chain with increased cell proliferation for wound dressing applications | [234] |
| Partial collagen sequence Partial fibronectin sequence |
[GERGDLGPQGIAGQRGVV(GER)3 GAS]8GPPGPCCGGG [TGRGDSPAS]8 |
Transgenic B. mori fibroin light-chain fusion protein | Films with increased mammalian cell-adhesive properties | [47] |
| Acidic fibroblast growth factor (FGF1) | MAAGSITTLPALPEDGGSGAFPPG HFKDPKRLYCKNGGFFLRIHPDGR VDGVREKSDPHIKLQLQAEERGVV SIKGVCANRYLAMKEDGRLLASKC VTDECFFFERLESNNYNTYRSRKY TSWYVALKRTGQYKLGSKTGPGQK AILFLPMSAKS |
Co-expression with B. mori sericin | Self-gelling and injectable sericin hydrogels loaded with FGF1 with increased cell proliferation and the absence of immunogenicity Sustained release of FGF1 |
[235] |
| Human platelet-derived growth factor (PDGF-BB) | MNRCWALFLSLCCYLRLVSAEGDP IPEELYEMLSDHSIRSFDDLQRLL HGDPGEEDGAELDLNMTRSHSGGE LESLARGRRSLGSLTIAEPAMIAE CKTRTEVFEISRRLIDRTNANFLV WPPCVEVQRCSGCCNNRNVQCRPT QVQLRPVQVRKIEIVRKKPIFKKA TVTLEDHLACKCETVAAARPVTRS PGGSQEQRAKTPQTRVTIRTVRVR RPPKGKHRKFKHTHDKTALKETLG A |
Co-expression with B. mori sericin | Temperature-induce gelation of sericin with a 13.1% w/w content of PDGF-BB Sustained release of PDFG-BB for over 30 days Increased cell proliferation |
[228] |
| Recombinant human lactoferrin (rhLF) | MKLVFLVLLFLGALGLCLAGRRRR SVQWCAVSQPEATKCFQWQRNMRR VRGPPVSCIKRDSPIQCIQAIAEN RADAVTLDGGFIYEAGLAPYKLRP VAAEVYGTERQPRTHYYAVAVVKK GGSFQLNELQGLKSCHTGLRRTAG WNVPIGTLRPFLNWTGPPEPIEAA VARFFSASCVPGADKGQFPNLCRL CAGTGENKCAFSSQEPYFSYSGAF KCLRDGAGDVAFIRESTVFEDLSD EAERDEYELLCPDNTRKPVDKFKD CHLARVPSHAVVARSVNGKEDAIW NLLRQAQEKFGKDKSPKFQLFGSP SGQKDLLFKDSAIGFSRVPPRIDS GLYLGSGYFTAIQN LRK - GenBank: M93150.1) with 6 × His in the C-terminal |
Co-expression with B. mori sericin | Hydrogels for sustained release and gastrointestinal delivery of rhLF in mice Protection against degradation of the rhLF |
[207, 236] |
| Refs | ||||
|---|---|---|---|---|
| N-terminal domain (exon 1) in native resilin (fruit fly Drosophila melanogaster resilin gene CG15920) | (GGRPSDSYGAPGGGN)18 | Carboxyl-terminal (CT) domain of major ampullate spidroin 1 (MaSp1) of the spider N. clavipes (NcCT) | Hydrogels with a pH-dependent release of Rhodamine B | [134] |
| Elastin-like (SELP-415K) | [(GVGVP)4GKGVP(GVGVP)11] | Recombinant B. mori silk | Sustained gastrointestinal delivery of GM-0111 for mucosal inflammatory diseases (murine model) | [71] |
| Refs | |||
|---|---|---|---|
| Glycerol or 2-pyrrolidone | Recombinant spider silk protein eADF4(C16) from A. diadematus fibroin | DDS for small bioactive drugs (paracetamol and tetracaine HCl), complex sugars (dextran), or model proteins (lysozyme and BSA) | [141] |
| Mesoporous silica nanoparticles | Recombinant spider silk protein eADF4(C16) from A. diadematus fibroin | Over 15 days release of antimicrobial peptide | [132] |
| Refs | ||||
|---|---|---|---|---|
| DTNB (5,5′-dithiobis-2-nitrobenzoic acid) | Disulfide bonds | Polyanionic recombinant silk protein from A. diadematus fibroin eADF4(C16), the polycationic variant eADF4(κ16) as well as the cysteine-bearing variants of both proteins | Redox-responsive Release DDS |
[37] |
| DMAB (para-dimethylaminobenzaldehyde) | Carboxyl groups via and hydrazine linker | pH-responsive carrier DDS | ||
| Different fluorophores and the antibiotic Levofloxacin | Click-chemistry | Recombinant spider silk 4RepCT with nonnatural methionine analogue L-Aha (4RepCT3Aha) | Multifunctional derivatization of recombinant silk with click-chemistry pH-labile linker for controlled drug-release |
[143] |
| Thrombin-sensitive peptide (TSP) | GFDPRGFPAGG | MaSp2 protein of N. clavipes | Vancomycin loaded particles Increase in thrombin-like activity during S. aureus infections due to its secretion of staphylocoagulase Improved bacterial clearing in a murine model |
[182] |
| Refs | |||
|---|---|---|---|
| Lysozyme, Bovine serum albumin (BSA), Horseradish peroxidase (HRP) | Recombinant spider silk fibroin eADF4(C16) from A. diadematus fibroin | Direct or diffusion mediated loading of hydrogels and particles | [130] |
| 6-mercaptopurine, FITC | recombinant spider silk eADF4(κ16) and eADF4(Ω16) from A. diadematus fibroin | Organic co-solvent incorporation of hydrophobic drugs in hydrogels and 3D printable systems | [80] |
| Doxorubicin and Rhodamine B | Recombinant eTuSp1 tubuliform spidroin from black widow spider | 1,1,1,3,3,3-Hexafluoro-2-propanol + Oil mixture Proteinase-enhanced release behaviors In vivo studies with Hela cells |
[131] |
| Mitoxantrone and etoposide | MaSp1 and MaSp2 protein of N. clavipes | MaSp2 spheres more dispersed, smaller, with solid core, higher beta-sheet structure content, Superior loading capacity of the MaSp2 variant MaSp1 were a better choice for doxorubicin DDS |
[176] |
| Mitoxantrone and etoposide | MaSp2 protein of N. clavipes | Negatively-charged variant of MaSp2 Increased loading capacity Burst release of mitoxantrone Sustained release of etoposide |
[35] |
Funding
We thank the National Institutes of Health (1R01NS094218-02, P41EB027062) and the Army Research Office (W911NF2120130) for support of the work that in part was included in this review.
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 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.
Declaration of Competing Interests
The authors declare no competing interests.
References
- [1].Homayun B, Lin X, Choi HJ, Challenges and Recent Progress in Oral Drug Delivery Systems for Biopharmaceuticals, Pharmaceutics, 11 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [2].Wu J, Shaidani S, Theodossiou SK, Hartzell EJ, Kaplan DL, Localized, on-demand, sustained drug delivery from biopolymer-based materials, Expert Opinion on Drug Delivery, 19 (2022) 1317–1335. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [3].Adepu S, Ramakrishna S, Controlled Drug Delivery Systems: Current Status and Future Directions, Molecules, 26 (2021) 5905. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [4].Zaman R, Islam RA, Ibnat N, Othman I, Zaini A, Lee CY, Chowdhury EH, Current strategies in extending half-lives of therapeutic proteins, Journal of Controlled Release, 301 (2019) 176–189. [DOI] [PubMed] [Google Scholar]
- [5].Yarger JL, Cherry BR, Vaart v.d.A., Uncovering the structure–function relationship in spider silk, Nature Reviews Materials, 3 (2018) 1–11. [Google Scholar]
- [6].Peng Q, Zhang Y, Lu L, Shao H, Qin K, Hu X, Xia X, Recombinant spider silk from aqueous solutions via a bio-inspired microfluidic chip, Scientific Reports, 6 (2016) 36473. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [7].Aigner TB, DeSimone E, Scheibel T, Biomedical Applications of Recombinant Silk-Based Materials, Advanced Materials, 30 (2018). [DOI] [PubMed] [Google Scholar]
- [8].Gosline JM, DeMont ME, Denny MW, The structure and properties of spider silk, Endeavour, 10 (1986) 37–43. [Google Scholar]
- [9].Gosline JM, Guerette PA, Ortlepp CS, Savage KN, The mechanical design of spider silks: from fibroin sequence to mechanical function, Journal of Experimental Biology, 202 (1999) 3295–3303. [DOI] [PubMed] [Google Scholar]
- [10].Tian M, Lewis RV, Tubuliform silk protein: A protein with unique molecular characteristics and mechanical properties in the spider silk fibroin family, Applied Physics A, 82 (2006) 265–273. [Google Scholar]
- [11].Hayashi CY, Blackledge TA, Lewis RV, Molecular and Mechanical Characterization of Aciniform Silk: Uniformity of Iterated Sequence Modules in a Novel Member of the Spider Silk Fibroin Gene Family, Molecular Biology and Evolution, 21 (2004) 1950–1959. [DOI] [PubMed] [Google Scholar]
- [12].Koh L-D, Cheng Y, Teng C-P, Khin Y-W, Loh X-J, Tee S-Y, Low M, Ye E, Yu H-D, Zhang Y-W, Han M-Y, Structures, mechanical properties and applications of silk fibroin materials, Progress in Polymer Science, 46 (2015) 86–110. [Google Scholar]
- [13].Hui Y, Yi X, Hou F, Wibowo D, Zhang F, Zhao D, Gao H, Zhao C-X, Role of Nanoparticle Mechanical Properties in Cancer Drug Delivery, ACS Nano, 13 (2019) 7410–7424. [DOI] [PubMed] [Google Scholar]
- [14].Simmons AH, Michal CA, Jelinski LW, Molecular orientation and two-component nature of the crystalline fraction of spider dragline silk, Science, 271 (1996) 84–87. [DOI] [PubMed] [Google Scholar]
- [15].Zhou C-Z, Confalonieri F, Jacquet M, Perasso R, Li Z-G, Janin J, Silk fibroin: Structural implications of a remarkable amino acid sequence, Proteins: Structure, Function, and Genetics, 44 (2001) 119–122. [DOI] [PubMed] [Google Scholar]
- [16].Naskar D, Sapru S, Ghosh AK, Reis RL, Dey T, Kundu SC, Nonmulberry silk proteins: multipurpose ingredient in bio-functional assembly, Biomedical Materials, 16 (2021). [DOI] [PubMed] [Google Scholar]
- [17].Weisman S, Haritos VS, Church JS, Huson MG, Mudie ST, Rodgers AJW, Dumsday GJ, Sutherland TD, Honeybee silk: Recombinant protein production, assembly and fiber spinning, Biomaterials, 31 (2010) 2695–2700. [DOI] [PubMed] [Google Scholar]
- [18].Shi J, Lua S, Du N, Liu X, Song J, Identification, recombinant production and structural characterization of four silk proteins from the Asiatic honeybee Apis cerana, Biomaterials, 29 (2008) 2820–2828. [DOI] [PubMed] [Google Scholar]
- [19].Kambe Y, Sutherland TD, Kameda T, Recombinant production and film properties of full-length hornet silk proteins, Acta Biomaterialia, 10 (2014) 3590–3598. [DOI] [PubMed] [Google Scholar]
- [20].Khamhaengpol A, Siri S, Composite Electrospun Scaffold Derived from Recombinant Fibroin of Weaver Ant (Oecophylla smaragdina) as Cell-Substratum, Applied Biochemistry and Biotechnology, 183 (2017) 110–125. [DOI] [PubMed] [Google Scholar]
- [21].Parker KD, Rudall KM, The Silk of the Egg-Stalk of the Green Lace-Wing Fly: Structure of the Silk of Chrysopa Egg-stalks, Nature, 179 (1957) 905–906. [Google Scholar]
- [22].Weisman S, Okada S, Mudie ST, Huson MG, Trueman HE, Sriskantha A, Haritos VS, Sutherland TD, Fifty years later: The sequence, structure and function of lacewing cross-beta silk, Journal of Structural Biology, 168 (2009) 467–475. [DOI] [PubMed] [Google Scholar]
- [23].Wu J, Sahoo JK, Li Y, Xu Q, Kaplan DL, Challenges in delivering therapeutic peptides and proteins: A silk-based solution, Journal of Controlled Release, 345 (2022) 176–189. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [24].Thurber AE, Omenetto FG, Kaplan DL, In vivo bioresponses to silk proteins, Biomaterials, 71 (2015) 145–157. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [25].Gholipourmalekabadi M, Sapru S, Samadikuchaksaraei A, Reis RL, Kaplan DL, Kundu SC, Silk fibroin for skin injury repair: Where do things stand?, Advanced Drug Delivery Reviews, 153 (2020) 28–53. [DOI] [PubMed] [Google Scholar]
- [26].Cao Y, Wang B, Biodegradation of Silk Biomaterials, International Journal of Molecular Sciences, 10 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [27].Park YR, Sultan MT, Park HJ, Lee JM, Ju HW, Lee OJ, Lee DJ, Kaplan DL, Park CH, NF-κB signaling is key in the wound healing processes of silk fibroin, Acta Biomaterialia, 67 (2018) 183–195. [DOI] [PubMed] [Google Scholar]
- [28].Martínez-Mora C, Mrowiec A, García-Vizcaíno EM, Alcaraz A, Cenis JL, Nicolás FJ, Fibroin and Sericin from Bombyx mori Silk Stimulate Cell Migration through Upregulation and Phosphorylation of c-Jun, PLOS ONE, 7 (2012) e42271. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [29].Sahu N, Baligar P, Midha S, Kundu B, Bhattacharjee M, Mukherjee S, Mukherjee S, Maushart F, Das S, Loparic M, Kundu SC, Ghosh S, Mukhopadhyay A, Nonmulberry Silk Fibroin Scaffold Shows Superior Osteoconductivity Than Mulberry Silk Fibroin in Calvarial Bone Regeneration, Advanced Healthcare Materials, 4 (2015) 1709–1721. [DOI] [PubMed] [Google Scholar]
- [30].Xue R, Liu Y, Zhang Q, Liang C, Qin H, Liu P, Wang K, Zhang X, Chen L, Wei Y, Kelly RM, Shape Changes and Interaction Mechanism of Escherichia coli Cells Treated with Sericin and Use of a Sericin-Based Hydrogel for Wound Healing, Applied and Environmental Microbiology, 82 (2016) 4663–4672. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [31].Kumar JP, Mandal BB, Antioxidant potential of mulberry and non-mulberry silk sericin and its implications in biomedicine, Free Radical Biology and Medicine, 108 (2017) 803–818. [DOI] [PubMed] [Google Scholar]
- [32].Aramwit P, Towiwat P, Srichana T, Anti-inflammatory Potential of Silk Sericin, Natural Product Communications, 8 (2013) 1934578X1300800424. [PubMed] [Google Scholar]
- [33].Kumar JP, Mandal BB, Silk sericin induced pro-oxidative stress leads to apoptosis in human cancer cells, Food and Chemical Toxicology, 123 (2019) 275–287. [DOI] [PubMed] [Google Scholar]
- [34].Kumar JP, Alam S, Jain AK, Ansari KM, Mandal BB, Protective Activity of Silk Sericin against UV Radiation-Induced Skin Damage by Downregulating Oxidative Stress, ACS Applied Bio Materials, 1 (2018) 2120–2132. [DOI] [PubMed] [Google Scholar]
- [35].Kucharczyk K, Weiss M, Jastrzebska K, Luczak M, Ptak A, Kozak M, Mackiewicz A, Dams-Kozlowska H, Bioengineering the spider silk sequence to modify its affinity for drugs, International Journal of Nanomedicine, 13 (2018) 4247–4261. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [36].Doblhofer E, Scheibel T, Engineering of Recombinant Spider Silk Proteins Allows Defined Uptake and Release of Substances, Journal of Pharmaceutical Sciences, 104 (2015) 988–994. [DOI] [PubMed] [Google Scholar]
- [37].Herold HM, Döbl A, Wohlrab S, Humenik M, Scheibel T, Designed Spider Silk-Based Drug Carrier for Redox- or pH-Triggered Drug Release, Biomacromolecules, 21 (2020) 4904–4912. [DOI] [PubMed] [Google Scholar]
- [38].Wohlrab S, Müller S, Schmidt A, Neubauer S, Kessler H, Leal-Egaña A, Scheibel T, Cell adhesion and proliferation on RGD-modified recombinant spider silk proteins, Biomaterials, 33 (2012) 6650–6659. [DOI] [PubMed] [Google Scholar]
- [39].Kozlowska AK, Florczak A, Smialek M, Dondajewska E, Mackiewicz A, Kortylewski M, Dams-Kozlowska H, Functionalized bioengineered spider silk spheres improve nuclease resistance and activity of oligonucleotide therapeutics providing a strategy for cancer treatment, Acta Biomaterialia, 59 (2017) 221–233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [40].Kucharczyk K, Florczak A, Deptuch T, Penderecka K, Jastrzebska K, Mackiewicz A, Dams-Kozlowska H, Drug affinity and targeted delivery: double functionalization of silk spheres for controlled doxorubicin delivery into Her2-positive cancer cells, Journal of Nanobiotechnology, 18 (2020) 56. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [41].Florczak A, Mackiewicz A, Dams-Kozlowska H, Functionalized Spider Silk Spheres As Drug Carriers for Targeted Cancer Therapy, Biomacromolecules, 15 (2014) 2971–2981. [DOI] [PubMed] [Google Scholar]
- [42].Isaacson KJ, Jensen MM, Steinhauff DB, Kirklow JE, Mohammadpour R, Grunberger JW, Cappello J, Ghandehari H, Location of stimuli-responsive peptide sequences within silk-elastinlike protein-based polymers affects nanostructure assembly and drug–polymer interactions, Journal of Drug Targeting, 28 (2020) 766–779. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [43].Bini E, Foo CWP, Huang J, Karageorgiou V, Kitchel B, Kaplan DL, RGD-Functionalized Bioengineered Spider Dragline Silk Biomaterial, Biomacromolecules, 7 (2006) 3139–3145. [DOI] [PubMed] [Google Scholar]
- [44].Pereira AM, Machado R, da Costa A, Ribeiro A, Collins T, Gomes AC, Leonor IB, Kaplan DL, Reis RL, Casal M, Silk-based biomaterials functionalized with fibronectin type II promotes cell adhesion, Acta Biomaterialia, 47 (2017) 50–59. [DOI] [PubMed] [Google Scholar]
- [45].Widhe M, Johansson U, Hillerdahl C-O, Hedhammar M, Recombinant spider silk with cell binding motifs for specific adherence of cells, Biomaterials, 34 (2013) 8223–8234. [DOI] [PubMed] [Google Scholar]
- [46].Holland C, Numata K, Rnjak-Kovacina J, Seib FP, The Biomedical Use of Silk: Past, Present, Future, Advanced Healthcare Materials, 8 (2019) 1800465. [DOI] [PubMed] [Google Scholar]
- [47].Yanagisawa S, Zhu Z, Kobayashi I, Uchino K, Tamada Y, Tamura T, Asakura T, Improving Cell-Adhesive Properties of Recombinant Bombyx mori Silk by Incorporation of Collagen or Fibronectin Derived Peptides Produced by Transgenic Silkworms, Biomacromolecules, 8 (2007) 3487–3492. [DOI] [PubMed] [Google Scholar]
- [48].Ma S-Y, Smagghe G, Xia Q-Y, Genome editing in Bombyx mori: New opportunities for silkworm functional genomics and the sericulture industry, Insect Science, 26 (2019) 964–972. [DOI] [PubMed] [Google Scholar]
- [49].Li Z, You L, Zhang Q, Yu Y, Tan A, A Targeted In-Fusion Expression System for Recombinant Protein Production in Bombyx mori, Frontiers in Genetics, 12 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [50].Bogush VG, Sokolova OS, Davydova LI, Klinov DV, Sidoruk KV, Esipova NG, Neretina TV, Orchanskyi IA, Makeev VY, Tumanyan VG, Shaitan KV, Debabov VG, Kirpichnikov MP, A Novel Model System for Design of Biomaterials Based on Recombinant Analogs of Spider Silk Proteins, Journal of Neuroimmune Pharmacology, 4 (2008) 17. [DOI] [PubMed] [Google Scholar]
- [51].Harris TI, Gaztambide DA, Day BA, Brock CL, Ruben AL, Jones JA, Lewis RV, Sticky Situation: An Investigation of Robust Aqueous-Based Recombinant Spider Silk Protein Coatings and Adhesives, Biomacromolecules, 17 (2016) 3761–3772. [DOI] [PubMed] [Google Scholar]
- [52].Scheller J, Henggeler D, Viviani A, Conrad U, Purification of Spider Silk-elastin from Transgenic Plants and Application for Human Chondrocyte Proliferation, Transgenic Research, 13 (2004) 51–57. [DOI] [PubMed] [Google Scholar]
- [53].Wen H, Lan X, Zhang Y, Zhao T, Wang Y, Kajiura Z, Nakagaki M, Transgenic silkworms (Bombyx mori) produce recombinant spider dragline silk in cocoons, Molecular Biology Reports, 37 (2010) 1815–1821. [DOI] [PubMed] [Google Scholar]
- [54].Stark M, Grip S, Rising A, Hedhammar M, Engström W, Hjälm G, Johansson J, Macroscopic Fibers Self-Assembled from Recombinant Miniature Spider Silk Proteins, Biomacromolecules, 8 (2007) 1695–1701. [DOI] [PubMed] [Google Scholar]
- [55].Jastrzebska K, Felcyn E, Kozak M, Szybowicz M, Buchwald T, Pietralik Z, Jesionowski T, Mackiewicz A, Dams-Kozlowska H, The method of purifying bioengineered spider silk determines the silk sphere properties, Scientific Reports, 6 (2016) 28106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [56].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, 43 (2004) 13604–13612. [DOI] [PubMed] [Google Scholar]
- [57].Xia X-X, Qian Z-G, Ki CS, Park YH, Kaplan DL, Lee SY, Native-sized recombinant spider silk protein produced in metabolically engineered Escherichia coli results in a strong fiber, Proceedings of the National Academy of Sciences, 107 (2010) 14059–14063. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [58].Huang W, Rollett A, Kaplan DL, Silk-elastin-like protein biomaterials for the controlled delivery of therapeutics, Expert Opinion on Drug Delivery, 12 (2015) 779–791. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [59].Huang J, Valluzzi R, Bini E, Vernaglia B, Kaplan DL, Cloning, Expression, and Assembly of Sericin-like Protein, Journal of Biological Chemistry, 278 (2003) 46117–46123. [DOI] [PubMed] [Google Scholar]
- [60].Thomas DS, Manoharan C, Rasalkar S, Mishra RK, Gopalapillai R, Expression of Silk Sericin-Cecropin B Fusion Protein in Pichia pastoris and Cell-Free System, Bioscience Biotechnology Research Communications, 13 (2020) 981–984. [Google Scholar]
- [61].Terada S, Nishimura T, Sasaki M, Yamada H, Miki M, Sericin, a protein derived from silkworms, accelerates the proliferation of several mammalian cell lines including a hybridoma, Cytotechnology, 40 (2002) 3–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [62].Tsujimoto K, Takagi H, Takahashi M, Yamada H, Nakamori S, Cryoprotective Effect of the Serine-Rich Repetitive Sequence in Silk Protein Sericin1, The Journal of Biochemistry, 129 (2001) 979–986. [DOI] [PubMed] [Google Scholar]
- [63].Thomas DS, Manoharan C, Rasalkar S, Mishra RK, Gopalapillai R, Recombinant expression of sericin-cecropin fusion protein and its functional activity, Biotechnology Letters, 42 (2020) 1673–1682. [DOI] [PubMed] [Google Scholar]
- [64].Thomas DS, Manoharan C, Rasalkar S, Mishra RK, Gopalapillai R, Enhanced antioxidant properties of sericin-cecropin fusion protein against oxidative stress in human adult dermal fibroblasts, Journal of Bioactive and Compatible Polymers, 36 (2020) 3–12. [Google Scholar]
- [65].Bostan F, Surmeli NB, Cloning, expression, and characterization of a novel sericin-like protein, Biotechnology and Applied Biochemistry. [DOI] [PubMed] [Google Scholar]
- [66].Kundu SC, Kundu B, Talukdar S, Bano S, Nayak S, Kundu J, Mandal BB, Bhardwaj N, Botlagunta M, Dash BC, Acharya C, Ghosh AK, Invited review nonmulberry silk biopolymers, Biopolymers, 97 (2012) 455–467. [DOI] [PubMed] [Google Scholar]
- [67].Asakura T, Suzuki Y, Nagano A, Knight D, Kamiya M, Demura M, Synthesis and Characterization of Water-Soluble Silk Peptides and Recombinant Silk Protein Containing Polyalanine, the Integrin Binding Site, and Two Glutamic Acids at Each Terminal Site as a Possible Candidate for Use in Bone Repair Materials, Biomacromolecules, 14 (2013) 3731–3741. [DOI] [PubMed] [Google Scholar]
- [68].Xia X-X, Xu Q, Hu X, Qin G, Kaplan DL, Tunable Self-Assembly of Genetically Engineered Silk–Elastin-like Protein Polymers, Biomacromolecules, 12 (2011) 3844–3850. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [69].Xia X-X, Wang M, Lin Y, Xu Q, Kaplan DL, Hydrophobic Drug-Triggered Self-Assembly of Nanoparticles from Silk-Elastin-Like Protein Polymers for Drug Delivery, Biomacromolecules, 15 (2014) 908–914. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [70].Parker RN, Wu WA, McKay TB, Xu Q, Kaplan DL, Design of Silk-Elastin-Like Protein Nanoparticle Systems with Mucoadhesive Properties, Journal of Functional Biomaterials, 10 (2019) 49. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [71].Steinhauff D, Jensen M, Talbot M, Jia W, Isaacson K, Jedrzkiewicz J, Cappello J, Oottamasathien S, Ghandehari H, Silk-elastinlike copolymers enhance bioaccumulation of semisynthetic glycosaminoglycan ethers for prevention of radiation induced proctitis, Journal of Controlled Release, 332 (2021) 503–515. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [72].Gomes SC, Leonor IB, Mano JF, Reis RL, Kaplan DL, Antimicrobial functionalized genetically engineered spider silk, Biomaterials, 32 (2011) 4255–4266. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [73].Currie HA, Deschaume O, Naik RR, Perry CC, Kaplan DL, Genetically Engineered Chimeric Silk–Silver Binding Proteins, Advanced Functional Materials, 21 (2011) 2889–2895. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [74].Gomes S, Leonor IB, Mano JF, Reis RL, Kaplan DL, Spider silk-bone sialoprotein fusion proteins for bone tissue engineering, Soft Matter, 7 (2011) 4964–4973. [Google Scholar]
- [75].Rabotyagova OS, Cebe P, Kaplan DL, Self-Assembly of Genetically Engineered Spider Silk Block Copolymers, Biomacromolecules, 10 (2009) 229–236. [DOI] [PubMed] [Google Scholar]
- [76].Slotta U, Tammer M, Kremer F, Koelsch P, Scheibel T, Structural Analysis of Spider Silk Films, Supramolecular Chemistry, 18 (2006) 465–471. [Google Scholar]
- [77].Lammel A, Schwab M, Slotta U, Winter G, Scheibel T, Processing conditions for the formation of spider silk microspheres, ChemSusChem, 1 (2008) 413–416. [DOI] [PubMed] [Google Scholar]
- [78].Lucke M, Mottas I, Herbst T, Hotz C, Römer L, Schierling M, Herold HM, Slotta U, Spinetti T, Scheibel T, Winter G, Bourquin C, Engert J, Engineered hybrid spider silk particles as delivery system for peptide vaccines, Biomaterials, 172 (2018) 105–115. [DOI] [PubMed] [Google Scholar]
- [79].Slotta UK, Rammensee S, Gorb S, Scheibel T, An engineered spider silk protein forms microspheres, Angewandte Chemie - International Edition, 47 (2008) 4592–4594. [DOI] [PubMed] [Google Scholar]
- [80].Neubauer VJ, Trossmann VT, Jacobi S, Döbl A, Scheibel T, Recombinant Spider Silk Gels Derived from Aqueous–Organic Solvents as Depots for Drugs, Angewandte Chemie International Edition, 60 (2021) 11847–11851. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [81].Spieß K, Wohlrab S, Scheibel T, Structural characterization and functionalization of engineered spider silk films, Soft Matter, 6 (2010) 4168–4174. [Google Scholar]
- [82].Hedhammar M, Bramfeldt H, Baris T, Widhe M, Askarieh G, Nordling K, von Aulock S, Johansson J, Sterilized Recombinant Spider Silk Fibers of Low Pyrogenicity, Biomacromolecules, 11 (2010) 953–959. [DOI] [PubMed] [Google Scholar]
- [83].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, 31 (2010) 9575–9585. [DOI] [PubMed] [Google Scholar]
- [84].Nilebäck L, Chouhan D, Jansson R, Widhe M, Mandal BB, Hedhammar M, Silk–Silk Interactions between Silkworm Fibroin and Recombinant Spider Silk Fusion Proteins Enable the Construction of Bioactive Materials, ACS Applied Materials & Interfaces, 9 (2017) 31634–31644. [DOI] [PubMed] [Google Scholar]
- [85].Chouhan D, Thatikonda N, Nilebäck L, Widhe M, Hedhammar M, Mandal BB, Recombinant Spider Silk Functionalized Silkworm Silk Matrices as Potential Bioactive Wound Dressings and Skin Grafts, ACS Applied Materials & Interfaces, 10 (2018) 23560–23572. [DOI] [PubMed] [Google Scholar]
- [86].Widhe M, Shalaly ND, Hedhammar M, A fibronectin mimetic motif improves integrin mediated cell biding to recombinant spider silk matrices, Biomaterials, 74 (2016) 256–266. [DOI] [PubMed] [Google Scholar]
- [87].Jansson R, Thatikonda N, Lindberg D, Rising A, Johansson J, Nygren P-Å, Hedhammar M, Recombinant Spider Silk Genetically Functionalized with Affinity Domains, Biomacromolecules, 15 (2014) 1696–1706. [DOI] [PubMed] [Google Scholar]
- [88].Chouhan D, Lohe T.-u., Thatikonda N, Naidu VGM, Hedhammar M, Mandal BB, Silkworm Silk Scaffolds Functionalized with Recombinant Spider Silk Containing a Fibronectin Motif Promotes Healing of Full-Thickness Burn Wounds, ACS Biomaterials Science & Engineering, 5 (2019) 4634–4645. [DOI] [PubMed] [Google Scholar]
- [89].Chen J, Hu J, Zuo P, Su X, Liu Z, Yang M, Tailor-made spider-eggcase-silk spheres for efficient lysosomal drug delivery, RSC Adv, 8 (2018) 9394–9401. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [90].Bauer F, Scheibel T, Artificial Egg Stalks Made of a Recombinantly Produced Lacewing Silk Protein, Angewandte Chemie - International Edition, 51 (2012) 6521–6524. [DOI] [PubMed] [Google Scholar]
- [91].Yavuz B, Chambre L, Kaplan DL, Extended release formulations using silk proteins for controlled delivery of therapeutics, Expert Opinion on Drug Delivery, 16 (2019) 741–756. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [92].Chambre L, Martín-Moldes Z, Parker RN, Kaplan DL, Bioengineered elastin- and silk-biomaterials for drug and gene delivery, Advanced Drug Delivery Reviews, 160 (2020) 186–198. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [93].Wenk E, Merkle HP, Meinel L, Silk fibroin as a vehicle for drug delivery applications, Journal of Controlled Release, 150 (2011) 128–141. [DOI] [PubMed] [Google Scholar]
- [94].Wongpinyochit T, Vassileiou AD, Gupta S, Mushrif SH, Johnston BF, Seib FP, Unraveling the Impact of High-Order Silk Structures on Molecular Drug Binding and Release Behaviors, The Journal of Physical Chemistry Letters, 10 (2019) 4278–4284. [DOI] [PubMed] [Google Scholar]
- [95].Pham DT, Tiyaboonchai W, Fibroin nanoparticles: a promising drug delivery system, Drug Deliv, 27 (2020) 431–448. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [96].Lammel AS, Hu X, Park S-H, Kaplan DL, Scheibel TR, Controlling silk fibroin particle features for drug delivery, Biomaterials, 31 (2010) 4583–4591. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [97].Marin MA, Mallepally RR, McHugh MA, Silk fibroin aerogels for drug delivery applications, The Journal of Supercritical Fluids, 91 (2014) 84–89. [Google Scholar]
- [98].Nguyen TP, Nguyen QV, Nguyen V-H, Le T-H, Huynh VQN, Vo D-VN, Trinh QT, Kim SY, Le QV, Silk Fibroin-Based Biomaterials for Biomedical Applications: A Review, Polymers (Basel), 11 (2019) 1933. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [99].Zhao Z-J, Wang Q, Zhang L, Liu Y-C, A Different Diffusion Mechanism for Drug Molecules in Amorphous Polymers, The Journal of Physical Chemistry B, 111 (2007) 4411–4416. [DOI] [PubMed] [Google Scholar]
- [100].Karve KA, Gil ES, McCarthy SP, Kaplan DL, Effect of β-sheet crystalline content on mass transfer in silk films, Journal of Membrane Science, 383 (2011) 44–49. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [101].Lu Q, Wang X, Hu X, Cebe P, Omenetto F, Kaplan DL, Stabilization and Release of Enzymes from Silk Films, Macromolecular Bioscience, 10 (2010) 359–368. [DOI] [PubMed] [Google Scholar]
- [102].Belda Marín C, Fitzpatrick V, Kaplan DL, Landoulsi J, Guénin E, Egles C, Silk Polymers and Nanoparticles: A Powerful Combination for the Design of Versatile Biomaterials, Frontiers in Chemistry, 8 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [103].Seib FP, Kaplan DL, Doxorubicin-loaded silk films: Drug-silk interactions and in vivo performance in human orthotopic breast cancer, Biomaterials, 33 (2012) 8442–8450. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [104].Chambre L, Parker RN, Allardyce BJ, Valente F, Rajkhowa R, Dilley RJ, Wang X, Kaplan DL, Tunable Biodegradable Silk-Based Memory Foams with Controlled Release of Antibiotics, ACS Applied Bio Materials, 3 (2020) 2466–2472. [DOI] [PubMed] [Google Scholar]
- [105].Ding Z, Zhou M, Zhou Z, Zhang W, Jiang X, Lu X, Zuo B, Lu Q, Kaplan DL, Injectable Silk Nanofiber Hydrogels for Sustained Release of Small-Molecule Drugs and Vascularization, ACS Biomaterials Science & Engineering, 5 (2019) 4077–4088. [DOI] [PubMed] [Google Scholar]
- [106].Zhu C, Ding Z, Lu Q, Lu G, Xiao L, Zhang X, Dong X, Ru C, Kaplan DL, Injectable Silk–Vaterite Composite Hydrogels with Tunable Sustained Drug Release Capacity, ACS Biomaterials Science & Engineering, 5 (2019) 6602–6609. [DOI] [PubMed] [Google Scholar]
- [107].Wu H, Liu S, Xiao L, Dong X, Lu Q, Kaplan DL, Injectable and pH-Responsive Silk Nanofiber Hydrogels for Sustained Anticancer Drug Delivery, ACS Applied Materials & Interfaces, 8 (2016) 17118–17126. [DOI] [PubMed] [Google Scholar]
- [108].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, 29 (2008) 894–903. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [109].Xu L, Weatherbee-Martin N, Liu X-Q, Rainey JK, Recombinant Silk Fiber Properties Correlate to Prefibrillar Self-Assembly, Small, 15 (2019) 1805294. [DOI] [PubMed] [Google Scholar]
- [110].Ramezaniaghdam M, Nahdi ND, Reski R, Recombinant Spider Silk: Promises and Bottlenecks, Frontiers in Bioengineering and Biotechnology, 10 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [111].Jaleel Z, Zhou S, Martín-Moldes Z, Baugh LM, Yeh J, Dinjaski N, Brown LT, Garb JE, Kaplan DL, Expanding Canonical Spider Silk Properties through a DNA Combinatorial Approach, Materials (Basel), 13 (2020) 3596. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [112].Geckil H, Xu F, Zhang X, Moon S, Demirci U, Engineering hydrogels as extracellular matrix mimics, Nanomedicine, 5 (2010) 469–484. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [113].Wang W, Liu Y, Wang S, Fu X, Zhao T, Chen X, Shao Z, Physically Cross-Linked Silk Fibroin-Based Tough Hydrogel Electrolyte with Exceptional Water Retention and Freezing Tolerance, ACS Applied Materials & Interfaces, 12 (2020) 25353–25362. [DOI] [PubMed] [Google Scholar]
- [114].Sun X, He S, Yao M, Wu X, Zhang H, Yao F, Li J, Fully-physically crosslinked silk fibroin/poly(hydroxyethyl acrylamide) hydrogel with high transparency and adhesive properties for wireless sensing and low-temperature strain sensing, Journal of Materials Chemistry C, 9 (2021) 1880–1887. [Google Scholar]
- [115].Luo K, Yang Y, Shao Z, Physically Crosslinked Biocompatible Silk-Fibroin-Based Hydrogels with High Mechanical Performance, Advanced Functional Materials, 26 (2016) 872–880. [Google Scholar]
- [116].Choi J, McGill M, Raia NR, Hasturk O, Kaplan DL, Silk Hydrogels Crosslinked by the Fenton Reaction, Advanced Healthcare Materials, 8 (2019) 1900644. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [117].Mu X, Sahoo JK, Cebe P, Kaplan DL, Photo-Crosslinked Silk Fibroin for 3D Printing, Polymers (Basel), 12 (2020) 2936. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [118].Kim MH, Park WH, Chemically cross-linked silk fibroin hydrogel with enhanced elastic properties, biodegradability, and biocompatibility, International Journal of Nanomedicine, 11 (2016) 2967–2978. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [119].Kim CS, Yang YJ, Bahn SY, Cha HJ, A bioinspired dual-crosslinked tough silk protein hydrogel as a protective biocatalytic matrix for carbon sequestration, NPG Asia Materials, 9 (2017) e391–e391. [Google Scholar]
- [120].Canchi DR, García AE, Cosolvent Effects on Protein Stability, Annual Review of Physical Chemistry, 64 (2013) 273–293. [DOI] [PubMed] [Google Scholar]
- [121].Eisoldt L, Smith A, Scheibel T, Decoding the secrets of spider silk, Materials Today, 14 (2011) 80–86. [Google Scholar]
- [122].Wang X, Kluge J, Leisk GG, Kaplan DL, Sonication-Induced Gelation of Silk Fibroin for Cell Encapsulation, Biomaterials, 29 (2008) 1054–1064. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [123].Toprakcioglu Z, Knowles TPJ, Shear-mediated sol-gel transition of regenerated silk allows the formation of Janus-like microgels, Scientific Reports, 11 (2021) 6673. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [124].Samal SK, Kaplan DL, Chiellini E, Ultrasound Sonication Effects on Silk Fibroin Protein, Macromolecular Materials and Engineering, 298 (2013) 1201–1208. [Google Scholar]
- [125].Zhao Y, Zhu ZS, Guan J, Wu SJ, Processing, mechanical properties and bio-applications of silk fibroin-based high-strength hydrogels, Acta Biomaterialia, 125 (2021) 57–71. [DOI] [PubMed] [Google Scholar]
- [126].Hasturk O, Jordan KE, Choi J, Kaplan DL, Enzymatically crosslinked silk and silk-gelatin hydrogels with tunable gelation kinetics, mechanical properties and bioactivity for cell culture and encapsulation, Biomaterials, 232 (2020) 119720. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [127].Su D, Yao M, Liu J, Zhong Y, Chen X, Shao Z, Enhancing Mechanical Properties of Silk Fibroin Hydrogel through Restricting the Growth of β-Sheet Domains, ACS Applied Materials & Interfaces, 9 (2017) 17489–17498. [DOI] [PubMed] [Google Scholar]
- [128].Partlow BP, Hanna CW, Rnjak-Kovacina J, Moreau JE, Applegate MB, Burke KA, Marelli B, Mitropoulos AN, Omenetto FG, Kaplan DL, Highly Tunable Elastomeric Silk Biomaterials, Advanced Functional Materials, 24 (2014) 4615–4624. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [129].Vargason AM, Anselmo AC, Mitragotri S, The evolution of commercial drug delivery technologies, Nature Biomedical Engineering, 5 (2021) 951–967. [DOI] [PubMed] [Google Scholar]
- [130].Kumari S, Bargel H, Anby MU, Lafargue D, Scheibel T, Recombinant Spider Silk Hydrogels for Sustained Release of Biologicals, ACS Biomaterials Science & Engineering, 4 (2018) 1750–1759. [DOI] [PubMed] [Google Scholar]
- [131].Chen J, Hu J, Zuo P, Shi J, Yang M, Facile preparation of recombinant spider eggcase silk spheres via an HFIP-on-Oil approach, International Journal of Biological Macromolecules, 116 (2018) 1146–1152. [DOI] [PubMed] [Google Scholar]
- [132].Kumari S, Bargel H, Scheibel T, Recombinant Spider Silk–Silica Hybrid Scaffolds with Drug-Releasing Properties for Tissue Engineering Applications, Macromolecular Rapid Communications, 41 (2020) 1900426. [DOI] [PubMed] [Google Scholar]
- [133].Kumari S, Lang G, DeSimone E, Spengler C, Trossmann VT, Lücker S, Hudel M, Jacobs K, Krämer N, Scheibel T, Engineered spider silk-based 2D and 3D materials prevent microbial infestation, Materials Today, 41 (2020) 21–33. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [134].Luo F, Qian Z-G, Xia X-X, Responsive Protein Hydrogels Assembled from Spider Silk Carboxyl-Terminal Domain and Resilin Copolymers, Polymers (Basel), 10 (2018) 915. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [135].Zhou L, Chen M, Tian L, Guan Y, Zhang Y, Release of Polyphenolic Drugs from Dynamically Bonded Layer-by-Layer Films, ACS Applied Materials & Interfaces, 5 (2013) 3541–3548. [DOI] [PubMed] [Google Scholar]
- [136].Pritchard EM, Szybala C, Boison D, Kaplan DL, Silk fibroin encapsulated powder reservoirs for sustained release of adenosine, Journal of Controlled Release, 144 (2010) 159–167. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [137].Hines DJ, Kaplan DL, Mechanisms of Controlled Release from Silk Fibroin Films, Biomacromolecules, 12 (2011) 804–812. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [138].Yucel T, Lovett ML, Kaplan DL, Silk-based biomaterials for sustained drug delivery, Journal of Controlled Release, 190 (2014) 381–397. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [139].Lu S, Wang X, Lu Q, Zhang X, Kluge JA, Uppal N, Omenetto F, Kaplan DL, Insoluble and Flexible Silk Films Containing Glycerol, Biomacromolecules, 11 (2010) 143–150. [DOI] [PubMed] [Google Scholar]
- [140].Li X, Zhang H, He L, Chen Z, Tan Z, You R, Wang D, Flexible nanofibers-reinforced silk fibroin films plasticized by glycerol, Composites Part B: Engineering, 152 (2018) 305–310. [Google Scholar]
- [141].Agostini E, Winter G, Engert J, Water-based preparation of spider silk films as drug delivery matrices, Journal of Controlled Release, 213 (2015) 134–141. [DOI] [PubMed] [Google Scholar]
- [142].Gustafsson L, Tasiopoulos CP, Jansson R, Kvick M, Duursma T, Gasser TC, van der Wijngaart W, Hedhammar M, Recombinant Spider Silk Forms Tough and Elastic Nanomembranes that are Protein-Permeable and Support Cell Attachment and Growth, Advanced Functional Materials, 30 (2020) 2002982. [Google Scholar]
- [143].Harvey D, Bardelang P, Goodacre SL, Cockayne A, Thomas NR, Antibiotic Spider Silk: Site-Specific Functionalization of Recombinant Spider Silk Using “Click” Chemistry, Advanced Materials, 29 (2017) 1604245. [DOI] [PubMed] [Google Scholar]
- [144].Neubauer VJ, Scheibel T, Spider Silk Fusion Proteins for Controlled Collagen Binding and Biomineralization, ACS Biomaterials Science & Engineering, 6 (2020) 5599–5608. [DOI] [PubMed] [Google Scholar]
- [145].Dinjaski N, Plowright R, Zhou S, Belton DJ, Perry CC, Kaplan DL, Osteoinductive recombinant silk fusion proteins for bone regeneration, Acta Biomaterialia, 49 (2017) 127–139. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [146].Liu J, Chen H, Wang Y, Li G, Zheng Z, Kaplan DL, Wang X, Wang X, Flexible Water-Absorbing Silk-Fibroin Biomaterial Sponges with Unique Pore Structure for Tissue Engineering, ACS Biomaterials Science & Engineering, 6 (2020) 1641–1649. [DOI] [PubMed] [Google Scholar]
- [147].Tamada Y, New Process to Form a Silk Fibroin Porous 3-D Structure, Biomacromolecules, 6 (2005) 3100–3106. [DOI] [PubMed] [Google Scholar]
- [148].Huson MG, Church JS, Poole JM, Weisman S, Sriskantha A, Warden AC, Campbell PM, Ramshaw JAM, Sutherland TD, Controlling the Molecular Structure and Physical Properties of Artificial Honeybee Silk by Heating or by Immersion in Solvents, PLOS ONE, 7 (2012) e52308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [149].Brown JE, Moreau JE, Berman AM, McSherry HJ, Coburn JM, Schmidt DF, Kaplan DL, Shape Memory Silk Protein Sponges for Minimally Invasive Tissue Regeneration, Advanced healthcare materials, 6 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [150].Hu X, Shmelev K, Sun L, Gil E-S, Park S-H, Cebe P, Kaplan DL, Regulation of Silk Material Structure by Temperature-Controlled Water Vapor Annealing, Biomacromolecules, 12 (2011) 1686–1696. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [151].Wen D-L, Sun D-H, Huang P, Huang W, Su M, Wang Y, Han M-D, Kim B, Brugger J, Zhang H-X, Zhang X-S, Recent progress in silk fibroin-based flexible electronics, Microsystems & Nanoengineering, 7 (2021) 35. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [152].Petrou G, Jansson R, Högqvist M, Erlandsson J, Wågberg L, Hedhammar M, Crouzier T, Genetically Engineered Mucoadhesive Spider Silk, Biomacromolecules, 19 (2018) 3268–3279. [DOI] [PubMed] [Google Scholar]
- [153].Florczak A, Deptuch T, Kucharczyk K, Dams-Kozlowska H, Systemic and Local Silk-Based Drug Delivery Systems for Cancer Therapy, Cancers, 13 (2021) 5389. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [154].Albanese A, Tang PS, Chan WCW, The Effect of Nanoparticle Size, Shape, and Surface Chemistry on Biological Systems, Annual Review of Biomedical Engineering, 14 (2012) 1–16. [DOI] [PubMed] [Google Scholar]
- [155].Soares S, Sousa J, Pais A, Vitorino C, Nanomedicine: Principles, Properties, and Regulatory Issues, Frontiers in Chemistry, 6 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [156].Schubert J, Chanana M, Coating Matters: Review on Colloidal Stability of Nanoparticles with Biocompatible Coatings in Biological Media, Living Cells and Organisms, Current Medicinal Chemistry, 25 4553–4586. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [157].Maeda H, Nakamura H, Fang J, The EPR effect for macromolecular drug delivery to solid tumors: Improvement of tumor uptake, lowering of systemic toxicity, and distinct tumor imaging in vivo, Advanced Drug Delivery Reviews, 65 (2013) 71–79. [DOI] [PubMed] [Google Scholar]
- [158].Yoo J, Park C, Yi G, Lee D, Koo H, Active Targeting Strategies Using Biological Ligands for Nanoparticle Drug Delivery Systems, Cancers, 11 (2019) 640. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [159].Wilhelm S, Tavares AJ, Dai Q, Ohta S, Audet J, Dvorak HF, Chan WCW, Analysis of nanoparticle delivery to tumours, Nature Reviews Materials, 1 (2016) 1–12. [Google Scholar]
- [160].Veber DF, Johnson SR, Cheng H-Y, Smith BR, Ward KW, Kopple KD, Molecular Properties That Influence the Oral Bioavailability of Drug Candidates, J. Med. Chem, 45 (2002) 2615–2623. [DOI] [PubMed] [Google Scholar]
- [161].Leeson PD, Springthorpe B, The influence of drug-like concepts on decision-making in medicinal chemistry, Nat Rev Drug Discov, 6 (2007) 881–890. [DOI] [PubMed] [Google Scholar]
- [162].Florczak A, Grzechowiak I, Deptuch T, Kucharczyk K, Kaminska A, Dams-Kozlowska H, Silk Particles as Carriers of Therapeutic Molecules for Cancer Treatment, Materials, 13 (2020) 4946. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [163].Ng CY, Chen C-B, Wu M-Y, Wu J, Yang C-H, Hui RC-Y, Chang Y-C, Lu C-W, Anticancer Drugs Induced Severe Adverse Cutaneous Drug Reactions: An Updated Review on the Risks Associated with Anticancer Targeted Therapy or Immunotherapies, Journal of Immunology Research, 2018 (2018) e5376476. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [164].Herrmann J, Adverse cardiac effects of cancer therapies: cardiotoxicity and arrhythmia, Nat Rev Cardiol, 17 (2020) 474–502. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [165].Tacar O, Sriamornsak P, Dass CR, Doxorubicin: an update on anticancer molecular action, toxicity and novel drug delivery systems, Journal of Pharmacy and Pharmacology, 65 (2013) 157–170. [DOI] [PubMed] [Google Scholar]
- [166].Florczak A, Jastrzebska K, Mackiewicz A, Dams-Kozlowska H, Blending two bioengineered spider silks to develop cancer targeting spheres, Journal of Materials Chemistry B, 5 (2017) 3000–3011. [DOI] [PubMed] [Google Scholar]
- [167].Florczak A, Deptuch T, Lewandowska A, Penderecka K, Kramer E, Marszalek A, Mackiewicz A, Dams-Kozlowska H, Functionalized silk spheres selectively and effectively deliver a cytotoxic drug to targeted cancer cells in vivo, Journal of Nanobiotechnology, 18 (2020) 177. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [168].Deptuch T, Florczak A, Lewandowska A, Leporowska E, Penderecka K, Marszalek A, Mackiewicz A, Dams-Kozlowska H, MS1-type bioengineered spider silk nanoparticles do not exhibit toxicity in an in vivo mouse model, Nanomedicine, (2021). [DOI] [PubMed] [Google Scholar]
- [169].Lammel A, Schwab M, Hofer M, Winter G, Scheibel T, Recombinant spider silk particles as drug delivery vehicles, Biomaterials, 32 (2011) 2233–2240. [DOI] [PubMed] [Google Scholar]
- [170].Andrew MacKay J, Chen M, McDaniel JR, Liu W, Simnick AJ, Chilkoti A, Self-assembling chimeric polypeptide–doxorubicin conjugate nanoparticles that abolish tumours after a single injection, Nature Materials, 8 (2009) 993–999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [171].Zhao P, Dong S, Bhattacharyya J, Chen M, iTEP Nanoparticle-Delivered Salinomycin Displays an Enhanced Toxicity to Cancer Stem Cells in Orthotopic Breast Tumors, Mol. Pharmaceutics, 11 (2014) 2703–2712. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [172].Nultsch K, Germershaus O, Matrix metalloprotease triggered bioresponsive drug delivery systems – Design, synthesis and application, European Journal of Pharmaceutics and Biopharmaceutics, 131 (2018) 189–202. [DOI] [PubMed] [Google Scholar]
- [173].Montecucco A, Zanetta F, Biamonti G, Molecular mechanisms of etoposide, EXCLI J, 14 (2015) 95–108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [174].Fox EJ, Mechanism of action of mitoxantrone, Neurology, 63 (2004) S15–S18. [DOI] [PubMed] [Google Scholar]
- [175].Fernández-Ramos AA, Marchetti-Laurent C, Poindessous V, Antonio S, Laurent-Puig P, Bortoli S, Loriot M-A, Pallet N, 6-mercaptopurine promotes energetic failure in proliferating T cells, Oncotarget, 8 (2017) 43048–43060. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [176].Jastrzebska K, Florczak A, Kucharczyk K, Lin Y, Wang Q, Mackiewicz A, Kaplan DL, Dams-Kozlowska H, Delivery of chemotherapeutics using spheres made of bioengineered spider silks derived from MaSp1 and MaSp2 proteins, Nanomedicine, 13 (2018) 439–454. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [177].Santos RS, Figueiredo C, Azevedo NF, Braeckmans K, De Smedt SC, Nanomaterials and molecular transporters to overcome the bacterial envelope barrier: Towards advanced delivery of antibiotics, Advanced Drug Delivery Reviews, 136–137 (2018) 28–48. [DOI] [PubMed] [Google Scholar]
- [178].Van Giau V, An SSA, Hulme J, Recent advances in the treatment of pathogenic infections using antibiotics and nano-drug delivery vehicles, Drug Des Devel Ther, 13 (2019) 327–343. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [179].Yeh Y-C, Huang T-H, Yang S-C, Chen C-C, Fang J-Y, Nano-Based Drug Delivery or Targeting to Eradicate Bacteria for Infection Mitigation: A Review of Recent Advances, Frontiers in Chemistry, 8 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [180].Pang X, Xiao Q, Cheng Y, Ren E, Lian L, Zhang Y, Gao H, Wang X, Leung W, Chen X, Liu G, Xu C, Bacteria-Responsive Nanoliposomes as Smart Sonotheranostics for Multidrug Resistant Bacterial Infections, ACS Nano, 13 (2019) 2427–2438. [DOI] [PubMed] [Google Scholar]
- [181].Qing G, Zhao X, Gong N, Chen J, Li X, Gan Y, Wang Y, Zhang Z, Zhang Y, Guo W, Luo Y, Liang X-J, Thermo-responsive triple-function nanotransporter for efficient chemo-photothermal therapy of multidrug-resistant bacterial infection, Nature Communications, 10 (2019) 4336. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [182].Mulinti P, Shreffler J, Hasan R, Dea M, Brooks AE, Infection Responsive Smart Delivery of Antibiotics Using Recombinant Spider Silk Nanospheres, Pharmaceutics, 13 (2021) 1358. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [183].Liu C, Bayer A, Cosgrove SE, Daum RS, Fridkin SK, Gorwitz RJ, Kaplan SL, Karchmer AW, Levine DP, Murray BE, Rybak MJ, Talan DA, Chambers HF, Clinical Practice Guidelines by the Infectious Diseases Society of America for the Treatment of Methicillin-Resistant Staphylococcus aureus Infections in Adults and Children: Executive Summary, Clinical Infectious Diseases, 52 (2011) 285–292. [DOI] [PubMed] [Google Scholar]
- [184].Panizzi P, Friedrich R, Fuentes-Prior P, Richter K, Bock PE, Bode W, Fibrinogen Substrate Recognition by Staphylocoagulase·(Pro)thrombin Complexes*, Journal of Biological Chemistry, 281 (2006) 1179–1187. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [185].Vanassche T, Kauskot A, Verhaegen J, Peetermans WE, Ryn v.J., Schneewind O, Hoylaerts MF, Verhamme P, Fibrin formation by staphylothrombin facilitates Staphylococcus aureus-induced platelet aggregation, Thromb Haemost, 107 (2012) 1107–1121. [DOI] [PubMed] [Google Scholar]
- [186].Dandapani S, Marcaurelle LA, Grand Challenge Commentary: Accessing new chemical space for ‘undruggable’ targets, Nat Chem Biol, 6 (2010) 861–863. [DOI] [PubMed] [Google Scholar]
- [187].Dang CV, Reddy EP, Shokat KM, Soucek L, Drugging the ‘undruggable’ cancer targets, Nat Rev Cancer, 17 (2017) 502–508. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [188].Lou B, De Koker S, Lau CYJ, Hennink WE, Mastrobattista E, mRNA Polyplexes with Post-Conjugated GALA Peptides Efficiently Target, Transfect, and Activate Antigen Presenting Cells, Bioconjugate Chem, 30 (2019) 461–475. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [189].Luther DC, Huang R, Jeon T, Zhang X, Lee Y-W, Nagaraj H, Rotello VM, Delivery of drugs, proteins, and nucleic acids using inorganic nanoparticles, Advanced Drug Delivery Reviews, 156 (2020) 188–213. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [190].Vaidya AM, Sun Z, Ayat N, Schilb A, Liu X, Jiang H, Sun D, Scheidt J, Qian V, He S, Gilmore H, Schiemann WP, Lu Z-R, Systemic Delivery of Tumor-Targeting siRNA Nanoparticles against an Oncogenic LncRNA Facilitates Effective Triple-Negative Breast Cancer Therapy, Bioconjugate Chem, 30 (2019) 907–919. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [191].Catalanotto C, Cogoni C, Zardo G, MicroRNA in Control of Gene Expression: An Overview of Nuclear Functions, International Journal of Molecular Sciences, 17 (2016) 1712. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [192].Akinc A, Maier MA, Manoharan M, Fitzgerald K, Jayaraman M, Barros S, Ansell S, Du X, Hope MJ, Madden TD, Mui BL, Semple SC, Tam YK, Ciufolini M, Witzigmann D, Kulkarni JA, van der Meel R, Cullis PR, The Onpattro story and the clinical translation of nanomedicines containing nucleic acid-based drugs, Nat. Nanotechnol, 14 (2019) 1084–1087. [DOI] [PubMed] [Google Scholar]
- [193].Alshaer W, Hillaireau H, Fattal E, Aptamer-guided nanomedicines for anticancer drug delivery, Advanced Drug Delivery Reviews, 134 (2018) 122–137. [DOI] [PubMed] [Google Scholar]
- [194].Li L, Xu S, Yan H, Li X, Yazd HS, Li X, Huang T, Cui C, Jiang J, Tan W, Nucleic Acid Aptamers for Molecular Diagnostics and Therapeutics: Advances and Perspectives, Angewandte Chemie International Edition, 60 (2021) 2221–2231. [DOI] [PubMed] [Google Scholar]
- [195].Leader B, Baca QJ, Golan DE, Protein therapeutics: a summary and pharmacological classification, Nat Rev Drug Discov, 7 (2008) 21–39. [DOI] [PubMed] [Google Scholar]
- [196].Geraldes DC, Beraldo-de-Araújo VL, Pardo BOP, Pessoa Junior A, Stephano MA, de Oliveira-Nascimento L, Protein drug delivery: current dosage form profile and formulation strategies, Journal of Drug Targeting, 28 (2020) 339–355. [DOI] [PubMed] [Google Scholar]
- [197].Muttenthaler M, King GF, Adams DJ, Alewood PF, Trends in peptide drug discovery, Nat Rev Drug Discov, 20 (2021) 309–325. [DOI] [PubMed] [Google Scholar]
- [198].Vaughan HJ, Green JJ, Tzeng SY, Cancer-Targeting Nanoparticles for Combinatorial Nucleic Acid Delivery, Advanced Materials, 32 (2020) 1901081. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [199].Numata K, Subramanian B, Currie HA, Kaplan DL, Bioengineered silk protein-based gene delivery systems, Biomaterials, 30 (2009) 5775–5784. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [200].Numata K, Hamasaki J, Subramanian B, Kaplan DL, Gene delivery mediated by recombinant silk proteins containing cationic and cell binding motifs, Journal of Controlled Release, 146 (2010) 136–143. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [201].Numata K, Kaplan DL, Silk-Based Gene Carriers with Cell Membrane Destabilizing Peptides, Biomacromolecules, 11 (2010) 3189–3195. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [202].Numata K, Reagan MR, Goldstein RH, Rosenblatt M, Kaplan DL, Spider Silk-Based Gene Carriers for Tumor Cell-Specific Delivery, Bioconjugate Chem, 22 (2011) 1605–1610. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [203].Tokareva OS, Glettig DL, Abbott RD, Kaplan DL, Multifunctional spider silk polymers for gene delivery to human mesenchymal stem cells, Journal of Biomedical Materials Research Part B: Applied Biomaterials, 103 (2015) 1390–1401. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [204].Blüm C, Nichtl A, Scheibel T, Spider Silk Capsules as Protective Reaction Containers for Enzymes, Advanced Functional Materials, 24 (2014) 763–768. [Google Scholar]
- [205].Xu S, Yang Q, Wang R, Tian C, Ji Y, Tan H, Zhao P, Kaplan DL, Wang F, Xia Q, Genetically engineered pH-responsive silk sericin nanospheres with efficient therapeutic effect on ulcerative colitis, Acta Biomaterialia, (2022). [DOI] [PubMed] [Google Scholar]
- [206].Cutone A, Ianiro G, Lepanto MS, Rosa L, Valenti P, Bonaccorsi di Patti MC, Musci G, Lactoferrin in the Prevention and Treatment of Intestinal Inflammatory Pathologies Associated with Colorectal Cancer Development, Cancers, 12 (2020) E3806. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [207].Xu S, Wang F, Wang Y, Wang R, Hou K, Tian C, Ji Y, Yang Q, Zhao P, Xia Q, A silkworm based silk gland bioreactor for high-efficiency production of recombinant human lactoferrin with antibacterial and anti-inflammatory activities, Journal of Biological Engineering, 13 (2019) 61. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [208].Faraday M, The Bakerian Lecture: Experimental relations of gold to light, Philos. Trans. R. Soc. Lond, 147 (1857) 145–181. [Google Scholar]
- [209].Huang H, Feng W, Chen Y, Shi J, Inorganic nanoparticles in clinical trials and translations, Nano Today, 35 (2020) 100972. [Google Scholar]
- [210].Friedrich RP, Cicha I, Alexiou C, Iron Oxide Nanoparticles in Regenerative Medicine and Tissue Engineering, Nanomaterials, 11 (2021) 2337. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [211].Li W, Cao Z, Liu R, Liu L, Li H, Li X, Chen Y, Lu C, Liu Y, AuNPs as an important inorganic nanoparticle applied in drug carrier systems, Artificial Cells, Nanomedicine, and Biotechnology, 47 (2019) 4222–4233. [DOI] [PubMed] [Google Scholar]
- [212].Billings C, Langley M, Warrington G, Mashali F, Johnson JA, Magnetic Particle Imaging: Current and Future Applications, Magnetic Nanoparticle Synthesis Methods and Safety Measures, International Journal of Molecular Sciences, 22 (2021) 7651. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [213].Avasthi A, Caro C, Pozo-Torres E, Leal MP, García-Martín ML, Magnetic Nanoparticles as MRI Contrast Agents, in: Puente-Santiago AR, Rodríguez-Padrón D, Puente-Santiago AR, Rodríguez-Padrón D (Eds.) Surface-modified Nanobiomaterials for Electrochemical and Biomedicine Applications, Cham, 2020, pp. 49–91. [Google Scholar]
- [214].Vallabani NVS, Singh S, Recent advances and future prospects of iron oxide nanoparticles in biomedicine and diagnostics, 3 Biotech, 8 (2018) 279. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [215].Ganguly S, Margel S, Review: Remotely controlled magneto-regulation of therapeutics from magnetoelastic gel matrices, Biotechnology Advances, 44 (2020) 107611. [DOI] [PubMed] [Google Scholar]
- [216].Vangijzegem T, Stanicki D, Laurent S, Magnetic iron oxide nanoparticles for drug delivery: applications and characteristics, Expert Opinion on Drug Delivery, 16 (2019) 69–78. [DOI] [PubMed] [Google Scholar]
- [217].Kucharczyk K, Rybka JD, Hilgendorff M, Krupinski M, Slachcinski M, Mackiewicz A, Giersig M, Dams-Kozlowska H, Composite spheres made of bioengineered spider silk and iron oxide nanoparticles for theranostics applications, PLOS ONE, 14 (2019) e0219790. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [218].Kucharczyk K, Kaczmarek K, Jozefczak A, Slachcinski M, Mackiewicz A, Dams-Kozlowska H, Hyperthermia treatment of cancer cells by the application of targeted silk/iron oxide composite spheres, Materials Science and Engineering: C, 120 (2021) 111654. [DOI] [PubMed] [Google Scholar]
- [219].Huang T, Kumari S, Herold H, Bargel H, Aigner TB, Heath DE, O’Brien-Simpson NM, O’Connor AJ, Scheibel T, Enhanced Antibacterial Activity of Se Nanoparticles Upon Coating with Recombinant Spider Silk Protein eADF4(κ16), International Journal of Nanomedicine, 15 (2020) 4275—4288. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [220].Lin Y, Xia X, Wang M, Wang Q, An B, Tao H, Xu Q, Omenetto F, Kaplan DL, Genetically Programmable Thermoresponsive Plasmonic Gold/Silk-Elastin Protein Core/Shell Nanoparticles, Langmuir, 30 (2014) 4406–4414. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [221].Cheng Y, Zhang S, Kang N, Huang J, Lv X, Wen K, Ye S, Chen Z, Zhou X, Ren L, Polydopamine-Coated Manganese Carbonate Nanoparticles for Amplified Magnetic Resonance Imaging-Guided Photothermal Therapy, ACS Applied Materials & Interfaces, 9 (2017) 19296–19306. [DOI] [PubMed] [Google Scholar]
- [222].Li L, Chen L, Huang L, Ye X, Lin Z, Wei X, Yang X, Yang Z, Biodegradable mesoporous manganese carbonate nanocomposites for LED light-driven cancer therapy via enhancing photodynamic therapy and attenuating survivin expression, Journal of Nanobiotechnology, 19 (2021) 310. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [223].Neubauer VJ, Kellner C, Gruen V, Schenk AS, Scheibel T, Recombinant major ampullate spidroin-particles as biotemplates for manganese carbonate mineralization, Multifunct. Mater, 4 (2021) 014002. [Google Scholar]
- [224].Kambe Y, Functionalization of silk fibroin-based biomaterials for tissue engineering, Polymer Journal, (2021). [Google Scholar]
- [225].Maeda S, Kawai T, Obinata M, Fujiwara H, Horiuchi T, Saeki Y, Sato Y, Furusawa M, Production of human α-interferon in silkworm using a baculovirus vector, Nature, 315 (1985) 592–594. [DOI] [PubMed] [Google Scholar]
- [226].Maruta R, Takaki K, Yamaji Y, Sezutsu H, Mori H, Kotani E, Effects of transgenic silk materials that incorporate FGF-7 protein microcrystals on the proliferation and differentiation of human keratinocytes, FASEB Bioadv, 2 (2020) 734–744. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [227].Wang Y, Wang F, Xu S, Wang R, Chen W, Hou K, Tian C, Wang F, Yu L, Lu Z, Zhao P, Xia Q, Genetically engineered bi-functional silk material with improved cell proliferation and anti-inflammatory activity for medical application, Acta Biomaterialia, 86 (2019) 148–157. [DOI] [PubMed] [Google Scholar]
- [228].Wang F, Hou K, Chen W, Wang Y, Wang R, Tian C, Xu S, Ji Y, Yang Q, Zhao P, Yu L, Lu Z, Zhang H, Li F, Wang H, He B, Kaplan DL, Xia Q, Transgenic PDGF-BB/sericin hydrogel supports for cell proliferation and osteogenic differentiation, Biomaterials Science, 8 (2020) 657–672. [DOI] [PubMed] [Google Scholar]
- [229].Chong ZX, Yeap SK, Ho WY, Transfection types, methods and strategies: a technical review, PeerJ, 9 (2021) e11165–e11165. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [230].Tamura T, Thibert C, Royer C, Kanda T, Eappen A, Kamba M, Kômoto N, Thomas J-L, Mauchamp B, Chavancy G, Shirk P, Fraser M, Prudhomme J-C, Couble P, Germline transformation of the silkworm Bombyx mori L. using a piggyBac transposon-derived vector, Nature Biotechnology, 18 (2000) 81–84. [DOI] [PubMed] [Google Scholar]
- [231].Kambe Y, Murakoshi A, Urakawa H, Kimura Y, Yamaoka T, Vascular induction and cell infiltration into peptide-modified bioactive silk fibroin hydrogels, Journal of Materials Chemistry B, 5 (2017) 7557–7571. [DOI] [PubMed] [Google Scholar]
- [232].Kambe Y, Yamaoka T, Biodegradation of injectable silk fibroin hydrogel prevents negative left ventricular remodeling after myocardial infarction, Biomater Science, 7 (2019) 4153–4165. [DOI] [PubMed] [Google Scholar]
- [233].Burger D, Yamada N, Uchino K, Shiomi K, Tamada Y, Production of recombinant silk fibroin with basic fibroblast growth factor binding affinity, The Journal of Silk Science and Technology of Japan, 29 (2021) 67–77. [Google Scholar]
- [234].Wu M, Huang S, Ye X, Ruan J, Zhao S, Ye J, Zhong B, Human epidermal growth factor-functionalized cocoon silk with improved cell proliferation activity for the fabrication of wound dressings, Journal of Biomaterials Applications, 36 (2021) 722–730. [DOI] [PubMed] [Google Scholar]
- [235].Wang F, Wang Y, Tian C, Xu S, Wang R, Hou K, Chen W, Zhao P, Yu L, Lu Z, Kaplan DL, Xia Q, Fabrication of the FGF1-functionalized sericin hydrogels with cell proliferation activity for biomedical application using genetically engineered Bombyx mori (B. mori) silk, Acta Biomaterialia, 79 (2018) 239–252. [DOI] [PubMed] [Google Scholar]
- [236].Xu S, Tan H, Yang Q, Wang R, Tian C, Ji Y, Zhao P, Xia Q, Wang F, Fabrication of a Silk Sericin Hydrogel System Delivering Human Lactoferrin Using Genetically Engineered Silk with Improved Bioavailability to Alleviate Chemotherapy-Induced Immunosuppression, ACS Applied Materials & Interfaces, 13 (2021) 45175–45190. [DOI] [PubMed] [Google Scholar]
