Next Article in Journal
Gαq Is the Specific Mediator of PAR-1 Transactivation of Kinase Receptors in Vascular Smooth Muscle Cells
Next Article in Special Issue
Biomolecule-Based Biomaterials and Their Application in Drug Delivery Systems
Previous Article in Journal
1,2,3,4,6-O-Pentagalloylglucose Protects against Acute Lung Injury by Activating the AMPK/PI3K/Akt/Nrf2 Pathway
Previous Article in Special Issue
Controlled Hydrolysis of Odorants Schiff Bases in Low-Molecular-Weight Gels
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Silk Fibroin as an Efficient Biomaterial for Drug Delivery, Gene Therapy, and Wound Healing

1
Department of Pharmaceutical Sciences, School of Applied Science and Technology, University of Kashmir, Jammu and Kashmir, Srinagar 190006, India
2
Department of Pharmaceutical Sciences, College of Pharmacy, AlMaarefa University, Ad Diriyah 13713, Saudi Arabia
3
Department of Pharmacognosy, College of Pharmacy, Prince Sattam Bin Abdulaziz University, Al-Kharj 11942, Saudi Arabia
4
Department of Pharmacy Practice, College of Pharmacy, AlMaarefa University, Ad Diriyah 13713, Saudi Arabia
5
Department of Pharmaceutics, College of Pharmacy, JSS Academy of Technical Education, Noida 201301, India
6
Department of Pharmacy Practice, East Point College of Pharmacy, Bangalore 560049, India
7
Department of Pharmaceutics, College of Pharmacy, King Saud University, Riyadh 11451, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(22), 14421; https://doi.org/10.3390/ijms232214421
Submission received: 27 March 2022 / Revised: 14 November 2022 / Accepted: 15 November 2022 / Published: 20 November 2022

Abstract

:
Silk fibroin (SF), an organic material obtained from the cocoons of a silkworm Bombyx mori, is used in several applications and has a proven track record in biomedicine owing to its superior compatibility with the human body, superb mechanical characteristics, and its controllable propensity to decay. Due to its robust biocompatibility, less immunogenic, non-toxic, non-carcinogenic, and biodegradable properties, it has been widely used in biological and biomedical fields, including wound healing. The key strategies for building diverse SF-based drug delivery systems are discussed in this review, as well as the most recent ways for developing functionalized SF for controlled or redirected medicines, gene therapy, and wound healing. Understanding the features of SF and the various ways to manipulate its physicochemical and mechanical properties enables the development of more effective drug delivery devices. Drugs are encapsulated in SF-based drug delivery systems to extend their shelf life and control their release, allowing them to travel further across the bloodstream and thus extend their range of operation. Furthermore, due to their tunable properties, SF-based drug delivery systems open up new possibilities for drug delivery, gene therapy, and wound healing.

1. Introduction

Polymeric formulations have developed as a modern alternative to conventional formulations of maintaining a supply of active pharmaceutical ingredients (APIs), optimizing their physicochemical properties, maximizing their effectiveness, and resolving many crucial issues in drug delivery, such as unique intracellular targeting transportation and biocompatibility in the process of optimizing therapeutic efficacy and patient quality of life [1,2,3,4]. In the case of incredibly harmful medications such as anti-cancer drugs, an optimal drug delivery mechanism would soothe the loaded drug, allowing it to regulate its kinetic releases, and reduce its unfavorable effects through tissue-specific targeting. For decades, silk has been recognized as an important natural material for fabric production; but in recent years, it piqued interest as a possible biopolymer for biomedical and pharmaceutical applications [5,6,7].
Several attempts have been made in recent years to develop and build different forms of nano-scale drug carriers to encapsulate corrective particles and thus unleash the drug under controlled conditions [8,9,10]. The development, as well as the formulation of new systems of drug delivery, is an exciting and rapidly growing area. The ease of access to appropriate material, the material’s complete safety to the host, and the material’s important physic-chemical and biomedical characteristics, as well as the ability to a breakdown in biological environments, are all important requirements for an outstanding drug delivery method. For drug distribution, a variety of processing materials, such as synthetic and natural polymers, have been used. So many synthesized polymers that are widely used are made with one or two monomers and have a low deterioration speed. Biodegradation in vivo can be predicted by factors such as the mass of molecules or the composition of monomers. In contrast to synthetic fibers, organic polymers have outstanding biocompatibilities as well. Numerous types of nano-particles made from artificial or organic polymers have been confirmed to be in initial clinical trials for the diagnosis of diabetes, tumor, and other sicknesses. Silk fibroin (SF) is a naturally occurring protein derived mostly from cocoons of the silkworm, Bombyx mori, and successfully utilized in biomedicine because of its supreme biocompatibility, mechanical behavior, and tunable biodegradability [11,12,13,14,15,16,17,18]. The rate of SF degradation is being managed by adjusting its molecular mass, crystalline size, or cross-linking [19,20,21].
A growing number of systems focused on SF are being used to store and distribute medicines in recent years [22,23,24,25,26,27,28]. The efficiency of drug distribution and loading in such silk drug systems is determined by its hydrophobic nature and charge, which point to several drug-release contours. Furthermore, diverse forms of delivery mechanisms such as films, hydrogels, microspheres, nanoparticles, and scaffolds may be made from silk fibroin (SF) solutions using various techniques [29]. Silk I, Silk II, and Silk III are the three structural forms of SF. Silk I is a hydrophilic protein that contains a great proportion of domains of α-helix as well as random coils [30]. Silk II, on the other hand, has a structure of β-sheet mostly that is extra stable and not water soluble, whereas Silk III dominates only at the water–air interface [31]. As a result, natural and artificial polymers can be used to develop and fabricate new therapeutics. Li and his colleagues used electrospray to build cisplatin-encapsulated SF nanostructures without utilizing organic solvents [32]. The drug particles were inserted into nanomaterials using a metal-polymer arrangement of bond sharing, as well as the drug could be released at slow speed and sustainably for fifteen days and more, according to the in vitro release studies.
The coupling of chemicals and the genomic alteration of silk-protein through chaining the amino acid sequence or inserting a segment to achieve a certain feature are the two major techniques for functionalizing silk proteins [33]. A huge number of antitumor dosage forms are formulated for parenteral delivery, resulting in close interaction with blood products. The therapeutic agents used in these preparations must not cause hematological poisoning or immune reactions [3], as a result, biocompatible polymers must be used in the preparation. Besides that, developing delivery mechanisms for therapeutic products including vaccines and antibodies necessitates ensuring both their physical integrity and biological activity, which is especially important for controlled-release mechanisms [34]. Utilizing non-modified or engineered SF protein production and formulating techniques, a broad variety of therapeutic agents of varying sizes and morphologies can be prepared (Figure 1). SF transports that have not been changed have been used to carry antitumor drugs such as doxorubicin [35], paclitaxel [36], curcumin [9,37], and cisplatin [38]. In the present review, the key techniques for developing various SF-based drug delivery systems, as well as the most recent methods for developing functionalized SF for managed or redirected drugs, gene therapy, and wound healing, are addressed.

2. Processing of SF Biomaterials or Silk Cocoon Processing—Generating Silk for Biomedical Applications

Degumming of the silk cocoon of Bombyx mori to extract sericin is one of the most important steps in producing silk appropriate for biomedical application. Enzymatic approaches (digestive process sericin not silk) or production of chemicals may be used to isolate sericin (example- treatment of alkaline). The latter process is generally utilized, and sodium bicarbonate is used to boil silk thread over 20–60 min [40].
Degumming periods as limited as five minutes are also adequate to eliminate sericin while mitigating silk destruction, which is typically caused by the disulfide’s cleavage connection between the heavy as well as light chains and separation of a sequence of amorphous silk throughout the heavy chain, resulting in polydispersed silk [41]. By dissolving the degummed cotton thread in such a chaotropic agent’s large concentration (for say, lithium bromide 9.3 M) over 60 °C for numerous hours, the silk framework of higher level can be completely reverse engineered. The resultant regenerated silk fibroin liquid is therefore dialyzed with water vigorously to create a silk aqueous solution which is steady at room temperature over half a month as well as at 4 °C over months (Figure 2) [40]. The whole reverse modified silk fibroin formulation has a lower solution conformation than native silk feedstock [42] and rheological properties are altered [43]. Novel silk materials, such as scaffolds, films, fibers, and (self-assembling) silk hydrogels, as well as (nano) fragments and (nano) abrasives, are frequently created using a reverse-optimized aqueous-regenerated silk fibroin solution within ambient circumstances. This gentle processing environment is perfect for maintaining biological function.

3. Properties of SF

SF has a rare balance of mechanical and biological attributes, due to its natural polymer-like characteristics [44,45]. Silk is often associated with a soft texture in the garment industry, but its tensile capacity and modulus make this one of the most durable natural biomaterials [46]. This property is crucial for polymers used in bone tissue reconstruction since the polymer’s mechanical efficiency is critical in such implementations [47]. Under the thermal stress (less than that of 250 °C), SF is highly stable. Lu et al. reported that differential scanning calorimetry results revealed that Silk I crystals had stable thermal properties up to 250 °C, without crystallization above the Tg, but degraded at lower temperatures than Silk II structure [48].

3.1. Physiochemical Characteristics of SF

SF has a one-of-a-kind blend of mechanical as well as biological characteristics, as well as unusual characteristics of both artificial and natural polymers [49,50]. While silk’s tensile strength and modulus make it one of nature’s most durable biomaterials, it is also synonymous with a soft texture in the garment industry [51]. This ability is essential for polymers used in bone tissue regeneration because mechanical performance is important in such implementations [52,53].

3.2. Mechanical Properties

For pharmaceutical and biomedical uses, mechanical toughness is an essential characteristic of SF-based formulations. For example, the strength of the SF product use for tissue engineering must equal that of the specific tissue. The hardness of the SF polymer will also influence its durability and degradability [54]. Numerous polymers used within drug distribution systems such as PLGA and collagen, are not solid enough mechanically. Cross-linking is a common technique for the mechanical properties of biopolymers such as collagen. The cross-linking interaction, on the other hand, may have unfavorable effects, such as immunogenicity and mitochondrial toxicity [55]. SF has a solid β-sheet construction that allows it to have outstanding mechanical characteristics without having any severe crosslinking procedures. SF can be transformed into a variety of formats, comprising liquid, hydrogels, and scaffolds depending on the material of the β-sheet [5]. Young’s modulus is commonly used to measure mechanical strength using nanoindentation strategies [56]. Because of its strong tensile strength and compressive force resistance, SF is an excellent material for drug delivery and tissue engineering [57]. Furthermore, extracting sericin during the degumming process increases the tensile strength of SF by 50% [58], rendering it more durable during physical pharmaceutical manufacturing.

3.3. Stability

Amongst the very important considerations in the manufacture of pharmaceutical formulations is the stability of polymeric products. Biopolymers are favored over synthetic alternatives for therapeutic uses because of their biocompatibility and biodegradability but they should also follow such durability requirements to be accepted for use in the pharmaceutical companies. Accumulation or gelation through extended preservation is among the most common issues with pure SF solutions. SF comes in two varieties: insoluble (high β-sheet content) and soluble (high α-helix and random coil content). Either type should be utilized and preserved, based on the medication preparation. As soluble SF is processed in extremely humid circumstances, it undergoes a transition from α-helix and random coil to β-film, that can result in gel formation and a reduction in the SF solution’s stability [59,60]. In comparison to other proteins, SF has superior heat durability. The temperature at which glass transitions (Tg), which is influenced by the β-sheet content of silk fiber, is the best measure of protein heat stability. SF film’s Tg is about 175 degrees Celsius, and the protein stays steady up to 250 degrees Celsius, which is ideal for product manufacturing. The Tg of frozen SF solution, on the other hand, is about −34~−20 °C [61], which is also advantageous in medication manufacturing at low temperatures.

3.4. Degradability and Biocompatibility

The FDA has formally approved silk fiber as a biocompatible substance for the application of a variety of nanotechnological instruments [62]. Silk’s biocompatibility has been thoroughly researched over the last two decades. In contrast to other commonly used biological polymers that degrade within the industry of pharmaceuticals, including collagen, poly(lactic-co-glycolic acid) (PLGA), and polylactide (PLA), the vast majority of researches have indicated a less immunogenic response and outstanding biocompatibility [63,64,65]. Mesenchymal stem cells, fibroblasts, hepatocytes, endothelial cells, and osteoblasts were shown to be highly compatible with SF configurations in cytocompatibility studies (MSCs) [66,67]. Methanol and hexafluoroisopropanol (HFIP) are used as organic solvents in SF manufacturing to cross-link SF by causing structural modification (α-helix to β-sheet), that linked to the SF formulations’ inflammatory ability [68]. To stop these inflammatory reactions, however, slightly different manufacturing methods have been used to avoid the use of organic solvents [69].
The capacity of biological materials to degrade is an essential attribute. Even though biodegradability is a key benefit of SF in medicinal implementations it also makes pure silk fiber particles susceptible to enzymes that break down proteins. The pace at which SF deteriorates can also be managed by adjusting the molecular mass, crystalline nature, morphological characteristics, or cross–linking [70]; however, cross-linking and the level of the crystalline phase is not the only option to consider to prevent SF from deteriorating. When SF sheets were subjected to collagenase IA in an in vitro enzymatic destruction test, the crystalline shape of the sheets changed from Silk II to Silk I. While the protease XIV enzyme was used, however, the bulk of the SF films were converted to Silk I, resulting in stronger crystallinity. Although both enzymes took 15 days to degrade, the degradation rate of protease XIV was slightly lower than that of collagenase IA [71]. Another research showed that SF deterioration induces a predictable loss of mechanical integrity [72].

4. SF-Based Biomaterial for Drug Delivery Systems

Many therapeutic applications need active pharmaceutical ingredients (APIs) to be delivered in continuous and managed release modes. The scale, composition, and other characteristics of the particles are determined by the form of the distribution device and the administration path. Furthermore, in such distribution systems, using polymers that are biocompatible and mechanically robust with moderate conditions of fabrication and manufacturing is beneficial for maintaining the loaded API’s bioactivity. As previously mentioned, SF fits both of these conditions, making it a successful candidate for drug delivery [5,73]. Nanofibers, films, lyophilized sponges, SF-coated polymeric particles, hydrogels, and micro-and nanoparticles are only some of the SF-based delivery systems of drugs that have been created. Many of the most commonly researched drug delivery of SF-based mechanisms are discussed in the following section.

4.1. Hydrogels

The SF aqueous phase was used in a variety of ways to make hydrogels. Physiochemical or chemical methods involving organic polymers or artificial reagents may initiate the conversion from solution to gel [74]. Water evaporation or osmotic stress, shearing (spinning), electric field, and warming are some of the physicochemical operations. The gel shape is stabilized by the stability in terms of thermodynamics β-sheets, which create a stable gel-type under physiological circumstances before enzymes or oxidative processes destroy it extensively [75]. Gel scaffolds containing curcumin designed by electro-gelation were used in a recent analysis for wound healing. Not only did the formulated gel formulation boosts protein adsorption and curcumin release, but it also boosted the bacterial inhibitory effect six-fold against S. aureus [76]. Protein adsorption on substrates has been known for a long time to have a vital factor in cell development and regeneration; SF gel scaffolds may help in tissue repair by encouraging cell proliferation.

4.2. Silk Films

As a biomaterial, SF has massive potential in medication formulations, tissue engineering, and in the processing of films from SF [77]. Casting an aqueous SF solution is a simple way to make SF films [78]. Certain SF film preparing strategies have been published such as a vertical deposition [79], spin coating [80], and spin-assisted layer-by-layer assembly [81]. Rather than binding independently to the hydrogel base, fibroblasts are aggregated on the rigid surface. Terada et al. [71] looked at how spin-coated SF films acted when subjected to varying ethanol concentrations. A jelly-like hydrogel coating was created with alcohol concentrations of less than 80%, while a solid film surface was created with alcohol concentrations of more than 90%. The binding of fibroblast cells to SF films was influenced by this shift in morphology [80].

4.3. Silk Particles

Nanoparticle drug delivery mechanisms have been researched the most, particularly for chemotherapeutic agents, among systems based on SF that are utilized for enclosing APIs and completing drug distribution modulation. The SF nanoparticles of lysosomotropic engineered by Seib et al. [23] for pH-dependent activation of the antitumor agent’s doxorubicin in sequence to counteract drug resistance are one example of such structures. SF nanomaterials are primarily used to deliver the primed drug to the target site in a controlled manner. SF nanoparticles can be made using a variety of processes, such as polyvinyl alcohol (PVA) blends, that can be designed to make silk fiber spheres of new shapes and sizes [81] (Table 1). The charge and lipophilicity of such systems are deciding factors for drug delivery and encapsulation quality. Different drug release profiles result from changing these conditions [82]. Furthermore, the inclusion of PVA greatly increases the morphology of the SF particles [83]. The salting-out process is one of the most common ways to make SF particles. Lammel et al., for example, used a salting-out agent potassium phosphate to create particles of SF with manageable sizes from 500 nm to 2 µm [84]. Tian et al. used the salting-out process to make SF nanoparticles, which were then filled with Fe3O4 magnetic nanomaterials and doxorubicin and guided to the tissue of interest utilizing an additional magnetic field to achieve tissue-specific guided delivery [85]. It was discovered that modifying the concentration of Fe3O4 in the formulation would affect the doxorubicin entrapment performance [85]. Moreover, current research on microfluidics used a desolvation approach to generate tinier SF spheres (150–300 nm) using a microfluidic setup (nano-assembler) [86]. The characteristics of SF nanoparticles were discovered to be influenced by two major factors: rate of flow and percentage of flow rate [87]. The use of a microfluidic device allowed for the development of SF nanomaterials with desired sizes for drug delivery that was fast, repeatable, and monitored. Monitoring the particle size and zeta potential allows for fine tuning of drug delivery.

5. SF-Based Biomaterials for Biomedical Applications

Minute particles of medicine [3], biological API drugs [91], and genes [91] have all been distributed via silk. Various silk production methods have been used to create numerous compositions for each category of therapeutic agents [92]. Stabilizing the formulated API and adjusting its circulation period to produce the desired therapeutic result are important key requirements of delivery systems based on SF. Furthermore, engineered formulations are often tailored for a specific use of drug distribution, such as preloaded drug stabilization, monitoring drug discharge, or optimizing cell adhesion [93]. The following section will include a description of SF implementations in the delivery of drugs and genes, with the review given in (Table 2 and Table 3).

5.1. SF Helps to Keep Drugs Stable

The primary aim of integrating active compounds such as minute particles or peptides into SF-based reservoirs is to stabilize them through a range of processes such as covalent interaction, adsorption, and/or enslavement [101]. Sustained drug release is difficult to perform without a safe relationship between the SF-based reservoir and drug to hold the drug active. Away from minute cases, such as the majority of stabilization and growth factors strategies focus on uniformly spreading the substance inside the SF-matrix/particles [26]. Temperature [102], humidity [103], and pH [104] all affect the stability of SF-based biomaterials. As a result, they have been extensively researched for improving the durability of other compounds, such as the encapsulation of antibiotics such as erythromycin, which has poor water stability.

5.2. Drug Delivery

Several investigators have concentrated on silk as a model or developing novel materials with customized characteristics and exceptional efficiency for a wide range of specialized applications such as tissue engineering techniques and drug delivery systems (DDSs), owing to its organizational nature and flexibility [105,106,107,108,109]. Regulated DDSs have piqued the attention of academic and industrial investigators since they were first authorized by FDA in 1990 [110]. As opposed to conventional medicine, medication systems can increase the bioavailability of the drug, maintain an adequate concentration of the drug, and reduce adverse reactions. PEGylated, Liposomes medication complex, systems based on PLGA, and protein-based mechanisms are all FDA-approved DDSs [111]. Because of their outstanding biocompatibility and pharmacokinetics, naturally existing polymer-based DDSs have recently gained a lot of publicity. Because of their uncommon combination of strong mechanical characteristics and manageable biodegradation ability, aqueous-based purification/processing, and medicine stabilization effect, SF-based substances are great options for the distribution of bioactive molecules among different groups of biopolymers. Therapeutic substances are generally maintained shipped to target locations, and distributed in a managed manner after being incorporated into the SF network [112].
Substances-based SF has been extensively studied for the delivery of antitumor substances that is the most significant title in biomedicine science, due to its drug stabilization potential. Poorly soluble in water chemotherapeutic agents can be stabilized, their bioactivity maintained after release, and therapy results improved thanks to the water-insoluble relationship with the fibroin network of β-sheet crystallites. Intratumoral and injectable administration have also been made possible with SF-based devices [113] because of the different formats that are available. For localized primary breast cancer treatment, Seib et al. used hydrogel of a SF-based to deliver doxorubicin (DOX) [114]. After sonication, aqueous SF fluids are combined with DOX and self-assembled into the thixotropic hydrogel. The packed DOX controllably discharged from the SF hydrogel after being infused locally into rats bearing breast cancer and it showed excellent tumor regression and decreased metastatic propagation. To induce the creation of nanofibers, Wu et al. prepared a DOX-loaded SF hydrogel and put it into a concentration–dilution cycle [115]. This device released DOX in a pH-responsive and concentration-dependent fashion, indicating that it may be a valuable method for regulating antitumor behavior. Phototherapy, over other chemotherapy and antitumor medications, is thought to be an effective therapeutic tool for removing tumors. Their photothermal effect and upconversion luminescence imaging output should be mixed; hydrogels based on SF have filled nanographene oxide compounds and lanthanide-doped special earth nanoparticles [116]. The theranostic hydrogels based on fibroin significantly decreased the measurement of treated tumors when exposed to NIR laser light. SF-based medicine carriers in the shape of nanomaterials are rapidly established and looked into for their possible application in intravenous, tumor care with systemic drug delivery, in contrast to hydrogel systems for cancer treatment on a local basis [117]. Furthermore, the capacity of anticancer drugs discharged from nanosized cargo to attack tumors will be advantageous for improved retention and permeability. Qu et al. described cisplatin-entrapped SF nanoparticles that were made by electrospraying [118]. Cisplatin may be distributed over 15 days thanks to its tight metal–ligand coordinate bonds with the SF matrix, and it had easy intracellular penetration and enhanced protective action on a lung tumor cell model. Tian et al. have created SF-based nanoparticles with DOX attached with the ability to target tumors magnetically [85]. The combined SF and superparamagnetic Fe3O4 nanoparticle method were rendered using the one-pot salting-out technique. DOX was delivered using external magnetic guides, resulting in good anticancer efficacy. MCF-7/ADR is a multidrug-resistant cancer cell line tumor, with a 30-day level of survival of up to 100 percent.
Nanoparticles from natural polymers, such as SF, stand out as ideal drug delivery systems due to their flexible nanostructures, biocompatibility compatibility, and custom degradation; however, the natural variability of polymers in the structure and release of drugs may limit their performance in selected conditions [117].
Larger particles are associated with the gradual release of the drug as the synthetic drug emanates from the nanoparticle area; in contrast, smaller particles indicate a faster release as the drug is closer to the nanoparticle surface [29]. In addition, small particles spread through tissues more easily than larger particles, leading to wider drug proliferation. Standing is also an important factor in the way nanoparticles behave. One study showed that circular nanoparticles had a much more efficient absorption than rod-shaped nanoparticles [23,119]. Chemical properties such as hydrophobicity and particle charge can determine the end of a target cell. Hydrophobic particles have a high potential for phagocytes, leading to targeted cell death [29].
Although the behavior of therapeutic nanoparticles is clearly multifactorial and sound conclusions cannot be drawn from in vitro therapeutic behavior, new research methods can be clarified. SF nanoparticles should be tested as a novel, in vivo system for drug delivery to wounds.
Investigators are particularly interested in using SF-based biomaterials to distribute natural possible anticancer compounds such as curcumin, in contrast to chemotherapy drugs such as DOX, paclitaxel, and cisplatin [69,120,121]. Apart from using hydrogels and nanoparticles as DDSs, there has been a lot of work put into using other SF-based biomaterials, such as medication distribution platforms. Coatings dependent on SF, for example, have been investigated as possible DDSs for tiny molecule medicines and biological agents. Bayraktar et al. coated theophylline tablets directly with a coating formulation for SF [122]. Biomedical uses include the distribution of biological particles such as proteins cell proliferation, genes, and peptides in contrast to the delivery of minute particles of drugs. As a result of its attributes, especially its tunable insolubility and very mild processing, SF may be a great member for immobilizing as well as the preservation of biological molecules. Li et al. detailed the beneficial stabilization SF biomaterials’ effect and process on biologics, such as enzymes, antibiotics, vaccines, and plasma molecules are only a few examples [94].

5.3. Controlled Drug Release

The goal of managed-release drug delivery systems is to release the embedded API in predetermined quantities over a predetermined period. Sustained-release rate is one use of such devices, which allows for the preservation of therapeutic drug doses in the bloodstream or at the action site for a prolonged period, which is important for the treatment of serious illnesses. PEG and PLGA are examples of the synthetic polymers used in the majority of presently available controlled-release formulations because they have attractive pharmacodynamic and pharmacokinetic characteristics [123]. Various functionalized SF for controlled or redirected medicines have been used (Table 4).
While the FDA has authorized PLGA as a safe ingredient in pharmaceuticals, processing conditions can limit its use in several controlled-release formulations. As a result, organic polymers such as SF, which provide kinetics of continuous release that can be tweaked and stabilization APIs that have been loaded, have currently attracted more interest for use in managed drug delivery systems. The capacity of SF to undergo various structural changes only at a molecular scale is one of its distinguishing characteristics. The increase in the ratio of alpha-helix to βeta-sheet material is the most studied structural transition in SF. The percentage of β-sheet formation, for example, influences the SF films’ permeability and release kinetics [7]. Hines and Kaplan previously explored the process of controlled release from science fiction films using numerous models [128]. The release of SF nanomaterials and microparticles under controlled conditions has been widely studied over the last decade. Song et al. [69] displayed curcumin release from SF nanoparticles is pH-controlled for up to 20 days to monitor SF molecules features, with lower pH facilitating the release. SF can be handled in aquatic conditions and crosslinked utilizing a variety of ways as a biopolymer. As a result, the SF solution has been utilized to cover a wide range of pharmaceutical formulations in single and multilayer coatings. The adenosine escape from SF-embedded powder reservoirs was computed as a characteristic of the reservoir coated layer, according to Pritchard et al. [129].

5.4. Gene Therapy

Bioengineered silk-elastin-like polymers (SELPs) have been commonly utilized as vectors for the transmission of plasmid DNA or adenoviral agents in modern genetic engineering scientific studies. SELPs’ chemical structure is made up of silk-like (Gly–Ala–Gly–Ala–Gly–Ser) and elastin-like (Gly–Val–Gly–Val–Pro) blocks that are tandemly replicated. Since they mix mechanical and thermal energy resilience of SF semi-crystalline frames with the solubility in water and durability of blocks of elastomeric elastin, silk-elastin-like polymers have been proposed. Furthermore, the ability to manipulate the monomer frames at the genomic level, as well as adenoviral viability is the unique feature of the substance that makes it ideal for gene therapy. Li et al. [130], used SELP hydrogels to distribute plasmid DNAs of various molecular weights and adeno-associated viral vector models. The DNA’s molecular mass and conformation, as well as the geometry of the hydrogel, regulated the drug release of encapsulated plasmid DNAs, and there was no substantial loss of bioactivity after 28 days. In a breast cancer-bearing mouse model, SELP hydrogel was used to deliver Renilla luciferase plasmid in vivo which resulted in a substantial improvement in luciferase gene expression and the preservation of transfection potency, which was the better option inside the tumor. Hatefi et al. [131] looked into adenoviral delivery using SELPs in vitro and in vivo. The stabilization virus interaction with silk or elastin units was due to the sustained release in vitro of imprisoned adenoviruses with the retained operation of infection after 28 days. Green fluorescent expression (GFP) was extended even 15 days after using virus-treated SELP samples after tumor treatments in vivo in xenograft murine systems, while the GFP term was reduced 11 days later with just the virus infusion, making this polymer helpful for localized adenovirus-mediated cancer therapy delivery. In the next, in a head and neck cancer xenograft model in mice, Ghandehari et al. [132] studied antitumors and the efficacy of an adenovirus-encoded SELP hydrogel. The SELP hydrogel with adenovirus was marginally more effective than the adenovirus/ganciclovir alone after 14 days of monitoring. The significant quantities of recent studies on the subject of utility, alteration, and construction of various SELPs have yielded promising results. Huang et al. studied the SELP-based patterned relationship of biomaterials of for the creation of different stimuli-responsive structures, as well as their present and future applications [133].
Pre- and post-loading is used to integrate pDNA into SF microcapsules to find the best delivery mechanism in terms of the encapsulation of drugs, toxicity of cells, discharge of drugs, and the transfection performance. Prior loading was performed by electrostatically before SF deposition, adsorbing pDNA onto bPEI25-coated PS particles, resulting in nucleic acid/bPEI25 complexes entangled within the microspheres, with the SF multi-layer shell serving as both a defensive and a diffusion impediment (Figure 3A,B). In the post-loading process, the cargo is loaded onto or into pre-fabricated capsules by adsorption and/or diffusion through the capsule casing (Figure 3A,B) [134]. Several studies [135,136,137] have successfully employed these techniques and the loading process and shell thickness is used to customize drug release characteristics [138].
Sustained pDNA release was observed in both classes, as measured by the decrease in Cy5-pDNA fluorescence associated with microcapsules, which was substantially in the presence of Protease XIV, the concentration is greater than in pure PBS. In the case of preloaded microcapsules, discharge is thought to occur as a result of absorption and desorption, while pDNA release after post-loading desorption is believed to be the cause (Figure 4A). According to Lu et al., the first hydrophilic blocks are damaged through protease, blocks with high crystallinity are left behind, eventually move into the solution as particles [139]. Since shell porosity, pDNA release, and desorption all increased as SF depleted, the protease was significantly both preloaded and post-loaded microcapsule improved. We also discovered that after pre- and post-loading, the transfection performance of 1 mm microcapsules was slightly higher than that of 4 mm microcapsules. We believe that the density of microcapsules on the cell surface causes this effect, which leads to a greater likelihood of encounters and better touch with the surface of the cell as capsule size is reduced (Figure 4B). Although further research is needed, this result may be another influence on capsule size in cytotoxicity. Even though it was less distinct than its effect on transfection efficiency, it was a significant parameter for balancing cytotoxicity and transfection quality.
When contrasted to pDNA/bPEI25 complexes, SF microcapsules filled with plasmid DNA transduced NIH/3T3 fibroblasts successfully, eliminating cytotoxic effects. The findings indicate that SF microcapsules have the potential to be a powerful carrier for regulated, localized gene delivery.

5.5. Wound Healing

Hemostasis, inflammation, replication, and remodeling are all part of the wound healing process, which is a fluid and dynamic process. Weak mechanical efficiency, high expense, collagen inconstancy, poor processability, restricted supply of elastins, soft silicons, and polyurethanes being non-biodegradable are all disadvantages of the currently used wound dressing products. Because of its availability, inherent biodegradation ability, biocompatibility, mechanical robustness, signaling molecules stability ability, high synthesis of water and oxygen, and poor immunogenicity, SF may be considered an outstanding wound healing material (Table 5). SF facilitates regeneration by more complex pathways and signaling pathways, which have been briefly summarised by Farokhi et al. in addition to the above-mentioned properties [140].
Injury covering matrices are normally filled with growth factors to facilitate epithelization, in contrast to antimicrobial compounds to avoid the growth of bacteria. Schneider et al. tested the tissue repair ability of epidermal growth factor-containing electrospun (EGF) silk mats on human skin-equivalent models [150]. Gil et al. studied the impact of various silk fiber content layouts and in related research, the effects of the drug loading approach on wound healing in vivo were studied and aimed at evaluating the feasibility of EGF/silver sulfadiazine/SF systems on wound covering applications [151]. Woong et al. reportedly immobilized a wound-healing antimicrobial peptide, Cys-KR12, onto an SF nanofiber layer [152]. It has been reported that insulin-loaded SF microparticles embedded within SF sponges are used to treat chronic cutaneous wounds [153]. Keratinocytes and endothelial cells can experience enhanced multiplication and differentiation as a result of insulin’s stimulatory impact. The use of SF-calcium alginate-carboxymethyl cellulose hydrogel, an SF-based blending composite, for the treatment of burn wounds was stated by Kim et al. [154]. Vasconcelos et al. also created a blending hydrogel device that merged elastin’s durability, elasticity, and silk’s tunable biodegradation and high mechanical strength made it ideal for use as a wound dressing medium [155]. Zhang et al. carried out a unique translational study in which clinically focused and extensive preclinical experiments were carried out on full-thickness skin defect models in rabbits and porcines, as well as randomized controlled clinical trials on human subjects [156].
According to Ju et al., [157] they used a modified electrospun device combined with a mimetics (i.e., sodium chloride crystal) dispensing apparatus to fabricate SF nano matrix with total density and wide holes, and we checked the burn tissue repair effect in rats using a deep second-degree burned mouse model. Histological findings were used to examine the wound recovery mechanism and an RT-PCR assay was used to establish the healing process in contrast to a freely available market dressing (i.e., polyurethane foam and Medifoam®). They looked at histological changes in damaged skin to see if the SF nano matrix treatment affected wound healing (Figure 5). On day one, the presence of a blister and edema indicated a second-degree burn without any tissue injury to the underlying fascia and muscle tissue.
As contrasted to the medical gauze-treated community, the collagen array in the SF nanomatrix and Medifoam®-treated groups was much thicker and far more constant [157]. Scar elimination, collagen and epithelialization, and PCNA expression are all factors to consider both confirmed that the SF nanomatrix could speed up the recovery of severe burns in rats (Figure 5, Figure 6 and Figure 7).
The mechanism of re-epithelialization started seven days after operation in both classes. The nano matrix party at SF, in particular, demonstrated quicker cell proliferation in the burned area (S), whereas clots of fibrin existed on the wounded area’s surface in the party that was treated with surgical gauze (C). Groups treated with SF nanomatrix and Medifoam® (S and M) demonstrated quicker processes of re-epithelialization than the surgical gauze-treated community over the entire healing cycle (C) (Figure 6).
The wounds managed with the SF nano matrix had morphogenesis and histology compared to usual skin after 14 days, and the wound region had regenerated without edema or granulation tissue. The medical gauze party, on the other hand, demonstrated extreme neutrophil and lymphocyte infiltration. On day 28, collagen accumulation at the injury site is seen in Figure 8.
Within 14 days after treatment, wound healing improved as the wounds were treated with SF nanomatrix skin regeneration and cell aggregation relative to injury healing with medical gauze, according to the wound size measurements (Figure 8). SF nanomatrix’s rapid wound healing may be due to a variety of factors. Human fibroblasts and keratinocytes have been shown to proliferate when exposed to SF [158], as well as collagen deposition [159].
Figure 5. Burn wound tissues stained with hematoxylin and eosin (C: surgical gauze, S: SF nanomatrix, and M: Medifoam®, scale: 100 µm). Reprinted with permission from reference [157].
Figure 5. Burn wound tissues stained with hematoxylin and eosin (C: surgical gauze, S: SF nanomatrix, and M: Medifoam®, scale: 100 µm). Reprinted with permission from reference [157].
Ijms 23 14421 g005
Figure 6. Photos of burn wound tissues stained with H&E at 7, 14, and 21 days. (A) Hospital gauze (14 days), (B) SF nanomatrix (7 days), and (C) Medifoam® (7 days). (Scale: 50 µm, NT: necrosis tissue, nt: natural tissue, circle: Keratinocytes). Reprinted with permission from reference [157].
Figure 6. Photos of burn wound tissues stained with H&E at 7, 14, and 21 days. (A) Hospital gauze (14 days), (B) SF nanomatrix (7 days), and (C) Medifoam® (7 days). (Scale: 50 µm, NT: necrosis tissue, nt: natural tissue, circle: Keratinocytes). Reprinted with permission from reference [157].
Ijms 23 14421 g006
Figure 7. At 7 days, PCNA expression was observed in the tissue covering the infected region. Medical gauze (A), SF nanomatrix (B), and Medifoam® (C) (scale: 50 µm). Reprinted with permission from reference [157].
Figure 7. At 7 days, PCNA expression was observed in the tissue covering the infected region. Medical gauze (A), SF nanomatrix (B), and Medifoam® (C) (scale: 50 µm). Reprinted with permission from reference [157].
Ijms 23 14421 g007
Figure 8. (A) Burn wound area on rat skin right after the creation. (B) Residual wound area change with healing time (28 days). (C) Gross findings of wound area treated with different wound dressing materials (C: medical gauze, S: SF nanomatrix, and M: Medifoam®) (scale: 50 µm). Reprinted with permission from reference [157].
Figure 8. (A) Burn wound area on rat skin right after the creation. (B) Residual wound area change with healing time (28 days). (C) Gross findings of wound area treated with different wound dressing materials (C: medical gauze, S: SF nanomatrix, and M: Medifoam®) (scale: 50 µm). Reprinted with permission from reference [157].
Ijms 23 14421 g008

6. Conclusions

Silk is a flexible biomaterial with several promises in terms of gene and drug delivery. SF has been used to make SF films, hydrogels, microparticles, and nanoparticles, among other drug delivery methods, employing a variety of manufacturing procedures. Every one of these SF-based frameworks has represented promise in a range of biomedical applications. Curcumin, doxorubicin, and ibuprofen, as well as pDNA, have all been delivered to different types of cells utilizing SF micro- and nanoparticles in a time and site-specific way. Drugs such as dextran and epirubicin, as well as biological agents such as IgG and HIV inhibitor 5P12-RANTES, have been controlled released using SF films. Furthermore, they have been used to keep biomedical agents such as horseradish peroxidase (HRP), oxidase of glucose, vaccines, and monoclonal antibodies fresh for longer. SF has also been used to extend the release and biological role of biomolecules including insulin and BMP-2. SF has been loaded with specific biological elements such as the RGD sequence, folate, and Her2 for tissue-specific drug delivery. SF has been used to cover the surfaces of polymer micro materials and liposomes to alter their release kinetics or improve cell adhesion, in comparison to drug carriers that depend on it. Another area that requires further study is changing the SF’s physicochemical and mechanical properties by mixing it with other inorganic fillers to create engineered SF-based biomaterials. Furthermore, due to their tunable properties, SF-based drug delivery systems open up new possibilities for drug delivery, gene therapy, and wound healing.

Author Contributions

Conceptualization, S.U.D.W. and F.S.; methodology, M.I.Z., M.H.M., H.G.S. and M.A.; software, S.A.; validation, F.S., S.A. and S.U.D.W.; formal analysis, M.M.G.; investigation, F.S., S.A., A.A. and P.A.; resources, S.A.; data curation, A.A.; writing—original draft preparation, S.U.D.W.; writing—review and editing, S.A., P.A. and F.S.; visualization, S.A.; supervision, F.S.; project administration, S.U.D.W. and F.S.; funding acquisition, S.A., M.M.G. and A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This work received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

This article did not report any data.

Acknowledgments

The authors are thankful to AlMaarefa University for their financial support for this article. The authors are also thankful to the Head, Department of Pharmaceutical Sciences, Srinagar, School of Applied Science and Technology, University of Kashmir, India for providing necessary support and encouragement to hoard the work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liechty, W.B.; Kryscio, D.R.; Slaughter, B.V.; Peppas, N.A. Polymers for drug delivery systems. Annu. Rev. Chem. Biomol. Eng. 2010, 1, 149–173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Torchilin, V.P. Multifunctional, stimuli-sensitive nanoparticulate systems for drug delivery. Nat. Rev. Drug Discov. 2014, 13, 813–827. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Pritchard, E.M.; Valentin, T.; Panilaitis, B.; Omenetto, F.; Kaplan, D.L. Antibiotic-releasing silk biomaterials for infection prevention and treatment. Adv. Funct. Mater. 2013, 23, 854–861. [Google Scholar] [CrossRef] [PubMed]
  4. Luo, Z.; Li, J.; Qu, J.; Sheng, W.; Yang, J.; Li, M. Cationized Bombyx mori silk fibroin as a delivery carrier of the VEGF165–Ang-1 coexpression plasmid for dermal tissue regeneration. J. Mater. Chem. B 2019, 7, 80–94. [Google Scholar] [CrossRef]
  5. Wenk, E.; Merkle, H.P.; Meinel, L. Silk fibroin as a vehicle for drug delivery applications. J. Control. Release 2011, 150, 128–141. [Google Scholar] [CrossRef]
  6. Kim, S.Y.; Naskar, D.; Kundu, S.C.; Bishop, D.P.; Doble, P.A.; Boddy, A.V.; Chan, H.-K.; Wall, I.B.; Chrzanowski, W. Formulation of biologically-inspired silk-based drug carriers for pulmonary delivery targeted for lung cancer. Sci. Rep. 2015, 5, 11878. [Google Scholar] [CrossRef] [Green Version]
  7. Choi, M.; Choi, D.; Hong, J. Multilayered controlled drug release silk fibroin nanofilm by manipulating secondary structure. Biomacromolecules 2018, 19, 3096–3103. [Google Scholar] [CrossRef]
  8. Karimi, M.; Zangabad, P.S.; Mehdizadeh, F.; Malekzad, H.; Ghasemi, A.; Bahrami, S.; Zare, H.; Moghoofei, M.; Hekmatmanesh, A.; Hamblin, M.R. Nanocaged platforms: Modification, drug delivery and nanotoxicity. Opening synthetic cages to release the tiger. Nanoscale 2017, 9, 1356–1392. [Google Scholar]
  9. Kamaly, N.; Yameen, B.; Wu, J.; Farokhzad, O.C. Degradable controlled-release polymers and polymeric nanoparticles: Mechanisms of controlling drug release. Chem. Rev. 2016, 116, 2602–2663. [Google Scholar] [CrossRef] [Green Version]
  10. Peer, D.; Karp, J.M.; Hong, S.; Farokhzad, O.C.; Margalit, R.; Langer, R. Nanocarriers as an emerging platform for cancer therapy. Nat. Nanotechnol. 2007, 2, 751–760. [Google Scholar] [CrossRef]
  11. Tao, H.; Kaplan, D.L.; Omenetto, F.G. Silk materials-a road to sustainable high technology. Adv. Mater. 2012, 24, 2824–2837. [Google Scholar] [CrossRef]
  12. Omenetto, F.G.; Kaplan, D.L. New opportunities for an ancient material. Science 2010, 329, 528–531. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Shao, Z.; Vollrath, F. Materials: Surprising strength of silkworm silk. Nature 2002, 418, 741. [Google Scholar] [CrossRef] [PubMed]
  14. Vepari, C.; Kaplan, D.L. Silk as a biomaterial. Prog. Polym. Sci. 2007, 32, 991–1007. [Google Scholar] [CrossRef] [PubMed]
  15. Kundu, B.; Rajkhowa, R.; Kundu, S.C.; Wang, X.G. Silk fibroin biomaterials for tissue regenerations. Adv. Drug Deliv. Rev. 2013, 65, 457–470. [Google Scholar] [CrossRef] [PubMed]
  16. Lin, S.; Ryu, S.; Tokareva, O.; Gronau, G.; Jacobsen, M.M.; Huang, W.; Rizzo, D.J.; Li, D.; Staii, C.; Pugno, N.M.; et al. Predictive modelling-based design and experiments for synthesis and spinning of bioinspired silk fibres. Nat. Commun. 2015, 6, 6892. [Google Scholar] [CrossRef] [Green Version]
  17. Kim, D.-H.; Viventi, J.; Amsden, J.J.; Xiao, J.; Vigeland, L.; Kim, Y.-S.; Blanco, J.A.; Panilaitis, B.; Frechette, E.S.; Contreras, D.; et al. Dissolvable films of silk fibroin for ultrathin conformal bio-integrated electronics. Nat. Mater. 2010, 9, 511–517. [Google Scholar] [CrossRef] [Green Version]
  18. Al Saqr, A.; Wani, S.U.D.; Gangadharappa, H.V.; Aldawsari, M.F.; Khafagy, E.-S.; Lila, A.S.A. Enhanced cytotoxic activity of docetaxel-loaded silk fibroin nanoparticles against breast cancer cells. Polymers 2021, 13, 1416. [Google Scholar] [CrossRef]
  19. Horan, R.L.; Antle, K.; Collette, A.L.; Wang, Y.; Huang, J.; Moreau, J.E.; Volloch, V.; Kaplan, D.L.; Altman, G.H. In vitro degradation of silk fibroin. Biomaterials 2005, 26, 3385–3393. [Google Scholar] [CrossRef]
  20. Li, M.-Y.; Zhao, Y.; Tong, T.; Hou, X.-H.; Fang, B.-S.; Wu, S.-Q.; Shen, X.-Y.; Tong, H. Study of the degradation mechanism of Chinese historic silk (Bombyx mori) for the purpose of conservation. Polym. Degrad. Stab. 2013, 98, 727–735. [Google Scholar] [CrossRef]
  21. You, R.; Zhang, Y.; Liu, Y.; Liu, G.; Li, M. The degradation behavior of silk fibroin derived from different ionic liquid solvents. Nat. Sci. 2013, 5, 10–19. [Google Scholar] [CrossRef]
  22. Huang, X.-W.; Liang, H.; Li, Z.; Zhou, J.; Chen, X.; Bai, S.-M.; Yang, H.-H. Monodisperse phase transfer and surface bioengineering of metal nanoparticles via a silk fibroin protein corona. Nanoscale 2017, 9, 2695–2700. [Google Scholar] [CrossRef] [PubMed]
  23. Seib, F.P.; Jones, G.T.; Rnjak-Kovacina, J.; Lin, Y.; Kaplan, D.L. pH-dependent anticancer drug release from silk nanoparticles. Adv. Healthc. Mater. 2013, 2, 1606–1611. [Google Scholar] [CrossRef] [Green Version]
  24. Wani, S.U.D.; Gautam, S.P.; Qadrie, Z.L.; Gangadharappa, H.V. Silk fibroin as a natural polymeric based bio-material for tissue engineering and drug delivery systems—A review. Int. J. Biol. Macromol. 2020, 163, 2145–2161. [Google Scholar] [CrossRef]
  25. Kundu, J.; Chung, Y.-I.; Kim, Y.H.; Tae, G.; Kundu, S.C. Silk fibroin nanoparticles for cellular uptake and control release. Int. J. Pharm. 2010, 388, 242–250. [Google Scholar] [CrossRef] [PubMed]
  26. Yucel, T.; Lovett, M.L.; Kaplan, D.L. Silk-based biomaterials for sustained drug delivery. J. Control. Release 2014, 190, 381–397. [Google Scholar] [CrossRef] [Green Version]
  27. Wani, S.U.D.; Gangadharappa, H.V. Silk fibroin based drug delivery applications: Promises and challenge. Curr. Drug Targets 2018, 19, 1177–1190. [Google Scholar] [CrossRef]
  28. Wani, S.U.D.; Gangadharappa, H.V.; Ashish, N.P. Formulation, development and characterization of drug delivery systems based telmisartan encapsulated in silk fibroin nanospheres. Int. J. Appl. Pharam. 2019, 11, 247–254. [Google Scholar] [CrossRef]
  29. Seib, F.P. Reverse-engineered silk hydrogels for cell and drug delivery. Ther. Deliv. 2018, 9, 469–487. [Google Scholar] [CrossRef] [Green Version]
  30. Jin, H.-J.; Kaplan, D.L. Mechanism of silk processing in insects and spiders. Nature 2003, 424, 1057–1061. [Google Scholar] [CrossRef]
  31. Crivelli, B.; Perteghella, S.; Bari, E.; Sorrenti, M.; Tripodo, G.; Chlapanidas, T.; Torre, M.L. Silk nanoparticles: From inert supports to bioactive natural carriers for drug delivery. Soft Matter 2018, 14, 546–557. [Google Scholar] [CrossRef] [PubMed]
  32. Yu, D.G.; Wang, X.; Li, X.Y.; Chian, W.; Li, Y.; Liao, Y.Z. Electrospun biphasic drug release polyvinylpyrrolidone/ethyl cellulose core/sheath nanofibers. Acta Biomater. 2013, 9, 5665–5672. [Google Scholar] [CrossRef] [PubMed]
  33. Deptuch, T.; Dams-Kozlowska, H. Silk materials functionalized via genetic engineering for biomedical applications. Materials 2017, 10, 1417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Guziewicz, N.A.; Massetti, A.J.; Perez-Ramirez, B.J.; Kaplan, D.L. Mechanisms of monoclonal antibody stabilization and release from silk biomaterials. Biomaterials 2013, 34, 7766–7775. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Totten, J.D.; Wongpinyochit, T.; Seib, F.P. Silk nanoparticles: Proof of lysosomotropic anticancer drug delivery at single-cell resolution. J. Drug Target. 2017, 25, 865–872. [Google Scholar] [CrossRef] [Green Version]
  36. Chen, M.; Shao, Z.; Chen, X. Paclitaxel-loaded silk fibroin nanospheres. J. Biomed. Mater. Res. Part A 2012, 100A, 203–210. [Google Scholar] [CrossRef]
  37. Tomeh, M.A.; Hadianamrei, R.; Zhao, X. A review of curcumin and its derivatives as anticancer agents. Int. J. Mol. Sci. 2019, 20, 1033. [Google Scholar] [CrossRef] [Green Version]
  38. Yavuz, B.; Zeki, J.; Taylor, J.; Harrington, K.; Coburn, J.M.; Ikegaki, N.; Kaplan, D.L.; Chiu, B. Silk Reservoirs for Local Delivery of Cisplatin for Neuroblastoma Treatment: In Vitro and In Vivo Evaluations. J Pharm. Sci. 2019, 108, 2748–2755. [Google Scholar] [CrossRef]
  39. Tomeh, M.A.; Hadianamrei, R.; Zhao, X. Silk fibroin as a functional biomaterial for drug and gene delivery. Pharmaceutics 2019, 11, 494. [Google Scholar] [CrossRef] [Green Version]
  40. Rockwood, D.N.; Preda, R.C.; Yucel, T.; Wang, X.; Lovett, M.L.; Kaplan, D.L. Materials fabrication from Bombyx mori silk fibroin. Nat. Protoc. 2011, 6, 1612–1631. [Google Scholar] [CrossRef] [Green Version]
  41. Wray, L.S.; Hu, X.; Gallego, J.; Georgakoudi, I.; Omenetto, F.G.; Schmidt, D.; Kaplan, D.L. Effect of processing on silk-based biomaterials: Reproducibility and biocompatibility. J. Biomed. Mater. Res. Part B 2011, 99, 89–101. [Google Scholar] [CrossRef] [PubMed]
  42. Greving, I.; Dicko, C.; Terry, A.; Callowc, P.; Vollrath, F. Small angle neutron scattering of native and reconstituted silk fibroin. Soft Matter 2010, 6, 4389–4395. [Google Scholar] [CrossRef]
  43. Holland, C.; Terry, A.E.; Porter, D.; Vollrath, F. Natural and Unnatural Silks. Polymer 2007, 48, 3388–3392. [Google Scholar] [CrossRef]
  44. Yao, D.; Dong, S.; Lu, Q.; Hu, X.; Kaplan, D.L.; Zhang, B.; Zhu, H. Salt-leached silk scaffolds with tunable mechanical properties. Biomacromolecules 2012, 13, 3723–3729. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Theodora, C.; Sara, P.; Silvio, F.; Alessandra, B.; Giuseppe, T.; Barbara, V.; Barbara, C.; Sabrina, R.; Silvia, D.; Stefania, P. Platelet lysate and adipose mesenchymal stromal cells on silk fibroin nonwoven mats for wound healing. J. Appl. Polym. Sci. 2016, 133, E42942. [Google Scholar] [CrossRef]
  46. Li, G.; Li, Y.; Chen, G.; He, J.; Han, Y.; Wang, X.; Kaplan, D.L. Silk-based biomaterials in biomedical textiles and fiber-based implants. Adv. Healthc. Mater. 2015, 4, 1134–1151. [Google Scholar] [CrossRef]
  47. Luo, K.; Yang, Y.; Shao, Z. Physically crosslinked biocompatible silk-fibroin-based hydrogels with high mechanical performance. Adv. Funct. Mater. 2016, 26, 872–880. [Google Scholar] [CrossRef]
  48. Lu, Q.; Hu, X.; Wang, X.; Kluge, J.A.; Lu, S.; Cebe, P.; Kaplan, D.L. Water-insoluble silk films with silk I structure. Acta Biomater. 2010, 6, 1380–1387. [Google Scholar] [CrossRef] [Green Version]
  49. Sun, W.; Gregory, D.A.; Tomeh, M.A.; Zhao, X. Silk fibroin as a functional biomaterial for tissue engineering. Int. J. Mol. Sci. 2021, 22, 1499. [Google Scholar] [CrossRef]
  50. Bossi, A.M.; Bucciarelli, A.; Maniglio, D. Molecularly imprinted silk fibroin nanoparticles. ACS Appl. Mater. Interfaces 2021, 13, 31431–31439. [Google Scholar] [CrossRef]
  51. Belbeoch, C.; Lejeune, J.; Vroman, P.; Salaun, F. Silkworm and spider silk electrospinning: A review. Environ. Chem. Lett. 2021, 19, 1737–1763. [Google Scholar] [CrossRef] [PubMed]
  52. Laity, P.R.; Holland, C. Thermo-rheological behavior of native silk feedstocks. Eur. Polym. J. 2017, 87, 519–534. [Google Scholar] [CrossRef] [Green Version]
  53. Lyons, J.G.; Plantz, M.A.; Hsu, W.K.; Hsu, E.L.; Minardi, S. Nanostructured biomaterials for bone regeneration. Front. Bioeng. Biotechnol. 2020, 8, 922. [Google Scholar] [CrossRef] [PubMed]
  54. Zhao, S.; Chen, Y.; Partlow, B.P.; Golding, A.S.; Tseng, P.; Coburn, J.; Applegate, M.B.; Moreau, J.E.; Omenetto, F.G.; Kaplan, D.L. Bio-functionalized silk hydrogel microfluidic systems. Biomaterials 2016, 93, 60–70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Bhrany, A.D.; Lien, C.J.; Beckstead, B.L.; Futran, N.D.; Muni, N.H.; Giachelli, C.M.; Ratner, B.D. Crosslinking of an oesophagus acellular matrix tissue scaffold. J. Tissue Eng. Regen. Med. 2008, 2, 365–372. [Google Scholar] [CrossRef] [PubMed]
  56. Kurland, N.E.; Drira, Z.; Yadavalli, V.K. Measurement of nanomechanical properties of biomolecules using atomic force microscopy. Micron 2012, 43, 116–128. [Google Scholar] [CrossRef]
  57. Altman, G.H.; Horan, R.L.; Lu, H.H.; Moreau, J.; Martin, I.; Richmond, J.C.; Kaplan, D.L. Silk matrix for tissue engineered anterior cruciate ligaments. Biomaterials 2002, 23, 4131–4141. [Google Scholar] [CrossRef]
  58. Kundu, B.; Kurland, N.E.; Bano, S.; Patra, C.; Engel, F.B.; Yadavalli, V.K.; Kundu, S.C. Silk proteins for biomedical applications: Bioengineering perspectives. Prog. Polym. Sci. 2014, 39, 251–267. [Google Scholar] [CrossRef]
  59. Min, B.M.; Jeong, L.; Lee, K.Y.; Park, W.H. Regenerated silk fibroin nanofibers: Water vapor-induced structural changes and their effects on the behavior of normal human cells. Macromol. Biosci. 2006, 6, 285–292. [Google Scholar] [CrossRef]
  60. Jin, H.J.; Park, J.; Karageorgiou, V.; Kim, U.J.; Valluzzi, R.; Cebe, P.; Kaplan, D.L. Water-stable silk films with reduced β-sheet content. Adv. Funct. Mater. 2005, 15, 1241–1247. [Google Scholar] [CrossRef]
  61. Li, M.; Lu, S.; Wu, Z.; Yan, H.; And, J.M.; Wang, L. Study on porous silk fibroin materials. I. Fine structure of freeze dried silk fibroin. J. Appl. Polym. Sci. 2001, 79, 2185–2191. [Google Scholar] [CrossRef]
  62. Melke, J.; Midha, S.; Ghosh, S.; Ito, K.; Hofmann, S. Silk fibroin as biomaterial for bone tissue engineering. Acta Biomater. 2016, 31, 1–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Makadia, H.K.; Siegel, S.J. Poly lactic-co-glycolic acid (PLGA) as biodegradable controlled drug delivery carrier. Polymers 2011, 3, 1377–1397. [Google Scholar] [CrossRef] [PubMed]
  64. Numata, K.; Hamasaki, J.; Subramanian, B.; Kaplan, D.L. Gene delivery mediated by recombinant silk proteins containing cationic and cell binding motifs. J. Control. Release 2010, 146, 136–143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Meinel, L.; Karageorgiou, V.; Hofmann, S.; Fajardo, R.; Snyder, B.; Li, C.; Zichner, L.; Langer, R.; Vunjak-Novakovic, G.; Kaplan, D.L. Engineering bone-like tissue in vitro using human bone marrow stem cells and silk scaffolds. J. Biomed. Mater. Res. A 2004, 71, 25–34. [Google Scholar] [CrossRef]
  66. Jin, H.J.; Chen, J.; Karageorgiou, V.; Altman, G.H.; Kaplan, D.L. Human bone marrow stromal cell responses on electrospun silk fibroin mats. Biomaterials 2004, 25, 1039–1047. [Google Scholar] [CrossRef]
  67. Meinel, L.; Hofmann, S.; Betz, O.; Fajardo, R.; Merkle, H.P.; Langer, R.; Evans, C.H.; Vunjak-Novakovic, G.; Kaplan, D.L. Osteogenesis by human mesenchymal stem cells cultured on silk biomaterials: Comparison of adenovirus mediated gene transfer and protein delivery of BMP-2. Biomaterials 2006, 27, 4993–5002. [Google Scholar] [CrossRef]
  68. Meinel, L.; Hofmann, S.; Karageorgiou, V.; Kirker-Head, C.; McCool, J.; Gronowicz, G.; Zichner, L.; Langer, R.; Vunjak-Novakovic, G.; Kaplan, D.L. The inflammatory responses to silk films in vitro and in vivo. Biomaterials 2005, 26, 147–155. [Google Scholar] [CrossRef]
  69. Hai-Yan, W.; Yu-Qing, Z.; Zheng-Guo, W. Dissolution and processing of silk fibroin for materials science. Crit. Rev. Biotechnol. 2021, 41, 406–424. [Google Scholar]
  70. Kundu, J.; Poole-Warren, L.A.; Martens, P.; Kundu, S.C. Silk fibroin/poly(vinyl alcohol) photocrosslimked hydrogels for delivery of macromolecular drugs. Acta Biomater. 2012, 8, 1720–1729. [Google Scholar] [CrossRef]
  71. Li, M.; Ogiso, M.; Minoura, N. Enzymatic degradation behavior of porous silk fibroin sheets. Biomaterials 2003, 24, 357–365. [Google Scholar] [CrossRef]
  72. Nguyen, T.P.; Nguyen, Q.V.; Nguyen, V.H.; Le, T.H.; Huynh, V.; Vo, D.N.; Trinh, Q.T.; Kim, S.Y.; Le, Q.V. Silk fibroin-based biomaterials for biomedical applications: A review. Polymers 2019, 11, 1933. [Google Scholar] [CrossRef] [PubMed]
  73. Liu, B.; Song, Y.-W.; Jin, L.; Wang, Z.-J.; Pu, D.-Y.; Lin, S.-Q.; Zhou, C.; You, H.-J.; Ma, Y.; Li, J.-M.; et al. Silk structure and degradation. Colloids Surf. B 2015, 131, 122–128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Chirila, T.V.; Suzuki, S.; Papolla, C. A comparative investigation of Bombyx mori silk fibroin hydrogels generated by chemical and enzymatic cross-linking. Biotechnol. Appl. Biochem. 2017, 64, 771–781. [Google Scholar] [CrossRef]
  75. Matsumoto, A.; Chen, J.; Collette, A.L.; Kim, U.-J.; Altman, G.H.; Cebe, P.; Kaplan, D.L. Mechanisms of silk fibroin sol-gel transitions. J. Phys. Chem. B 2006, 110, 21630–21638. [Google Scholar] [CrossRef] [PubMed]
  76. Karahaliloglu, Z. Curcumin-loaded silk fibroin e-gel scaffolds for wound healing applications. Mater. Technol. 2018, 33, 276–287. [Google Scholar] [CrossRef]
  77. Lee, M.C.; Kim, D.-K.; Lee, O.J.; Kim, J.-H.; Ju, H.W.; Lee, J.M.; Moon, B.M.; Park, H.J.; Kim, D.W.; Kim, S.H.; et al. Fabrication of silk fibroin film using centrifugal casting technique for corneal tissue engineering. J. Biomed. Mater. Res. Part B 2016, 104, 508–514. [Google Scholar] [CrossRef]
  78. Qi, Y.; Wang, H.; Wei, K.; Yang, Y.; Zheng, R.Y.; Kim, I.S.; Zhang, K.Q. A Review of Structure Construction of Silk Fibroin Biomaterials from Single Structures to Multi-Level Structures. Int. J. Mol. Sci. 2017, 18, 237. [Google Scholar] [CrossRef]
  79. Sagnella, A.; Pistone, A.; Bonetti, S.; Donnadio, A.; Saracino, E.; Nocchetti, M.; Dionigi, C.; Ruani, G.; Muccini, M.; Posati, T.; et al. Effect of different fabrication methods on the chemo-physical properties of silk fibroin films and on their interaction with neural cells. RSC Adv. 2016, 6, 9304–9314. [Google Scholar] [CrossRef]
  80. Terada, D.; Yokoyama, Y.; Hattori, S.; Kobayashi, H.; Tamada, Y. The outermost surface properties of silk fibroin films reflect ethanol-treatment conditions used in biomaterial preparation. Mater. Sci. Eng. C 2016, 58, 119–126. [Google Scholar] [CrossRef]
  81. Wang, X.; Yucel, T.; Lu, Q.; Hu, X.; Kaplan, D.L. Silk nanospheres and microspheres from silk/pva blend films for drug delivery. Biomaterials 2010, 31, 1025–1035. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Cao, Y.; Liu, F.; Chen, Y.; Yu, T.; Lou, D.; Guo, Y.; Li, P.; Wang, Z.; Ran, H. Drug release from core-shell PVA/silk fibroin nanoparticles fabricated by one-step electrospraying. Sci. Rep. 2017, 7, 11913. [Google Scholar] [CrossRef]
  83. Shi, P.; Goh, J.C.H. Self-assembled silk fibroin particles: Tunable size and appearance. Powder Technol. 2012, 215–216, 85–90. [Google Scholar] [CrossRef]
  84. Lammel, A.S.; Hu, X.; Park, S.H.; Kaplan, D.L.; Scheibel, T.R. Controlling silk fibroin particle features for drug delivery. Biomaterials 2010, 31, 4583–4591. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Tian, Y.; Jiang, X.; Chen, X.; Shao, Z.; Yang, W. Doxorubicin-loaded magnetic silk fibroin nanoparticles for targeted therapy of multidrug-resistant cancer. Adv. Mater. 2014, 26, 7393–7398. [Google Scholar] [CrossRef] [PubMed]
  86. Wongpinyochit, T.; Totten, J.D.; Johnston, B.F.; Seib, F.P. Microfluidic-assisted silk nanoparticle tuning. Nanoscale Adv. 2019, 1, 873–883. [Google Scholar] [CrossRef] [Green Version]
  87. Guziewicz, N.; Best, A.; Perez-Ramirez, B.; Kaplan, D.L. Lyophilized silk fibroin hydrogels for the sustained local delivery of therapeutic monoclonal antibodies. Biomaterials 2011, 32, 2642–2650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Myung, S.J.; Kim, H.-S.; Kim, Y.; Chen, P.; Jin, H.-J. Fluorescent silk fibroin nanoparticles prepared using a reverse microemulsion. Macromol. Res. 2008, 16, 604–608. [Google Scholar] [CrossRef]
  89. Wongpinyochit, T.; Uhlmann, P.; Urquhart, A.J.; Seib, F.P. PEGylated Silk Nanoparticles for Anticancer Drug Delivery. Biomacromolecules 2015, 16, 3712–3722. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Gholami, A.; Tavanai, H.; Moradi, A.R. Production of fibroin nanopowder through electrospraying. J. Nanopart. Res. 2011, 13, 2089–2098. [Google Scholar] [CrossRef]
  91. Numata, K.; Reagan, M.R.; Goldstein, R.H.; Rosenblatt, M.; Kaplan, D.L. Spider silk-based gene carriers for tumor cell-specific delivery. Bioconjug. Chem. 2011, 22, 1605–1610. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Meinel, L.; Kaplan, D.L. Silk constructs for delivery of musculoskeletal therapeutics. Adv. Drug Deliv. Rev. 2012, 64, 1111–1122. [Google Scholar] [CrossRef] [PubMed]
  93. Dong, Y.; Dong, P.; Huang, D.; Mei, L.; Xia, Y.; Wang, Z.; Pan, X.; Li, G.; Wu, C. Fabrication and characterization of silk fibroin-coated liposomes for ocular drug delivery. Eur. J. Pharm. Biopharm. 2015, 91, 82–90. [Google Scholar] [CrossRef] [PubMed]
  94. Li, A.B.; Kluge, J.A.; Guziewicz, N.A.; Omenetto, F.G.; Kaplan, D.L. Silk-based stabilization of biomacromolecules. J. Control. Release 2015, 219, 416–430. [Google Scholar] [CrossRef] [Green Version]
  95. Wang, X.; Wenk, E.; Matsumoto, A.; Meinel, L.; Li, C.; Kaplan, D.L. Silk microspheres for encapsulation and controlled release. J. Control. Release 2007, 117, 360–370. [Google Scholar] [CrossRef]
  96. Numata, K.; Subramanian, B.; Currie, H.A.; Kaplan, D.L. Bioengineered silk protein-based gene delivery systems. Biomaterials 2009, 30, 5775–5784. [Google Scholar] [CrossRef] [Green Version]
  97. Li, L.; Puhl, S.; Meinel, L.; Germershaus, O. Silk fibroin layer-by-layer microcapsules for localized gene delivery. Biomaterials 2014, 35, 7929–7939. [Google Scholar] [CrossRef]
  98. Numata, K.; Kaplan, D.L. Silk-based gene carriers with cell membrane destabilizing peptides. Biomacromolecules 2010, 11, 3189–3195. [Google Scholar] [CrossRef] [Green Version]
  99. Song, W.; Gregory, D.A.; Al-janabi, H.; Muthana, M.; Cai, Z.; Zhao, X. Magnetic-silk/polyethyleneimine core-shell nanoparticles for targeted gene delivery into human breast cancer cells. Int. J. Pharm. 2019, 555, 322–336. [Google Scholar] [CrossRef]
  100. Zhang, Y.-Q.; Ma, Y.; Xia, Y.-Y.; Shen, W.-D.; Mao, J.-P.; Zha, X.-M.; Shirai, K.; Kiguchi, K. Synthesis of silk fibroin-insulin bioconjugates and their characterization and activities in vivo. J. Biomed. Mater. Res. Part B 2006, 79B, 275–283. [Google Scholar] [CrossRef]
  101. Pritchard, E.M.; Dennis, P.B.; Omenetto, F.; Naik, R.R.; Kaplan, D.L. Review physical and chemical aspects of stabilization of compounds in silk. Biopolymers 2012, 97, 479–498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Liu, Y.; Zheng, Z.; Gong, H.; Liu, M.; Guo, S.; Li, G.; Wang, X.; Kaplan, D.L. DNA preservation in silk. Biomater. Sci. 2017, 5, 1279–1292. [Google Scholar] [CrossRef] [PubMed]
  103. Zhang, X.; Berghe, I.V.; Wyeth, P. Heat and moisture promoted deterioration of raw silk estimated by amino acid analysis. J. Cult. Herit. 2011, 12, 408–411. [Google Scholar] [CrossRef]
  104. Zong, X.H.; Zhou, P.; Shao, Z.Z.; Chen, S.M.; Chen, X.; Hu, B.W.; Deng, F.; Yao, W.H. Effect of pH and copper(II) on the conformation transitions of silk fibroin based on EPR, NMR, and Raman spectroscopy. Biochemistry 2004, 43, 11932–11941. [Google Scholar] [CrossRef]
  105. Xiao, S.; Wang, Z.; Ma, H.; Yang, H.; Xu, W. Effective removal of dyes from aqueous solution using ultrafine silk fibroin powder. Adv. Powder Technol. 2014, 25, 574–581. [Google Scholar] [CrossRef]
  106. Ciocci, M.; Cacciotti, I.; Seliktar, D.; Melino, S. Injectable silk fibroin hydrogels functionalized with microspheres as adult stem cells-carrier systems. Int. J. Biol. Macromol. 2018, 108, 960–971. [Google Scholar] [CrossRef] [PubMed]
  107. Moin, A.; Wani, S.U.D.; Osmani, R.A.; Lila, A.S.A.; Khafagy, E.-S.; Arab, H.H.; Gangadharappa, H.V.; Allam, A.N. Formulation, characterization, and cellular toxicity assessment of tamoxifen-loaded silk fibroin nanoparticles in breast cancer. Drug Deliv. 2021, 28, 1626–1636. [Google Scholar] [CrossRef]
  108. Zhou, Z.; Shi, Z.; Cai, X.; Zhang, S.; Corder, S.G.; Li, X.; Zhang, Y.; Zhang, G.; Chen, L.; Liu, M.; et al. The use of functionalized silk fibroin films as a platform for optical diraction-based sensing applications. Adv. Mater. 2017, 29, 1605471. [Google Scholar] [CrossRef]
  109. Wang, C.; Li, X.; Gao, E.; Jian, M.; Xia, K.; Wang, Q.; Xu, Z.; Ren, T.; Zhang, Y. Carbonized Silk Fabric for Ultrastretchable, Highly Sensitive, and Wearable Strain Sensors. Adv. Mater. 2016, 28, 6640–6648. [Google Scholar] [CrossRef]
  110. Allen, T.M.; Cullis, P.R. Drug delivery systems: Entering the mainstream. Science 2004, 303, 1818–1822. [Google Scholar] [CrossRef] [Green Version]
  111. Zhang, Y.; Chan, H.F.; Leong, K.W. Advanced materials and processing for drug delivery: The past and the future. Adv. Drug Deliv. Rev. 2013, 65, 104–120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. 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.; et al. Structures, mechanical properties and applications of silk fibroin materials. Prog. Polym. Sci. 2015, 46, 86–110. [Google Scholar] [CrossRef]
  113. Jastrzebska, K.; Kucharczyk, K.; Florczak, A.; Dams-kozlowska, H. Silk as an innovative biomaterial for cancer therapy. Rep. Pract. Oncol. Radiother. 2014, 20, 87–98. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Seib, F.P.; Pritchard, E.M.; Kaplan, D.L. Self-assembling doxorubicin silk hydrogels for the focal treatment of primary breast cancer. Adv. Funct. Mater. 2013, 23, 58–65. [Google Scholar] [CrossRef]
  115. Wu, H.; Liu, S.; Xiao, L.; Dong, C.; Lu, Q.; Kaplan, D.L. Injectable and pH-responsive silk nanofiber hydrogels for sustained anticancer drug delivery. ACS Appl. Mater. Interfaces 2016, 8, 17118–17126. [Google Scholar] [CrossRef]
  116. He, W.; Li, P.; Zhu, Y.; Liu, M.; Huang, X.; Qi, H. An injectable silk fibroin nanofiber hydrogel hybrid system for tumor upconversion luminescence imaging and photothermal therapy. New J. Chem. 2018, 43, 2213–2219. [Google Scholar] [CrossRef]
  117. Mottaghitalab, F.; Farokhi, M.; Shokrgozar, M.A.; Atyabi, F.; Hosseinkhani, H. Silk fibroin nanoparticle as a novel drug delivery system. J. Control. Release 2015, 206, 161–176. [Google Scholar] [CrossRef]
  118. Qu, J.; Liu, Y.; Yu, Y.; Li, J.; Luo, J.; Li, M. Silk fibroin nanoparticles prepared by electrospray as controlled release carriers of cisplatin. Mater. Sci. Eng. C 2014, 44, 166–174. [Google Scholar] [CrossRef]
  119. Chithrani, B.D.; Chan, W.C. Elucidating the mechanism of cellular uptake and removal of protein-coated gold nanoparticles of different sizes and shapes. Nano Lett. 2007, 7, 1542–1550. [Google Scholar] [CrossRef]
  120. Gupta, V.; Aseh, A.; Ríos, C.N.; Aggarwal, B.B.; Mathur, A.B. Fabrication and characterization of silkfibroin-derived curcumin nanoparticles for cancer therapy. Int. J. Nanomed. 2009, 4, 115–122. [Google Scholar] [CrossRef] [Green Version]
  121. Montalbán, M.G.; Coburn, J.M.; Lozano-Perez, A.A.; Cenis, J.L.; Villora, G.; Kaplan, D.L. Production of curcumin-loaded silk fibroin nanoparticles for cancer therapy. Nanomaterials 2018, 8, 126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Bayraktar, O.; Malay, Ö.; Özgarip, Y.; Batigün, A. Silk fibroin as a novel coating material for controlled release of theophylline. Eur. J. Pharm. Biopharm. 2005, 60, 373–381. [Google Scholar] [CrossRef] [PubMed]
  123. Mao, S.; Guo, C.; Shi, Y.; Li, L.C. Recent advances in polymeric microspheres for parenteral drug delivery—Part 1. Expert Opin. Drug Deliv. 2012, 9, 1161–1176. [Google Scholar] [CrossRef] [PubMed]
  124. Goudarzi, A.; Sadrnezhaad, S.K.; Johari, N. The Prominent Role of Fully-Controlled Surface Co-Modification Procedure Using Titanium Nanotubes and Silk Fibroin Nanofibers in the Performance Enhancement of Ti6Al4V Implants. Surf. Coat. Technol. 2021, 412, 127001. [Google Scholar] [CrossRef]
  125. Wenk, E.; Wandrey, A.J.; Merkle, H.P.; Meinel, L. Silk Fibroin Spheres as a Platform for Controlled Drug Delivery. J. Control. Release 2008, 132, 26–34. [Google Scholar] [CrossRef] [PubMed]
  126. Uebersax, L.; Mattotti, M.; Papaloïzos, M.; Merkle, H.P.; Gander, B.; Meinel, L. Silk Fibroin Matrices for the Controlled Release of Nerve Growth Factor (NGF). Biomaterials 2007, 28, 4449–4460. [Google Scholar] [CrossRef] [PubMed]
  127. Ma, Y.; Canup, B.S.B.; Tong, X.; Dai, F.; Xiao, B. Multi-Responsive Silk Fibroin-Based Nanoparticles for Drug Delivery. Front. Chem. 2020, 8, 585077. [Google Scholar] [CrossRef] [PubMed]
  128. Hines, D.J.; Kaplan, D.L. Mechanisms of controlled release from silk fibroin films. Biomacromolecules 2011, 12, 804–812. [Google Scholar] [CrossRef] [Green Version]
  129. Pritchard, E.M.; Szybala, C.; Boison, D.; Kaplan, D.L. Silk fibroin encapsulated powder reservoirs for sustained release of adenosine. J. Control. Release 2010, 144, 159–167. [Google Scholar] [CrossRef] [Green Version]
  130. Megeeda, Z.; Haidera, M.; Li, D.; O’Malley, B.W., Jr.; Cappello, J.; Ghandehari, H. In vitro and in vivo evaluation of recombinant silk-elastinlike hydrogels for cancer gene therapy. J. Control. Release 2004, 94, 433–445. [Google Scholar] [CrossRef]
  131. Hatefi, A.; Cappello, J.; Ghandehari, H. Adenoviral gene delivery to solid tumors by recombinant silk-elastinlike protein polymers. Pharm. Res. 2007, 24, 773–779. [Google Scholar] [CrossRef] [PubMed]
  132. Greish, K.; Araki, K.; Li, D.; O’Malley, B.W., Jr.; Dandu, R.; Frandsen, J.; Cappello, J.; Ghandehari, H. Silk-elastinlike protein polymer hydrogels for localized adenoviral gene therapy of head and neck tumors. Biomacromolecules 2009, 10, 2183–2188. [Google Scholar] [CrossRef] [PubMed]
  133. Huang, W.; Rollett, A.; Kaplan, D.L. Silk-elastin-like protein biomaterials for the controlled delivery of therapeutics. Expert Opin. Drug Deliv. 2015, 12, 779–791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Vergaro, V.; Scarlino, F.; Bellomo, C.; Rinaldi, R.; Vergara, D.; Maffia, M.; Baldassarre, F.; Giannelli, G.; Zhang, X.; Lvov, Y.M.; et al. Drugloaded polyelectrolyte microcapsules for sustained targeting of cancer cells. Adv. Drug Deliv. Rev. 2011, 63, 847–863. [Google Scholar] [CrossRef] [PubMed]
  135. Ng, S.L.; Such, G.K.; Johnston, A.P.R.; Antequera-Garcia, G.; Caruso, F. Controlled release of DNA from poly(vinylpyrrolidone) capsules using cleavable linkers. Biomaterials 2011, 32, 6277–6284. [Google Scholar] [CrossRef]
  136. Kakade, S.; Manickam, D.S.; Handa, H.; Mao, G.Z.; Oupicky, D. Transfection activity of layer-by-layer plasmid DNA/poly(ethylenimine) films deposited on PLGA microparticles. Int. J. Pharm. 2009, 365, 44–52. [Google Scholar] [CrossRef] [Green Version]
  137. Jewell, C.M.; Lynn, D.M. Multilayered polyelectrolyte assemblies as platforms for the delivery of DNA and other nucleic acid-based therapeutics. Adv. Drug Deliv. Rev. 2008, 60, 979–999. [Google Scholar] [CrossRef] [Green Version]
  138. She, Z.; Antipina, M.N.; Li, J.; Sukhorukov, G.B. Mechanism of protein release from polyelectrolyte multilayer microcapsules. Biomacromolecules 2010, 11, 1241–1247. [Google Scholar] [CrossRef]
  139. Lu, Q.; Zhang, B.; Li, M.; Zuo, B.; Kaplan, D.L.; Huang, Y.; Zhu, H. Degradation mechanism and control of silk fibroin. Biomacromolecules 2011, 12, 1080–1086. [Google Scholar] [CrossRef] [Green Version]
  140. Farokhi, M.; Mottaghitalab, F.; Fatahi, Y.; Khademhosseini, A.; Kaplan, D.L. Overview of Silk Fibroin Use in Wound Dressings. Trends Biotechnol. 2018, 36, 907–922. [Google Scholar] [CrossRef]
  141. Niemiec, S.M.; Louiselle, A.E.; Hilton, S.A.; Dewberry, L.C.; Zhang, L.; Azeltine, M.; Xu, J.; Singh, S.; Sakthivel, T.S.; Seal, S.; et al. Nanosilk Increases the Strength of Diabetic Skin and Delivers CNP-miR146a to Improve Wound Healing. Front. Immunol. 2020, 11, 590285. [Google Scholar] [CrossRef] [PubMed]
  142. Galateanu, B.; Hudita, A.; Zaharia, C.; Bunea, M.-C.; Vasile, E.; Buga, M.-R.; Costache, M. Silk-Based Hydrogels for Biomedical Applications. In Cellulose-Based Superabsorbent Hydrogels; Polymers and Polymeric Composites: A Reference Series; Springer: Cham, Switzerland, 2019; pp. 1791–1817. [Google Scholar]
  143. Pritchard, E.M.; Kaplan, D.L. Silk fibroin biomaterials for controlled release drug delivery. Expert Opin. Drug Deliv. 2011, 8, 797–811. [Google Scholar] [CrossRef] [PubMed]
  144. Yu, R.; Yang, Y.; He, J.; Li, M.; Guo, B. Novel supramolecular self-healing silk fibroin-based hydrogel via host–guest interaction as wound dressing to enhance wound healing. Chem. Eng. J. 2021, 417, 128278. [Google Scholar] [CrossRef]
  145. Mottaghitalab, F.; Kiani, M.; Farokhi, M.; Kundu, S.C.; Reis, R.L.; Gholami, M.; Bardania, H.; Dinarvand, R.; Geramifar, P.; Beiki, D. Targeted delivery system based on gemcitabine-loaded silk fibroin nanoparticles for lung cancer therapy. ACS Appl. Mater. Interfaces 2017, 9, 31600–31611. [Google Scholar] [CrossRef]
  146. Mathur, A.B.; Gupta, V. Silk fibroin-derived nanoparticles for biomedical applications. Nanomedicine 2010, 5, 807–820. [Google Scholar] [CrossRef]
  147. Zhang, Y.-Q.; Wang, Y.-J.; Wang, H.-Y.; Zhu, L.; Zhou, Z.-Z. Highly efficient processing of silk fibroin nanoparticle-l-asparaginase bioconjugates and their characterization as a drug delivery system. Soft Matter 2011, 7, 9728–9736. [Google Scholar] [CrossRef]
  148. Zhang, Y.-Q.; Shen, W.-D.; Gu, R.-A.; Zhu, J.; Xue, R.-Y. Amperometric biosensor for uric acid based on uricase-immobilized silk fibroin membrane. Anal. Chim. Acta 1998, 369, 123–128. [Google Scholar] [CrossRef]
  149. Votyakova, T.V.; Reynolds, I.J. Detection of hydrogen peroxide with Amplex Red: Interference by NADH and reduced glutathione auto-oxidation. Arch. Biochem. Biophys. 2004, 431, 138–144. [Google Scholar] [CrossRef]
  150. Schneider, A.; Wang, X.Y.; Kaplan, D.L.; Garlick, J.A.; Egles, C. Biofunctionalized electrospun silk mats as a topical bioactive dressing for accelerated wound healing. Acta Biomater. 2009, 5, 2570–2578. [Google Scholar] [CrossRef] [Green Version]
  151. Gil, E.S.; Panilaitis, B.; Bellas, E.; Kaplan, D.L. Functionalized silk biomaterials for wound healing. Adv. Healthc. Mater. 2013, 2, 206–217. [Google Scholar] [CrossRef] [Green Version]
  152. Woong, D.; Kim, S.H.; Kim, H.H.; Lee, K.H.; Ki, C.S.; Park, Y.H. Multi-biofunction of antimicrobial peptide-immobilized silk fibroin nanofiber membrane: Implications for wound healing. Acta Biomater. 2016, 39, 146–155. [Google Scholar]
  153. Li, X.; Liu, Y.; Zhang, J.; You, R.; Qu, J.; Li, M. Functionalized silk fibroin dressing with topical bioactive insulin release for accelerated chronic wound healing. Mater. Sci. Eng. C 2017, 72, 394–404. [Google Scholar] [CrossRef] [PubMed]
  154. Ju, H.W.; Lee, O.J.; Moon, B.M.; Sheikh, F.A.; Lee, J.M.; Kim, J.H.; Park, H.J.; Kim, D.W.; Lee, M.C.; Soo, H.K. Silk fibroin based hydrogel for regeneration of burn induced wounds. Tissue Eng. Regener. Med. 2014, 11, 203–210. [Google Scholar] [CrossRef]
  155. Vasconcelos, A.; Gomes, A.C.; Cavaco-Paulo, A. Novel silk fibroin/elastin wound dressings. Acta Biomater. 2012, 8, 3049–3060. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Zhang, W.; Chen, L.; Chen, J.; Wang, L.; Gui, X.; Ran, J.; Xu, G.; Zhao, H.; Zeng, M.; Ji, J.; et al. Silk fibroin biomaterial shows safe and effective wound healing in animal models and a randomized controlled clinical trial. Adv. Healthc. Mater. 2017, 6, 1700121. [Google Scholar] [CrossRef]
  157. Ju, H.W.; Lee, O.J.; Lee, J.M.; Moon, B.M.; Park, H.J.; Park, Y.R.; Lee, M.C.; Kim, S.H.; Chao, J.R.; Ki, C.S.; et al. Wound healing effect of electrospun silk fibroin nanomatrix in burn-model. Int. J. Biol. Macromol. 2016, 85, 29–39. [Google Scholar] [CrossRef]
  158. Min, B.M.; Lee, G.; Kim, S.H.; Nam, Y.S.; Lee, T.S.; Park, W.H. Electrospinning of silkfibroin nanofibers and its effect on the adhesion and spreading of normalhuman keratinocytes and fibroblasts in vitro. Biomaterials 2014, 8, 1289–1297. [Google Scholar]
  159. Roh, D.H.; Kang, S.Y.; Kim, J.Y.; Kwon, Y.B.; Kweon, H.Y.; Lee, K.G.; Park, Y.H.; Baek, R.M.; Heo, C.Y.; Choe, J.; et al. Wound healing effect of silkfibroin/alginate-blended sponge in full thickness skin defect of rat. J. Mater. Sci. Mater. Med. 2006, 17, 547–552. [Google Scholar] [CrossRef]
Figure 1. For pharmaceutical and biomedical implementations, a versatile range of functional arrangements and chemical treatment for manufacturing a variety of silk fibroin (SF) formats are available. Image reproduced from reference [39] which was published under a CC BY license.
Figure 1. For pharmaceutical and biomedical implementations, a versatile range of functional arrangements and chemical treatment for manufacturing a variety of silk fibroin (SF) formats are available. Image reproduced from reference [39] which was published under a CC BY license.
Ijms 23 14421 g001
Figure 2. Processing of SF.
Figure 2. Processing of SF.
Ijms 23 14421 g002
Figure 3. SF microcapsules of plasmid DNA are packaged. (A) A diagram depicting the initialization of pDNA before and after it has been loaded. Preloading: pDNA was adsorbed onto bPEI25 functionalized PS particles; SF was mounted on the pDNA-coated particles using LbL; the SF was stabilised, and the core was removed. PS particles were coated with bPEI25 after loading; SF was assembled LbL onto bPEI25-coated PS particles; the center was removed after pDNA was adsorbed onto the bPEI25eSF casing after another bPEI25 coating. (B) Photos obtained with a fluorescence microscope of 4 mm SF microcapsules that had been pre- or post-loaded with pDNA. FITC (green) was used to mark SF, and Cy5 was used to label pDNA (red). 10 mm scale bar. Reprinted with permission from reference [97].
Figure 3. SF microcapsules of plasmid DNA are packaged. (A) A diagram depicting the initialization of pDNA before and after it has been loaded. Preloading: pDNA was adsorbed onto bPEI25 functionalized PS particles; SF was mounted on the pDNA-coated particles using LbL; the SF was stabilised, and the core was removed. PS particles were coated with bPEI25 after loading; SF was assembled LbL onto bPEI25-coated PS particles; the center was removed after pDNA was adsorbed onto the bPEI25eSF casing after another bPEI25 coating. (B) Photos obtained with a fluorescence microscope of 4 mm SF microcapsules that had been pre- or post-loaded with pDNA. FITC (green) was used to mark SF, and Cy5 was used to label pDNA (red). 10 mm scale bar. Reprinted with permission from reference [97].
Ijms 23 14421 g003
Figure 4. The results of various loading methods and measures of pDNA loaded SF micro capsules on release activity and cell capsule interactions are depicted in this diagram. (A) Preloading: pDNAebPEI complexes are released from SF capsules through diffusion. After charging, desorption from its membrane of SF capsules releases complexes. (B) For the same total amount of pDNA, varying sizes of gene carriers have various transmission densities on the cell surface, affecting cell viability and transfection. Reprinted with permission from reference [97].
Figure 4. The results of various loading methods and measures of pDNA loaded SF micro capsules on release activity and cell capsule interactions are depicted in this diagram. (A) Preloading: pDNAebPEI complexes are released from SF capsules through diffusion. After charging, desorption from its membrane of SF capsules releases complexes. (B) For the same total amount of pDNA, varying sizes of gene carriers have various transmission densities on the cell surface, affecting cell viability and transfection. Reprinted with permission from reference [97].
Ijms 23 14421 g004
Table 1. Preparation method of silk fibroin (SF) micro- and nanoparticles.
Table 1. Preparation method of silk fibroin (SF) micro- and nanoparticles.
S. N.Preparation TechniqueAdvantagesDisadvantagesParticle SizeRef.
01Freeze dryingPorous particlesTemperature dependent490–940 nm [14]
02Self-assemblySimple and safe technique
Avoidance of toxic solvents
Prevent intermolecular100–200 nm [30]
03PVA Blending methodTime and energy efficient
No use of organic solvent
PVA filtrate300–400 nm [81]
04Salting outEconomical technique
The drug can be encapsulated at the time of particle formation
Salting out agents filtrate500 nm–2 µm [84]
05Microfluidic methodsRapid technique
Mild operation procedures
Controlled particle size
Complex process150–300 nm [86]
06EmulsificationControllable particle size
Low-cost method
Residual surfactant170 nm [88]
07DesolvationSimple and quick method
Small particle size
Reproduceable technique
Easy to amassed; low drug load35–170 nm [89]
08ElectrosprayingHigh-purity particles
Very good monodispersity
Requires post handling to make insolubility of SF59–80 nm[90]
Table 2. SF-based drug delivery systems.
Table 2. SF-based drug delivery systems.
Form of Drug
Delivery System
Linked APIOutcomeRef.
SF spongesErythromycinSustained drug release and extended antimicrobial effects against S. Aureus[3]
SF nanoparticlesCurcuminModified drug release pattern and increased
cellular uptake
[68]
Modified the release profileIbuprofenIncreased adhesion and tunable drug release[93]
SF filmsEpirubicinControlled drug release[94]
SF microspheresHorseradish peroxidase (HRP)Modified the release profile[95]
Table 3. SF-based formulations for gene delivery.
Table 3. SF-based formulations for gene delivery.
FormulationGeneCell lineOutcomeRef.
Bioengineered silk filmspDNA (GFP)Human HEK cellsBeneficial for refining together the transfection efficiency and cell viability[96]
SF layer-by-layer assembled
microcapsules
pDNA-Cy5NIH/3T3 fibroblastsUniting low cyto-toxicity and high transfection effectiveness[97]
Bioengineered silk–polylysine–ppTG1
nanoparticles
pDNAHuman HEK and
MDA-MB-435 cells
Improves transfection efficiency[98]
Magnetic-SF/polyethyleneimine
core-shell nanoparticles
c-Myc12 antisense
ODNs
MDA-MB-231 cellsMeaningfully advanced inhibition effect[99]
3D porous scaffoldAdenovirus Ad-BMP7Human BMSCsExtended term compatibility for growth factor[100]
Abbreviations: GFP = green fluorescent protein, ODN = oligodeoxynucleotides, BMP = bone morphogenic protein.
Table 4. Various functionalized SF for controlled or redirected medicines.
Table 4. Various functionalized SF for controlled or redirected medicines.
SF BiomaterialApplicationsReferences
SF nanofibersDrug delivery system[115,124]
SF spheresControlled drug delivery[81,125]
SF matricesDrug delivery and controlled release[126]
SF NanoparticlesDrug delivery[127]
Table 5. SF-based drug delivery systems for wound healing.
Table 5. SF-based drug delivery systems for wound healing.
SF BiomaterialApplicationsReferences
SF solutionsSkin wound repair[40,141]
SF hydrogelsWound healing drug delivery[142,143,144]
SF NanoparticlesDrug delivery[145,146,147]
SF biosensorsWound monitoring[148,149]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wani, S.U.D.; Zargar, M.I.; Masoodi, M.H.; Alshehri, S.; Alam, P.; Ghoneim, M.M.; Alshlowi, A.; Shivakumar, H.G.; Ali, M.; Shakeel, F. Silk Fibroin as an Efficient Biomaterial for Drug Delivery, Gene Therapy, and Wound Healing. Int. J. Mol. Sci. 2022, 23, 14421. https://doi.org/10.3390/ijms232214421

AMA Style

Wani SUD, Zargar MI, Masoodi MH, Alshehri S, Alam P, Ghoneim MM, Alshlowi A, Shivakumar HG, Ali M, Shakeel F. Silk Fibroin as an Efficient Biomaterial for Drug Delivery, Gene Therapy, and Wound Healing. International Journal of Molecular Sciences. 2022; 23(22):14421. https://doi.org/10.3390/ijms232214421

Chicago/Turabian Style

Wani, Shahid Ud Din, Mohammed Iqbal Zargar, Mubashir Hussain Masoodi, Sultan Alshehri, Prawez Alam, Mohammed M. Ghoneim, Areej Alshlowi, H. G. Shivakumar, Mohammad Ali, and Faiyaz Shakeel. 2022. "Silk Fibroin as an Efficient Biomaterial for Drug Delivery, Gene Therapy, and Wound Healing" International Journal of Molecular Sciences 23, no. 22: 14421. https://doi.org/10.3390/ijms232214421

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop