Next Article in Journal
Strigolactones Might Regulate Ovule Development after Fertilization in Xanthoceras sorbifolium
Next Article in Special Issue
Whey Protein Isolate as a Substrate to Design Calendula officinalis Flower Extract Controlled-Release Materials
Previous Article in Journal
Molecular Detection and Genetic Diversity of Cytomegaloviruses and Lymphocryptoviruses in Free-Roaming and Captive African Green Monkeys (Chlorocebus sabaeus)
Previous Article in Special Issue
Solanum lycopersicum (Tomato)-Derived Nanovesicles Accelerate Wound Healing by Eliciting the Migration of Keratinocytes and Fibroblasts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Natural Compounds and Biomimetic Engineering to Influence Fibroblast Behavior in Wound Healing

by
Charlotte E. Berry
1,
Camille Brenac
1,
Caroline E. Gonzalez
2,
Carter B. Kendig
1,
Thalia Le
1,
Nicholas An
1 and
Michelle F. Griffin
1,*
1
Hagey Laboratory for Pediatric Regenerative Medicine, Division of Plastic and Reconstructive Surgery, Department of Surgery, Stanford University School of Medicine, 257 Campus Drive West, Stanford, CA 94305, USA
2
Feinberg School of Medicine, Northwestern University, Chicago, IL 60611, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(6), 3274; https://doi.org/10.3390/ijms25063274
Submission received: 5 February 2024 / Revised: 27 February 2024 / Accepted: 11 March 2024 / Published: 14 March 2024

Abstract

:
Throughout history, natural products have played a significant role in wound healing. Fibroblasts, acting as primary cellular mediators in skin wound healing, exhibit behavioral responses to natural compounds that can enhance the wound healing process. Identifying bioactive natural compounds and understanding their impact on fibroblast behavior offers crucial translational opportunities in the realm of wound healing. Modern scientific techniques have enabled a detailed understanding of how naturally derived compounds modulate wound healing by influencing fibroblast behavior. Specific compounds known for their wound healing properties have been identified. Engineered biomimetic compounds replicating the natural wound microenvironment are designed to facilitate normal healing. Advanced delivery methods operating at micro- and nano-scales have been developed to effectively deliver these novel compounds through the stratum corneum. This review provides a comprehensive summary of the efficacy of natural compounds in influencing fibroblast behavior for promoting wound regeneration and repair. Additionally, it explores biomimetic engineering, where researchers draw inspiration from nature to create materials and devices mimicking physiological cues crucial for effective wound healing. The review concludes by describing novel delivery mechanisms aimed at enhancing the bioavailability of natural compounds. Innovative future strategies involve exploring fibroblast-influencing pathways, responsive biomaterials, smart dressings with real-time monitoring, and applications of stem cells. However, translating these findings to clinical settings faces challenges such as the limited validation of biomaterials in large animal models and logistical obstacles in industrial production. The integration of ancient remedies with modern approaches holds promise for achieving effective and scar-free wound healing.

1. Introduction

Wound healing comprises three distinct yet interconnected stages: inflammation, proliferation, and remodeling [1]. While the ultimate goal of wound repair is functional skin regeneration, in adults, the process often results in the formation of nonfunctional fibrotic tissue, commonly referred to as “scars” [2,3]. Pathological wound healing can manifest as underhealing (delayed repair) or overhealing. Excessive wound healing leads to hypertrophic/keloid scars, causing patients significant physical and psychological morbidity [1]. Cellular and molecular pathways, particularly those involving fibroblasts, are instrumental in mediating wound healing and scar formation.
Fibroblasts play a crucial role in scar formation and fibrosis, making them a focal point in research aimed at improving wound healing outcomes [4]. These cells, abundantly present in wounds, coordinate the healing process at every stage by producing regulatory molecules and recruiting immune cells (Figure 1) [5]. Therefore, the dysfunction of fibroblasts could lead to delayed wound closure or excessive fibrosis. Initially, wound healing involves coagulation and inflammation in response to a damaged extracellular matrix (ECM) and damaged resident cells. This first inflammatory phase includes the vasoconstriction of injured vessels and a complex sequence of reactions in response to blood contact, including plasma factor VII/VIIa binding to tissue factor 3, ultimately producing a fibrin and platelet clot [6].
Platelets, in addition to stimulating the coagulation cascade, secrete growth factors such as transforming growth factor-β (TGF-β) and platelet-derived growth factor (PDGF), activating the recruitment of fibroblasts and immune cells [5]. Platelets play an important role in enhancing blood vessel permeability through the secretion of histamine and serotonin, alongside nitric oxide production by endothelial cells, which facilitates vasodilation. Concurrently, platelet-derived signals, complement activation, and factors from the coagulation cascade trigger chemotaxis, prompting immune cell migration to the wound bed for the elimination of potential pathogens or foreign debris. Although neutrophils are not deemed necessary for wound healing, they generate inflammatory products vital for wound repair and undertake the phagocytosis of pathogens. Conversely, both skin-resident and monocyte-derived macrophages are pivotal for wound healing, emerging as the dominant cell population orchestrating the release of growth factors within the wound environment until fibroblast proliferation [7]. Fibroblasts, typically arriving at the wound site between days 5 and 7, alongside lymphocytes, release inflammatory cytokines and chemokines, thereby regulating the wound microenvironment [8].
During the proliferation phase, fibroblasts play a pivotal role in wound remodeling by secreting matrix metalloproteinases (MMPs) to break down the fibrin clot [9]. Fibroblast growth factor (FGF) and TGF-β induce fibroblast proliferation and granulation tissue formation, facilitating the production of ECM molecules for keratinocyte migration [10]. In the remodeling phase, fibroblasts remodel the ECM and differentiate into myofibroblasts, regulating mechanical wound contraction [11]. This transformation is governed by cytokines such as TGF-β1 and CXC motif chemokine ligand 8, along with ECM stiffness. Myofibroblasts, acquiring contractile properties, enhance the organization and tensile strength of the ECM. As the final stage of wound healing, the remodeling phase can last from weeks to years [4].
Given the fundamental role of fibroblasts in normal tissue repair, dysregulation in their behavior can lead to delayed wound closure or exaggerated fibrosis [5]. Conditions associated with altered fibroblast behavior include chronic ulcers in metabolic diseases, chronic infections, and hypertrophic scarring, particularly in the context of high mechanical force [11,12]. Mechanoreceptors transduce tension across wounds, influencing fibroblast migration, orientation, elongation, and proliferation [2,13,14,15,16]. Additionally, mechanical loading reduces myofibroblast apoptosis, contributing to hypertrophic scar formation [11].
In response to the demonstrated relationship between mechanical force and fibroblast behavior, various synthetic external tension off-loading devices and dressings have been developed for wound healing and scar management [11]. As cellular pathways associated with mechanotransduction have been identified, new avenues have been opened for targets of pharmaceutical drug development. Future therapies may target sites of mechanotransduction, like mechanosensitive calcium-dependent channels, Rho-GTPases, and the integrin-like kinase-phosphoinositide 3-kinase/protein kinase B pathway, which are activated by stretch and integrin stress [17,18,19]. Similarly, focal adhesion kinase (FAK), a non-receptor tyrosine kinase, is phosphorylated upon mechanical injury, initiating inflammatory signaling, fibroblast recruitment, and collagen production [20]. Downstream pathways, such as those associated with wnt, β-catenin, and yes-associated protein (YAP), alter transcription and initiate profibrotic processes [21]. There is significant heterogeneity in dermal fibroblast responses to wnt/β-catenin signaling pathway, which has previously been associated with the role of wound macrophages on hair follicle stem cell activation [22,23,24,25,26,27,28,29].
In parallel, natural compounds from diverse biological reservoirs have been investigated for their capacity to promote wound tissue regeneration and repair [30,31,32,33]. These natural compounds can be used directly as treatments or as drug carriers for the delivery of other therapeutics. Natural bioactive agents have recently received attention in wound management due to their efficacy and minimal side effects, though the history of natural compounds in the field of wound healing is rich and diverse. Although natural remedies have been used for thousands of years, recent advancements in skin delivery, the understanding of phytochemistry and biologic activity, and the characterization of the mechanism of action through modern scientific techniques, such as high-throughput transcriptomic and proteomic analyses, 3D organotypic cell culture models, live cell imaging, and integrative omics approaches, have allowed for a deeper appreciation of how their therapeutic effects are often mediated by fibroblast behavior [34,35,36,37,38,39]. For example, honey’s phenolic compound constituents have been shown to modify fibroblast behavior, enhancing wound healing outcomes [31,40]. Understanding how these compounds affect wound repair and fibroblast modulation is considered crucial [41]. This review aims to highlight natural compounds capable of modulating fibroblast behavior, consequently facilitating wound regeneration and repair. Additionally, it delves into the realm of biomimetic engineering, where researchers draw inspiration from nature to design materials and devices that replicate physiological cues essential for optimal wound repair (Figure 2).

2. Natural Compounds

2.1. Botanicals and Plant-Derived Products

The utilization of botanicals in the context of wound healing has a rich legacy within medicinal practices, with their therapeutic properties well documented in the ethnobotanical literature [31,42,43,44,45,46]. This reliance on natural remedies persists today, with an estimated 70–80% of the global population depending on plant-derived medicines for disease treatment [47]. Moreover, the prevalence of non-healing wounds, affecting approximately six million individuals worldwide, presents a significant healthcare challenge and an economic cost of USD 3 billion annually [45]. Given the costs and unfavorable side effects associated with conventional medicines, the quest for natural compounds with healing properties is a prominent focus of green chemistry [45,47].
Notable advantages of these natural remedies include their cost-effectiveness, wide availability, and reduced risk of adverse side effects [35]. Consequently, there has been a growing interest in understanding the intricate mechanisms underlying the therapeutic actions of herbal medicines in tissue regeneration and repair processes, with many attributing the reparative and regenerative properties of medicinal plants to their bioactive phytochemical constituents influencing mechanotransduction processes (Table 1. Natural compounds) [31,35,45,48]. Premarathna et al. tested 23 seaweed species that have historically been used medicinally, almost all of which contained bioactive compounds such as saponins, tannins, flavonoids, and alkaloids. In vitro cell migration assays found the Sargassum illicifolium and Haliimeda opuntin extracts to be the most efficacious, a result that was mirrored by expedited healing during the in vivo treatment of excisional wounds. This result was connected to the ability of these compounds to enhance fibroblast proliferation and increase the action of myofibroblasts [49].
Several botanicals have been recognized as potential regulators of fibroblast behavior and mechanotransduction pathways related specifically to TGF-β1. Sarıçiçek (Achillea biebersteinii Afan), a member of the Asteraceae family high in phenolic acids and flavonoids, widely utilized for wound healing in Mediterranean regions, inhibits scar formation by modulating the expression of TGF-β1 and basic fibroblast growth factor (bFGF) at both the gene and protein levels in murine embryonic fibroblasts [45,50]. These growth factors promote wound healing via fibroblast proliferation and granulation tissue formation [10]. Similarly, traditional Chinese herbs rich in saponins and flavonoids, such as Astragali Radix, Rehmanniae Radix, and Centella Asiatica, have been shown to modulate human skin fibroblast migration activities through the TGF-β1 pathway, ECM synthesis through the Suppressor of Mothers Against Decaptentaplegic (Smad) pathway, and cell motility through the Ras/MAPK (non-Smad) pathway [52,70].
Flavonoids such as quercetin and polysaccharides such as APS2-1 from Astragalus membranaceus upregulate the expression of TGF-β1, bFGF, and epidermal growth factor (EGF) [31,67]. Additionally, terpenoids like terpinolene and alpha phellandrene, along with alkaloids like taspine, are associated with enhanced fibroblast migration. Terpinolene and alpha phellandrene as well as curcuminoids and polyphenols such as tannins have demonstrated the capacity to enhance fibroblast proliferation, while taspine plays a crucial role in promoting the expression of keratinocyte growth factor (KGF), a member of the FGF family [31,58,63,64,65]. Furthermore, flavonoids including naringin, morin, and icariin have been shown to accelerate collagen deposition [46,86]. These properties of botanical plants and their derivatives highlight their potential significance as mechanotransduction modulators capable of augmenting tissue regeneration and repair.

2.2. Bacterial-Derived Products: Botulinum Toxin A

Established clinical applications of botulinum toxin are diverse, including wrinkle improvement, facial and body contouring, as well as treatment for migraines, hyperhidrosis, and cervical dystonia [87]. Derived from the anaerobic gram-positive bacteria Clostridium Botulinum, Botulinum toxin exhibits seven serotypes, with type A (BTA) being the most widely employed clinically [88]. BTA functions by blocking the presynaptic release of acetylcholine, leading to neuromuscular junction inhibition and temporary muscle paralysis lasting from two to six months. Recently, the compound’s paralytic capabilities have been aimed at alleviating tension during wound healing, consequently minimizing scar formation.
In the context of scars and wound healing, BTA, through its mechanism of reducing tension on the dermis caused by underlying musculature movement, shows promise in enhancing healing outcomes. Numerous studies support the efficacy of BTA injections in enhancing the aesthetic and functional aspects of surgical scars [89,90]. Gassner et al. observed substantial cosmetic improvements in facial excision scars following BTA treatment in both primates and humans [90], while Yue et al. demonstrated significant reductions in wound tension, improved functional outcome scores, and enhanced facial scar appearance with BTA injection [89]. Besides its mechanical impact, various studies have highlighted a direct or indirect influence of BTA on fibroblast proliferation and differentiation [89,91].
In vitro studies indicate that BTA may reduce TGF-β activity by inhibiting the expression of TGF-β1, as well as lowering the transcription and expression of other profibrotic cytokines like connective tissue growth factor (CTGF). TGF-β, associated with hypertrophic scars, stimulates fibroblasts, increases collagen deposition, and inhibits fibroblast apoptosis [92]. Consequently, BTA may inhibit scar hyperplasia and treat hypertrophic scars by reducing fibroblast proliferation [93,94]. Liu et al.’s study demonstrated a dose-dependent cellular reduction in wounds [92]. Zhou also revealed that BTA could reduce angiogenesis by inhibiting the expression of related cytokines, such as VEGF, in hypertrophic scars [95]. Additionally, BTA modulates the differentiation of fibroblasts into myofibroblasts, as seen in in vitro studies indicating a decrease in alpha-smooth muscle actin mRNA and protein levels in fibroblasts, resulting in reduced myofibroblast differentiation and wound contractility [92]. Therefore, beyond its mechanical off-loading properties, BTA is a valuable natural compound in wound healing and scar management due to its lesser-known effects on fibroblast behavior.

2.3. Animal-Derived Products

The therapeutic potential of compounds derived from insects and aquatic animals in wound healing are being explored more in depth [51,74]. Insects, historically used in folk medicine, provide royal jelly, beeswax, pollen, and propolis, which have been used to treat diabetes and arthritis [51,71]. Recent studies have shed light on the remarkable ability of these insect-derived medicinal products, such as honey, maggot excretions/secretions (ES) from Phaenicia sericata, and sericin from Bombyx mori silkworm cocoons, to modulate tissue regeneration and repair through mechanotransduction pathways [51,71,72,73]. Notably, honey and maggot ES have been found to stimulate fibroblast proliferation, with honey additionally promoting fibroblast migration [51,71,73]. On the other hand, sericin has demonstrated its efficacy in preventing scar tissue formation by decreasing the expression of TGF-β1 and TGF-β3 [96]. While the benefits of these insect-derived products are well established in the history of wound healing, recent efforts have focused on incorporating them in novel ways, such as in the development of natural rubber latex membranes infused with three types of propolis [97].
The marine ecosystem has also emerged as a valuable source of biologically active compounds conducive to tissue regeneration and repair [51,74]. Marine collagen and sea cucumbers exhibit the capacity to invoke keratinocyte and fibroblast migration, improve skin vascularity, and modulate mechanotransduction pathways [33,51,72,74]. Both marine collagen and sea cucumbers have shown the ability to promote fibroblast chemotaxis, while sea cucumbers can contribute to the breakdown of ECM proteins and indirectly induce fibroblast proliferation through the activation of T-lymphocytes, which secrete TGF-β [33,51,72,74]. Consequently, these animal-derived products have been recognized as significant regulators of mechanotransduction and fibroblast behavior.

2.4. Exogenous Growth Factors and Growth Factor-Rich Products

The wound healing process is orchestrated through the coordinated interplay of various cells, guided by a complex signaling network of growth factors (GF) [76]. Recent years have witnessed a growing body of evidence supporting the use of exogenous GFs and GF-rich products to enhance wound healing outcomes [72,77,98,99].
These studies have elucidated specific GFs that play critical roles in modulating mechanotransduction and fibroblast behavior in pathways crucial for tissue regeneration and repair. Specifically, TGF-β, FGF, and PDGF have emerged as key drivers of phenotypic changes in fibroblasts, leading to their transformation into myofibroblasts and promoting wound contraction [75,76]. TGF-β1 and TGF-β2 can stimulate the production and deposition of ECM, while bFGF inhibits TGF-β1/Smad-dependent pathways [77,79]. This regulation of ECM synthesis and degradation is achieved through collagen distribution, α-smooth muscle actin (α-SMA) expression, and TGF-β1 modulation [77,79]. TGF-β3, PDGF, insulin-like growth factor (IGF)-1, EGF, and neuregulin 1 (NRG1), an EGF family member, have all been validated as inducers of fibroblast proliferation [63,76,77,80,81]. PDGF, EGF, and NRG1 exhibit the ability to induce fibroblast chemotaxis to wound sites, while both PDGF and IGF-1 stimulate collagen synthesis, further enhancing the wound healing process [77,78,80,81].
In addition to these GFs, various GF-rich products, such as Platelet-Rich Plasma (PRP), Platelet-Rich Fibrin Matrix (PRFM), and Decellularized and Dehydrated Human Amniotic Membrane (DDHAM), have demonstrated potential in modulating mechanotransduction during wound healing due to their GF constituents [77,82,83,84,85]. PRP is rich in PDGF and TGF-β1, and both PRFM and DDHAM products stimulate the secretion of PDGF, TGF-β1, EGF, and FGF-2 [77,82,83,84]. These findings underscore the diverse array of exogenous GFs and GF-rich products that hold promise in influencing mechanotransduction pathways integral to enhancing tissue regeneration and repair outcomes.

3. Biomimetic Engineering of Biomaterials

Biomimetic engineering involves the development of technologies inspired by natural structures and biological phenomena. These methods are applied to numerous fields including art, architecture, engineering, and medicine. Well-known examples in the industrial world include burr-inspired Velcro, bird-inspired airplanes, and sharkskin-inspired Olympic swimsuits [100]. In the medical field, drug-delivering nanoparticles (NPs) mimic vesicles and liposomes [100,101]. Additionally, biomimetic engineering plays a promising role in tissue engineering, where scaffolds mimic tissue structures and/or cell–cell interactions. Biomimetic applications have been demonstrated to improve bone, cardiac, nerve, hair, and skin tissue regeneration, among others [102,103,104,105].
Recent advancements in bioengineering have focused on the development of biomimetic dressings that closely resemble the natural skin environment, aiming to enhance wound healing [106,107]. Conventional dressings often require frequent re-application, potentially hindering re-epithelialization and prolonging inflammation [107]. Biomimetic dressings, designed to mimic the ECM of healthy skin, seek to replicate the skin’s inherent regenerative mechanisms [108,109,110,111,112,113]. The ECM plays a crucial role in influencing fibroblast proliferation, differentiation, migration, and adhesion [113,114,115,116,117]. While biomimetic dressings also exhibit the potential for anti-microbial activity, drug delivery, and temperature/humidity regulation [107,118], this review focuses on how they specifically regulate fibroblast activity to reduce fibrosis by reproducing cell–ECM interactions in healing wounds [113,114,119].
Electrospun nanofibers are the preferred method to mimic the natural wound microenvironment. However, other technologies discussed elsewhere in this review, such as hydrogels and NPs, similarly possess favorable biomimetic properties. Electrospun nanofibers exhibit a high surface area and modifiable topography, which make them favorable biomimetic substrates [115]. Electrospun nanofiber dressings can be composed of many different biocompatible synthetic or natural polymers. Indeed, over 100 distinct polymers have been successfully electrospun. These polymers must fulfill several criteria for successful electrospinning, including a high molecular weight, the ability to be dissolved in a low-dielectric-constant solvent, and optimal electrical conductivity [115]. In general, natural polymers such as chitosan, silk fibroin (SF), and collagen are selected because of their bioactivity contributing to a favorable wound microenvironment. However, natural polymers must typically be mixed with synthetic polymers such as poly(lactic acid), polyurethane (PU), or poly-caprolactone (PCL) to achieve the mechanical parameters necessary for electrospinning [115]. Each polymer exhibits unique properties, allowing the electrospun nanofiber meshes to be tailored to the goal of the scaffold. For example, human skin fibroblasts bind favorably to PU, poly(3-hydroxybuturate-co-3-hydroxyvalerate) (PHBV) has high elasticity and inhibits myofibroblast differentiation, PCL has high plasticity and ductility but lower cell adhesion rates, and SF is highly biocompatibility with low immunogenicity [119,120,121]. Variations in nanofiber polymers, as well as variations in mesh alignment, pore size, and fiber diameter impact fibroblast activity and scar formation. Wounds treated with nanofibers have been found to have a higher density of fibroblasts, as well as inflammatory and epithelial cells, and vascular tissue [122].
Perpendicularly aligned electrospun nanofibers offer topographical cues, relieving wound tension and preventing myofibroblast differentiation and scar formation [113]. In healing wounds, collagen fibers align linearly with wound tension. Both mechanical tension and topographical cues promote myofibroblast differentiation and proliferation, leading to pathological scar formation [114,117,123,124]. A dressing developed by Chen et al. using PCL and SF aligned perpendicularly disrupted pathologic topographical cues, resulting in minimal scar formation. Compared to parallel cues, perpendicular topographical cues reduced fibrosis-related proteins such as FN, TGFβ-1, and α-SMA, and the mechanosensitive protein YAP-1. This reduction indicates a decrease in the differentiation of resting fibroblasts into pro-fibrotic myofibroblasts [113].
Kim et al. demonstrated how a PHBV biodegradable electrospun nanofiber mesh, resembling the fibrous matrix of skin ECM, reduced mechanical stress on wounds. The mesh’s pores allowed cell migration, inducing fibroblast proliferation while reducing myofibroblast differentiation. This resulted in the ordered deposition of a new ECM and reduced scar formation in healed wounds [119].
Modifying the diameter of electrospun fibers alters the topography in healing wounds, influencing fibroblast migration and proliferation [125]. Gao et al. observed increased cell migration and expression of wound-healing genes on larger-diameter PU/Poly(lactide-co-caprolactone)/Polyethylene oxide nanofibers, while smaller diameters increased fibroblast proliferation. Although limited to in vitro experiments, these findings highlight the importance of topographical cues on fibroblast activity, offering an additional tunable variable in biomimetic dressing design.
Electrospun nanofibers, mimicking skin ECM cues, influence fibroblast proliferation, migration, and differentiation. Fibroblasts are crucial in early wound healing, but their dysregulation contributes to fibrosis [114]. Ideal biomimetic dressings express ECM-like cues that enhance fibroblast migration and proliferation while limiting myofibroblast differentiation, leading to faster healing with limited scarring. These studies showcase how tunable characteristics of biomimetic dressings, such as polymer type, fiber alignment, and fiber diameter, influence fibroblast phenotypes and fibrosis development in healing wounds.

4. Natural Compound Delivery Systems

While preclinical in vivo findings regarding the use of natural compounds to modulate fibroblast behavior are often promising, numerous compounds exhibit insufficient efficacy regarding wound healing in clinical trials. This discrepancy may stem from various issues, such as poor bioavailability within the wound environment, challenges in drug retention, and difficulties in penetrating the stratum corneum, among other obstacles related to drug delivery challenges in vivo [126]. Indeed, the journey of topical compounds to penetrate the wound site is limited by several physiological barriers, and passive drug diffusion into the skin is generally impeded by the outermost lipophilic layer of the skin, the stratum corneum. This tight layer is composed of lipids including triglycerides, cholesterol, and free fatty acids that are linked by covalent bonds with corneocytes and provide a defensive barrier to the external environment [127].
Furthermore, only moderately lipophilic compounds can perform transcutaneous passage [128]. To address these limitations, both natural and synthetic biomaterials have been developed to fuse with stratum corneum, disrupt the skin surface, or accumulate in the hair follicles to act as a long-term reservoir. These biomaterials must be delivered in a controlled manner to maintain therapeutic drug concentrations in order to target specific cells and cell components (e.g., fibroblasts) to achieve a robust local effect while minimizing potential side effects, (Table 2. Delivery systems). However, the enhancement of transdermal drug delivery could be achieved using penetration enhancers, skin disruption, or physical techniques. Innovative formulation designs using new technologies have emerged as effective, reliable methods for the delivery of natural compounds [129,130].
The incorporation of biomaterial-based scaffolds into the field of wound healing has gained prominence since the late 1990s, coinciding with the initial definition of “tissue engineering” [159]. These scaffolds, constructed from various synthetic and natural polymers, possess key features that are ideal for the wound healing process. Notably, they can be functionalized with various agents to enhance wound healing and serve as effective delivery systems on the macroscale (Figure 3) [160]. Rapidly, new therapeutic alternatives using engineering have been developed at the micro and nanoscale. In addition to the size difference, there is a range of physical and chemical properties that differ between those particles, such as melting point, dissolution, and solubility, allowing them to be used in different indications.
These drug delivery systems not only act as physical support or barriers for wounds but also actively release therapeutic compounds, improving bioavailability, optimizing pharmacokinetics, and reducing dosing frequency, thus enhancing and optimizing the healing process [161].

4.1. Micro Delivery Systems

Micro delivery systems are designed to overcome limitations relative to the ability of larger particles to penetrate deeply into tissues for maximum effectiveness at the cellular level [162]. However, these microscale systems can transport relatively large quantities of natural compounds due to their larger relative size and can be produced at a lower cost compared to nano-delivery systems.
Adhirajan et al. developed a gelatin microsphere conjugated to an MMP inhibitor and loaded with doxycycline into a collagen dressing. Their ability to reduce the MMP level was assessed in a rat wound model [163]. Other authors produced chitosan microparticles loaded with EGF and VEGF and suspended them in dextran hydrogel for application to burn wounds. This new therapy provided a beneficial effect in a burn wound model in rats; however, repeated applications were needed, indicating a potentially reduced duration effect [164].

4.2. Nano-Drug Delivery Systems

Following the advent of nanotechnology and recognizing the prevalence of biological phenomena at the nanoscale, numerous nano-drug delivery systems have been introduced to the realm of skin and wound regeneration [165,166].
Due to their microscopic size (1–100 nm), these biomaterials can exert unique effects at the molecular and cellular levels, with some demonstrating the ability to enter the cytoplasmic space or activate specific transport mechanisms [166]. Nano-drug delivery systems (NDDS) exhibit biocompatibility, biodegradability, non-toxicity, non-immunogenicity, and the ability to create a moist environment beneficial in the field of wound healing. Additionally, they possess inherent anti-bacterial properties, a high surface area-to-volume ratio, and can encapsulate both hydrophilic and hydrophobic drugs [166]. NDDS can integrate natural bioactive molecules, shielding them from degradation, delivering them to the application area, and sustaining their release. This capability enhances the therapeutic efficacy of drugs, reduces the frequency of administration, lowers overall costs, and improves patient compliance [167,168]. At present, nanomaterials stand out as the most promising delivery systems due to their distinct advantages, offering alterations in physical and chemical properties in the field of wound healing.
Bioactive molecules incorporated into NDDS are diverse, encompassing anti-microbial agents, growth factors, and genes, and can be either of natural or synthetic origin. Various macroscopic nanomaterials, including NPs but at a macroscopic size (e.g., electrospun nanofibers, nanosheets, nanoemulsions, carbon nanotubes-based, or graphene-based nanocomposites), and nano-sized biomaterials (e.g., NPs, ions, molecules, nucleic acids, functional peptides, proteins, oligosaccharides, or polysaccharides), have been documented [169]. As a result, natural compound-eluting nanomaterials or natural encapsulated nanobioactive molecules represent a particularly promising avenue for synergistically enhancing the healing process of wounds by acting on fibroblasts at the nano-scale [161].

4.3. Liposomes and Transfersomes

Liposomes, characterized as bilayer amphiphilic vesicles, have found extensive use in wound dressings and are emerging as promising nanocarriers [170]. In addition to safeguarding hydrophilic drugs like growth factors, hormones, and metabolites, they create a conducive moist environment for wound healing [134,167]. The topical formulation of insulin in the field of wound healing has encountered multiple problems, as it cannot be delivered safely and in a controlled manner. To overcome this limitation, Dawoud et al. devised insulin-loaded chitosan nanoparticle liposomes, effectively prolonging insulin release and demonstrating improved wound healing in human patients [135]. In another study, Xu et al. developed a unique liposome with an SF hydrogel core incorporating bFGF, enhancing the stability of bFGF and expediting wound healing by stimulating angiogenesis [171]. Additionally, usnic acid, a type of lichen metabolite first isolated in 1844, has been utilized in a liposome-embedded gelatin-based membrane by Nunes et al. and applied in a porcine burn wound model. This product was compared to two different standard dressings for burn wounds. The authors observed a highly fibroblastic component in the granulation tissue of the treated group on day 18, along with extensive condensation of thick collagen fibers and narrower interfibrillar spaces compared to those of the other groups, indicating induced fibroblast proliferation and synthesis [133].
Transfersomes, deformable liposomes featuring an edge activator such as sodium, have been designed to enhance their ability to traverse the stratum corneum and reach deeper layers of the epidermis [167]. Manconi et al. innovatively developed a self-assembling core–shell gellan transfersome, functioning as a nanohydrogel, to encapsulate baicalin for wound treatment in mice. Their findings demonstrated that these delivery systems could optimize skin drug deposition, with approximately 11% of baicalin retained in the entire skin, including 8% in the dermis, ultimately promoting enhanced skin regeneration. Baicalin-loaded formulations provided the highest Tumor Necrosis Factor alpha (TNF-α) and TGF-B1 inhibition in a mice wound model, which led to significant improvements in wound healing [136].

4.4. Lipid Nanoparticles

Introduced to overcome the limitation of liposomes, lipid NPs are also receiving increasing interest because they permit the administration of different drugs, such as growth factors or anti-bacterial peptides [145]. Lipid NPs are prepared with physiological lipids or lipid molecules without any organic solvent and can be either solid lipid NPs or nanostructured lipid carriers [167]. Gainza et al. produced solid lipid NPs and nanostructured lipids loaded with rhEGF, and tested them in vitro and in vivo in a murine full-thickness wound model, which resulted in expedited wound closure and re-epithelialization, primarily through increased fibroblast proliferation [167].

4.5. Polymeric Nanoparticles

Polymeric NPs serve as effective carriers for encapsulating natural agents, providing controlled release and protection from degradation by proteases in the wound microenvironment [138,157]. Poly(lactic-co-glycolic acid) (PLGA) has demonstrated its potential in promoting wound healing by expediting neovascularization. Recent developments include polymeric NPs incorporating anti-microbial agents. Chereddy et al. introduced a PLGA nanoparticle loaded with the defense peptide LL37, showcasing its modulation of wound healing, angiogenesis, and infection prevention. In the context of full-thickness wounds, the application of PLGA-LL37 NPs significantly accelerated the healing process by influencing collagen deposition, epithelialization, and neovascularization. By day 10, the extent of collagen deposition was higher in the treated group, and the deposited collagen was thicker and denser [156].
Manukumar et al. conducted a study on thymol-loaded chitosan silver NPs (AgNPs) against biofilm-associated proteins in methicillin-resistant Staphylococcus aureus (S. aureus), demonstrating excellent anti-bacterial activity and anti-biofilm properties without inducing toxicity. These findings underscore the potential application of this drug in the realm of wound healing [139].

4.6. Inorganic Nanoparticles

Inorganic natural NPs, encompassing metal NPs, mineral NPs, biogenic NPs, ceramic NPs, and more, are prevalent in various environments such as geological formations, soils, and biological systems [140,172]. Natural NPs exhibit an improved duration of action and biocompatibility compared with other nanoparticles or pure drugs [158,173,174]. Their utility extends to the field of wound healing, where they serve as intrinsic anti-bacterial agents, wound healing enhancers, and drug delivery systems [167]. AgNPs, a widely utilized inorganic variant, play a crucial role in preventing infections and significantly enhancing the healing of chronic wounds by disrupting the intracellular activity and cell membrane of bacteria [170].
To enhance full-thickness dermal wounds in mice, Chigurupati et al. developed water-soluble cerium oxide NPs (metal NPs). In vitro, the migration and proliferation of fibroblasts was increased compared with those in the control group, and the growth rates of both fibroblasts and keratinocytes were significantly increased in wounds treated by NPs. In vivo, the rate of wound closure was significantly higher in mice treated with NPs, and immunostaining showed a higher rate of α-SMA expression and thus an enhancement of myofibroblast differentiation in those wounds [141].
Ziv Polat et al. utilized maghemite NPs to stabilize thrombin, applied clinically for topical hemostasis and wound healing, in an incisional wound model [140]. Results indicated that the NP-treated group exhibited the fewest inflammatory cells, the least granulation tissue along the surgical scar, and the highest values of skin tensile strength compared with the free thrombin group. The evidence supported the notion that thrombin-bonded maghemite NPs significantly advanced the wound healing stage and achieved better overall healing quality [140].
Kim et al. used gold NPs coated on a hydrocolloid membrane (HCM) to treat wound injuries in a rat model. The authors showed a decreased time to wound closure as well as an increased expression of collagen and a decreased expression of collagen-degrading enzyme at the early stage of the wound skin repair process [143].

4.7. Nanofibrous Structures (Nanofibers/Nanoscaffold)

Nanofibers, composed of both natural and synthetic continuous polymer chains, serve as versatile materials, forming nanofibrous sheets or 3D scaffolds applied in tissue engineering [157]. These nanofibrous structures, designed to emulate the ECM, create conducive conditions for cell attachment and interactions with therapeutic agents. The nanofibrous structure facilitates the transfer of various therapeutic agents, including GFs, nucleic acids, and anti-microbial agents [157]. Electrospinning stands out as the most widely adopted technique for nanofiber production.
To address the poor solubility of andrographolide in aqueous media, Jia et al. incorporated mesoporous silica NPs loaded with andrographolide into a PLGA matrix. Unlike lipid NPs previously used for andrographolide encapsulation and incorporation into sponge scaffolds for topical applications, the use of silica NPs was preferred due to their high drug encapsulation capacity, and PLGA was chosen for its good biocompatibility, adjustable mechanical properties, and tunable degradation rate. The PLGA/Andro-MSNs demonstrated an accelerated wound healing rate with increased epidermal cell adhesion and a reduction in the inflammation process compared with other samples, showcasing the effects of sustained andrographolide release [146]. The incorporation of mesoporous silica NPs into PLGA also enhanced the hydrophilicity of the nanofibrous membranes [146].
Shan et al. evaluated the effectiveness of an SF/gelatin nanofibrous dressing loaded with astragaloside IV on deep partial-thickness burn wounds [147]. This dressing not only promoted cell adhesion and proliferation with good biocompatibility in vitro but also demonstrated efficacy in wound healing and reduced scar formation in vivo in mice [147].

4.8. Nanohydrogel and Hydrogels Loaded with Nanoparticles

While hydrogels excel at effectively encapsulating, protecting, and sustaining the release of compounds, the integration of NPs into hydrogels has emerged, leveraging their benefits synergistically within a single system [175]. To enhance hydrogel adhesion to the skin and investigate the impact of phenytoin on wound healing, Cardoso et al. designed chitosan hydrogels loaded with nanoencapsulated phenytoin. In a porcine wound model treated with phenytoin nanocarriers, the authors observed a significant presence of fibroblasts and collagen deposition on day 6, indicating the proliferation and synthetic activity of fibroblasts [153].
Mahmoud et al. characterized and loaded gold NPs into thermosensitive hydrogels. Demonstrating the advantageous combination of the properties of gold NPs and hydrogel stability, the authors showcased excellent prolonged drug release behavior. In a murine wound model, the treated group exhibited enhanced skin re-epithelialization, granulation tissue, vascularization, and collagen deposition. Furthermore, in vitro studies highlighted the drug’s exceptional anti-bacterial activity and its impact on the expression of inflammatory and anti-inflammatory mediators [154].
Aly et al. formulated hydrogel loaded with polymeric NPs of simvastatin for topical application. Their study revealed that 81% of simvastatin was released into the skin after 24 h. Histopathological assessments confirmed a notable wound healing effect, characterized by the formation of a normal epithelial layer by day 11 [157].
Organic NPs have also been designed for use in wound healing. Shalaby et al. transformed collagen derived from tilapia fish scales and incorporated it into PRP gel, resulting in a novel nanomaterial. This demonstrated faster re-epithelialization and wound closure in vivo. Treatment with collagen NPs improved wound contraction and increased the number of myofibroblasts in the wound environment [158].
Nanohydrogels are often engineered to possess enhanced properties, including increased surface area, enabling them to encapsulate a wide variety of drugs with high compatibility and efficacy. This significantly affects skin regeneration due to their improved interaction with biological tissues and cells, including fibroblasts, at the nanoscale.
Dehkordi et al. utilized nanocrystalline cellulose human leukocyte antigens enriched with granulocyte-macrophage colony-stimulating factor (GM-CSF)-loaded chitosan to control the delivery of this growth factor to the wound site in rats. By day 13 of wound healing, the treated group exhibited significantly higher granulation tissue formation, attributed to fibroblast proliferation and differentiation. Treated wounds also displayed a lower inflammatory reaction, enhanced epithelialization, and a reduced wound closure time [150].

5. Future Directions for the Field

As natural compounds are introduced to the field of wound healing, future research should prioritize innovative and effective therapeutic strategies that exploit recently described signaling pathways, gene expression patterns, and interactions crucial in influencing fibroblast behavior. The current therapeutic landscape of biomaterials in wound healing, including many of the biomaterials previously discussed, often consists of a single therapeutic and lacks the dynamic, multi-component delivery necessary for effective wound healing. A new area of exploration lies in scaffolds that can respond to the dynamic wound microenvironment to optimize therapeutic outcomes [176].

5.1. Micro-Environment-Responsive Biomaterials

Responsive biomaterials designed for micro-environments exhibit adaptability to various endogenous biological/biochemical (i.e., reactive oxygen species (ROS), pH, enzymes, glucose, and glutathione), and exogenous physical conditions (temperature, mechanical forces, and pressure), or a combination mode to optimize therapeutic responses [176]. For instance, ROS-cleavable linkers, ROS-responsive polymers, and ROS-sensitive indicators can be used for the real-time monitoring of ROS and modulating redox balance in the wound environment [176,177]. Wound pH can vary depending on the stage of healing, infection status, and the presence of exudates. Under the influence of hypoxia and lactic acid stimulation, acute wounds are induced to become acidic [176]. In contrast, chronic wounds consistently maintain an alkaline pH within the range of 7.15–8.90 [178]. Biomaterials responding to pH could release drugs or bioactive agents as appropriate according to the characteristics of the wound [179,180]. Additionally enzymatic stimuli, such as MMPs, are involved in various aspects of wound healing, including ECM remodeling, cell migration, and angiogenesis [181]. Chronic wounds, particularly diabetic ulcers, are characterized by elevated glucose levels. Glucose-responsive biomaterials can be designed to release insulin or other therapeutic agents in response to high glucose concentrations [182]. However, endogenous physiological stimuli exhibit a semi-passive nature, creating a challenge for controlled dosage regulation.
Conversely, drug delivery dressings activated by exogenous stimuli can be modulated through mechanical force, temperature, light, ultrasound, magnetic, and electric fields, providing more predicted responses. Li et al. developed a thermo-sensitive mechanically active hydrogel that actively contracted the tissue surrounding the wound in response to body temperature in a rat model [183]. Histological analyses revealed that the hydrogel treatment mechanically contracted the wound, reduced inflammation, and improved wound healing quality.
Pressure-sensitive hydrogels with anti-bacterial properties can detect skin pressure and transmit electrical signals, making them ideal for monitoring and treating pressure ulcers [184]. Another innovative wound dressing material is based on moisture-adaptive contractile fibers. Dong et al. utilized moisture-adaptive contractile fibers fabricated through a simple wet spinning technique and knitted into textiles to create wound dressings that conform to the shape of the wound [185]. These fibers absorb wound tissue fluid, elongate to create tensile force, and, upon drying, contract to promote wound closure, accelerating healing and preventing infection [185].

5.2. Smart Dressings

While existing biomaterial systems typically monitor only one parameter at a time, future research aims to develop biomaterials equipped with multi-sensor capabilities capable of adapting and responding to various biomarkers within the dynamic wound microenvironment. This involves integrating real-time sensors into the wound microenvironment to enable the precise and controlled delivery of therapeutics, referred to as smart dressings. For example, Zhang et al. utilized an innovative integrated smart dressing consisting of three layers: a biomimetic nanofiber membrane, a gelatin methacryloyl and β-cyclodextrin ultraviolet crosslinked hydrogel, and a sensor chip [186]. The nanofiber membrane promotes cell migration and skin regeneration, while the hydrogel facilitates the controlled release of growth factors and cytokines for tissue remodeling, granulation tissue remodeling, and angiogenesis [186]. The sensor ensures accurate, stable measurements, and biocompatibility for optimal dose release [186]. Liang et al. demonstrated the use of pH/glucose dual-responsive metformin-released hydrogel dressings in a rat diabetic model [187]. These dressings respond to the physiological conditions of diabetic wounds, releasing metformin in both acidic and high-glucose environments. The hydrogel displayed superior self-healing, anti-oxidant, anti-inflammatory, and pro-angiogenic properties, effectively promoting the healing of chronic diabetic foot wounds in rats [187].
Smart dressings have immersive potential for clinical applications in telemedicine, artificial intelligence-based diagnosis, and personalized precision medicine for wound healing. However, the early stages of wound infection often exhibit subtle changes in biomarker levels that are susceptible to external noise, guiding inaccurate sensor responses. Therefore, future biomaterials and wearable smart dressings will likely work to optimize accuracy, sensitivity, and stability.

5.3. Stem Cells

While bone marrow mesenchymal stem cells have been extensively studied, there is a growing interest in exploring the potential of stem cells from fat tissue and hair follicles for regenerating wounds [188,189,190]. Natural compounds and biomaterials can serve as conductors to guide stem cell differentiation and functionality. The combined effects of stem cells and fibroblasts in improving skin healing and reducing scarring, especially with different cell lineages of mesenchymal stem cells (MSC), have been extensively explored. Vojtassák et al., 2006, conducted a study on chronic diabetic foot ulcers, combining autologous fibroblasts and MSCs on a biodegradable collagen membrane, resulting in decreased wound size, increased vascularity, and dermal thickness [191]. Yates et al. produced a similar result by experimenting on a full-thickness excisional wound model in mice to assess the effects of simultaneously applying MSC and fibroblasts in a collagen-tenascin-C matrix on wound healing and scarring [192]. The results confirmed that MSC and fibroblast co-transplantation increased the expression of collagen, fibronectin, and other matrix-related genes, and decreased the expression of pro-inflammatory cytokine. Moreover, MSCs can influence collagen organization, leading to a thinner, more flexible scar with a basket weave structure compared with the dense, disorganized scar tissue typically formed in control wounds. This improved structure is associated with greater tensile strength in the healed tissue [193]. Nonetheless, several challenges confront current cell-based wound therapies, including issues like immune rejection, complex treatment protocols, an incomplete understanding of the underlying molecular pathways, and the intricacies of managing cell viability and function.

6. Limitations

The integration of traditional healing agents with modern natural or synthetic biomaterials, such as nanofibers containing silver nanoparticles, aloe vera-loaded alginate hydrogels, propolis in dressing films, and hydrogel sheets containing honey, has been explored in various studies [194,195]. Although they have demonstrated efficacy in clinical trials, testing primarily involved the direct application of agents as topical solutions or in combination with gels and dressings [194], with a limited exploration of their effectiveness in conjunction with more modern drug delivery systems such as nanoparticles, biomaterials, and smart dressings.
As previously noted, nanofiber-based substrates have demonstrated the ability to enhance wound healing by targeting pathways influenced by fibroblasts. However, the efficacy of most nanofibrous materials in wound healing has primarily been studied in 2D cultures, which lack the complexities of the wound microenvironment observed in vivo 3/12/2024 9:24:00 AM. Since biologically significant signaling pathways, particularly those involving the interplay between adhesion and growth, function optimally when cells are organized spatially within 3D tissues, there has been a growing trend in current and future studies towards transitioning from 2D cell culture to 3D bioscaffolding [196,197]. The risk of over-proliferation and scar formation has already begun to be addressed by embedding nanofiber substrates with drugs [198] or cell components [199] that inhibit fibroblast differentiation into myofibroblasts. For instance, electrospun nanofibrous scaffolds, crafted from a combination of poly(ε-caprolactone), gelatin, and palmatine, were tested on wounds in rabbit ears and demonstrated promising anti-scarring properties beyond their influence on fibroblasts and collagen density [198]. Scar elevation index measurements and histological analyses revealed a significant reduction in hypertrophic scaring formation compared with that in the control group, and a shorter wound healing duration [198]. Nevertheless, there are still currently limited biomaterials investigated for anti-scar and skin regeneration benefits in clinical practice.
Furthermore, many experiments conducted on biomaterial applications in wound healing have utilized small animal models. Considering that the epidermis and dermis of rodent animals are thinner and primarily healed via contraction instead of regeneration like in humans, it is necessary to validate the results in large animal models and, ultimately, clinical trials with long-term follow-ups [176]. The expenses and logistical obstacles associated with the industrial production of biomaterials for clinical applications continue to pose a significant barrier [176]. The proteolytic nature of the wound environment necessitates the careful selection of a dressing tailored to the specific wound type, whether acute, chronic, ulcerative, or resulting from burns. This characteristic also presents challenges for conducting blinded or randomized studies [53]. Lastly, the costliness of current advanced wound care products restricts their accessibility in hospitals and clinics [53,54]. Addressing this issue requires enhanced collaboration among industry stakeholders, researchers, and healthcare providers.

7. Conclusions

Medicine has made strides in understanding the function of fibroblasts in wound healing, though this field remains in its infancy in translating that knowledge to the clinical environment. Traditional medicines and cultures worldwide have employed plant and animal-derived compounds with fibroblast activity, laying the groundwork for future therapies. Modern scientific approaches have allowed for a more thorough understanding of how these natural substances result in physical, chemical, and cellular signals that modify fibroblast behavior. Biomimetic materials and cutting-edge delivery systems harness these effects and allow for an interplay between therapeutic delivery and the dynamic wound microenvironment. Utilizing compounds both ancient and new, harnessing the power of fibroblast activity allows for therapies promoting rapid and scar-free wound healing.

Author Contributions

C.E.B., C.B. and M.F.G. contributed to the ideation and conceptualization of the presented manuscript. C.E.B., C.B., C.E.G., T.L., N.A. and C.B.K. contributed to the manuscript’s background research and writing. C.E.B. and C.B. contributed to figure creation. All authors assisted in editing and made substantial contributions to the written content included in the final manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Written informed consent for publication has been obtained from all necessary person(s).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

AgNPsSilver nanoparticles
bFGFBasic fibroblast growth factor
BTABotulinum toxin type A
CTGFConnective tissue growth factor
CXCL8C-X-C motif chemokine ligand 8
DDHAMDecellularized and dehydrated human amniotic membrane
ECM Extracellular matrix
EGFEpidermal growth factor
ESExcretions/secretions
FAKFocal adhesion kinase
FGFFibroblast growth factor
GFGrowth factors
GM-CSFGranulocyte-macrophage colony-stimulating factor
HCMHydrocolloid membrane
IGFInsulin-like growth factor
ILInterleukin
KGFKeratinocyte growth factor
MMPsMatrix metalloproteinases
MSCMesenchymal stem cells
NPsNanoparticles
NRG1Neuregulin 1
PCLPolycaprolactone
PDGFPlatelet-derived growth factor
PHBVPoly(3-hydroxybutyrate-co-3-hydroxyvalerate)
PLGAPoly(lactic-co-glycolic acid)
PRFMPlatelet-rich fibrin matrix
PRPPlatelet-rich plasma
PUPolyurethane
ROSReactive oxygen species
SFSilk fibroin
SmadSuppressor of mothers against decapentaplegic
TGF-βTransforming growth factor beta
TIMPsTissue inhibitors of metalloproteinases
TNF-αTumor necrosis factor alpha
YAPYes-associated protein
α-SMAα-smooth muscle actin

References

  1. Gurtner, G.C.; Werner, S.; Barrandon, Y.; Longaker, M.T. Wound repair and regeneration. Nature 2008, 453, 314–321. [Google Scholar] [CrossRef]
  2. Aarabi, S.; Bhatt, K.A.; Shi, Y.; Paterno, J.; Chang, E.I.; Loh, S.A.; Holmes, J.W.; Longaker, M.T.; Yee, H.; Gurtner, G.C. Mechanical load initiates hypertrophic scar formation through decreased cellular apoptosis. FASEB J. 2007, 21, 3250–3261. [Google Scholar] [CrossRef] [PubMed]
  3. Talbott, H.E.; Mascharak, S.; Griffin, M.; Wan, D.C.; Longaker, M.T. Wound healing, fibroblast heterogeneity, and fibrosis. Cell Stem Cell 2022, 29, 1161–1180. [Google Scholar] [CrossRef]
  4. Cialdai, F.; Risaliti, C.; Monici, M. Role of fibroblasts in wound healing and tissue remodeling on Earth and in space. Front. Bioeng. Biotechnol. 2022, 10, 958381. [Google Scholar] [CrossRef] [PubMed]
  5. Ahmadi, S.E.; Shabannezhad, A.; Kahrizi, A.; Akbar, A.; Safdari, S.M.; Hoseinnezhad, T.; Zahedi, M.; Sadeghi, S.; Mojarrad, M.G.; Safa, M. Tissue factor (coagulation factor III): A potential double-edge molecule to be targeted and re-targeted toward cancer. Biomark. Res. 2023, 11, 60. [Google Scholar] [CrossRef] [PubMed]
  6. Ellis, S.; Lin, E.J.; Tartar, D. Immunology of Wound Healing. Curr. Dermatol. Rep. 2018, 7, 350–358. [Google Scholar] [CrossRef] [PubMed]
  7. Langevin, H.M.; Nedergaard, M.; Howe, A.K. Cellular control of connective tissue matrix tension. J. Cell. Biochem. 2013, 114, 1714–1719. [Google Scholar] [CrossRef] [PubMed]
  8. Mirastschijski, U.; Schnabel, R.; Claes, J.; Schneider, W.; Ågren, M.S.; Haaksma, C.; Tomasek, J.J. Matrix metalloproteinase inhibition delays wound healing and blocks the latent transforming growth factor-β1-promoted myofibroblast formation and function. Wound Repair Regen. 2010, 18, 223–234. [Google Scholar] [CrossRef]
  9. Thulabandu, V.; Chen, D.; Atit, R.P. Dermal fibroblast in cutaneous development and healing. Wiley Interdiscip. Rev. Dev. Biol. 2018, 7, e307. [Google Scholar] [CrossRef]
  10. Berry, C.E.; Downer, M.; Morgan, A.G.; Griffin, M.; Liang, N.E.; Kameni, L.; Parker, J.B.L.; Guo, J.; Longaker, M.T.; Wan, D.C. The effects of mechanical force on fibroblast behavior in cutaneous injury. Front. Surg. 2023, 10, 1167067. [Google Scholar] [CrossRef]
  11. Carmichael, S.W. The tangled web of Langer’s lines. Clin. Anat. 2014, 27, 162–168. [Google Scholar] [CrossRef] [PubMed]
  12. Webb, K.; Hitchcock, R.W.; Smeal, R.M.; Li, W.; Gray, S.D.; Tresco, P.A. Cyclic strain increases fibroblast proliferation, matrix accumulation, and elastic modulus of fibroblast-seeded polyurethane constructs. J. Biomech. 2006, 39, 1136–1144. [Google Scholar] [CrossRef] [PubMed]
  13. Barnes, L.A.; Marshall, C.D.; Leavitt, T.; Hu, M.S.; Moore, A.L.; Gonzalez, J.G.; Longaker, M.T.; Gurtner, G.C. Mechanical Forces in Cutaneous Wound Healing: Emerging Therapies to Minimize Scar Formation. Adv. Wound Care 2018, 7, 47–56. [Google Scholar] [CrossRef] [PubMed]
  14. Bao, G.; Suresh, S. Cell and molecular mechanics of biological materials. Nat. Mater. 2003, 2, 715–725. [Google Scholar] [CrossRef] [PubMed]
  15. Miranti, C.K.; Brugge, J.S. Sensing the environment: A historical perspective on integrin signal transduction. Nature 2002, 4, E83–E90. [Google Scholar] [CrossRef] [PubMed]
  16. Mascharak, S.; Desjardins-Park, H.E.; Davitt, M.F.; Griffin, M.; Borrelli, M.R.; Moore, A.L.; Chen, K.; Duoto, B.; Chinta, M.; Foster, D.S.; et al. Preventing Engrailed-1 activation in fibroblasts yields wound regeneration without scarring. Science 2021, 372, eaba2374. [Google Scholar] [CrossRef] [PubMed]
  17. Li, G.; Li, Y.-Y.; Sun, J.-E.; Lin, W.-H.; Zhou, R.-X. ILK–PI3K/AKT pathway participates in cutaneous wound contraction by regulating fibroblast migration and differentiation to myofibroblast. Lab Investig. 2016, 96, 741–751. [Google Scholar] [CrossRef]
  18. Dupont, S.; Morsut, L.; Aragona, M.; Enzo, E.; Giulitti, S.; Cordenonsi, M.; Zanconato, F.; Le Digabel, J.; Forcato, M.; Bicciato, S.; et al. Role of YAP/TAZ in mechanotransduction. Nature 2011, 474, 179–183. [Google Scholar] [CrossRef]
  19. Kuehlmann, B.; Bonham, C.A.; Zucal, I.; Prantl, L.; Gurtner, G.C. Mechanotransduction in Wound Healing and Fibrosis. J. Clin. Med. 2020, 9, 1423. [Google Scholar] [CrossRef]
  20. Modelling and Targeting Mechanical Forces in Organ Fibrosis|Nature Reviews Bioengineering. Available online: https://www.nature.com/articles/s44222-023-00144-3 (accessed on 29 January 2024).
  21. Rognoni, E.; Gomez, C.; Pisco, A.O.; Rawlins, E.L.; Simons, B.D.; Watt, F.M.; Driskell, R.R. Inhibition of β-catenin signalling in dermal fibroblasts enhances hair follicle regeneration during wound healing. Development 2016, 143, 2522–2535. [Google Scholar] [CrossRef]
  22. Mastrogiannaki, M.; Lichtenberger, B.M.; Reimer, A.; Collins, C.A.; Driskell, R.R.; Watt, F.M. β-Catenin Stabilization in Skin Fibroblasts Causes Fibrotic Lesions by Preventing Adipocyte Differentiation of the Reticular Dermis. J. Investig. Dermatol. 2016, 136, 1130–1142. [Google Scholar] [CrossRef]
  23. Hamburg-Shields, E.; DiNuoscio, G.J.; Mullin, N.K.; Lafyatis, R.; Atit, R.P. Sustained β-catenin activity in dermal fibroblasts promotes fibrosis by up-regulating expression of extracellular matrix protein-coding genes. J. Pathol. 2014, 235, 686–697. [Google Scholar] [CrossRef] [PubMed]
  24. Han, J.; Lin, K.; Choo, H.; He, J.; Wang, X.; Wu, Y.; Chen, X. β-Catenin Signaling Evokes Hair Follicle Senescence by Accelerating the Differentiation of Hair Follicle Mesenchymal Progenitors. Front. Cell Dev. Biol. 2022, 10, 839519. [Google Scholar] [CrossRef]
  25. Rahmani, W.; Sinha, S.; Biernaskie, J. Immune modulation of hair follicle regeneration. NPJ Regen. Med. 2020, 5, 9. [Google Scholar] [CrossRef]
  26. Chen, H.; Wang, X.; Han, J.; Fan, Z.; Sadia, S.; Zhang, R.; Guo, Y.; Jiang, Y.; Wu, Y. AKT and its related molecular feature in aged mice skin. PLoS ONE 2017, 12, e0178969. [Google Scholar] [CrossRef]
  27. Chen, H.; Wang, X.; Chen, Y.; Han, J.; Kong, D.; Zhu, M.; Fu, X.; Wu, Y. Pten loss in Lgr5+ hair follicle stem cells promotes SCC development. Theranostics 2019, 9, 8321–8331. [Google Scholar] [CrossRef]
  28. Wang, X.; Chen, H.; Tian, R.; Zhang, Y.; Drutskaya, M.S.; Wang, C.; Ge, J.; Fan, Z.; Kong, D.; Wang, X.; et al. Macrophages induce AKT/β-catenin-dependent Lgr5+ stem cell activation and hair follicle regeneration through TNF. Nat. Commun. 2017, 8, 14091. [Google Scholar] [CrossRef] [PubMed]
  29. Scepankova, H.; Combarros-Fuertes, P.; Fresno, J.M.; Tornadijo, M.E.; Dias, M.S.; Pinto, C.A.; Saraiva, J.A.; Estevinho, L.M. Role of Honey in Advanced Wound Care. Molecules 2021, 26, 4784. [Google Scholar] [CrossRef]
  30. Falbo, F.; Spizzirri, U.G.; Restuccia, D.; Aiello, F. Natural Compounds and Biopolymers-Based Hydrogels Join Forces to Promote Wound Healing. Pharmaceutics 2023, 15, 271. [Google Scholar] [CrossRef]
  31. Al-Qattan, M.M.; Al-Shanawani, B.N.; Alshomer, F. Botulinum toxin type A: Implications in wound healing, facial cutaneous scarring, and cleft lip repair. Ann. Saudi Med. 2013, 33, 482–488. [Google Scholar] [CrossRef]
  32. Geahchan, S.; Baharlouei, P.; Rahman, A. Marine Collagen: A Promising Biomaterial for Wound Healing, Skin Anti-Aging, and Bone Regeneration. Mar. Drugs 2022, 20, 61. [Google Scholar] [CrossRef]
  33. Ryall, C.; Duarah, S.; Chen, S.; Yu, H.; Wen, J. Advancements in Skin Delivery of Natural Bioactive Products for Wound Management: A Brief Review of Two Decades. Pharmaceutics 2022, 14, 1072. [Google Scholar] [CrossRef]
  34. Vitale, S.; Colanero, S.; Placidi, M.; Di Emidio, G.; Tatone, C.; Amicarelli, F.; D’alessandro, A.M. Phytochemistry and Biological Activity of Medicinal Plants in Wound Healing: An Overview of Current Research. Molecules 2022, 27, 3566. [Google Scholar] [CrossRef]
  35. Ataide, J.A.; Cefali, L.C.; Croisfelt, F.M.; Shimojo, A.A.M.; Oliveira-Nascimento, L.; Mazzola, P.G. Natural actives for wound healing: A review. Phytother. Res. 2018, 32, 1664–1674. [Google Scholar] [CrossRef] [PubMed]
  36. Zhang, G.; Ji, P.; Xia, P.; Song, H.; Guo, Z.; Hu, X.; Guo, Y.; Yuan, X.; Song, Y.; Shen, R.; et al. Identification and targeting of cancer-associated fibroblast signature genes for prognosis and therapy in Cutaneous melanoma. Comput. Biol. Med. 2023, 167, 107597. [Google Scholar] [CrossRef] [PubMed]
  37. Chen, K.H.; Boettiger, A.N.; Moffitt, J.R.; Wang, S.; Zhuang, X. Spatially resolved, highly multiplexed RNA profiling in single cells. Science 2015, 348, aaa6090. [Google Scholar] [CrossRef] [PubMed]
  38. Woodley, J.P.; Lambert, D.W.; Asencio, I.O. Understanding Fibroblast Behavior in 3D Biomaterials. Tissue Eng. Part B Rev. 2022, 28, 569–578. [Google Scholar] [CrossRef]
  39. Combarros-Fuertes, P.; Fresno, J.M.; Estevinho, M.M.; Sousa-Pimenta, M.; Tornadijo, M.E.; Estevinho, L.M. Honey: Another Alternative in the Fight against Antibiotic-Resistant Bacteria? Antibiotics 2020, 9, 774. [Google Scholar] [CrossRef]
  40. Paudel, N.; Rai, M.; Adhikari, S.; Thapa, A.; Bharati, S.; Maharjan, B.; Shrestha, R.L.S.; Rav, K.; Singh, A.V. Green Extraction, Phytochemical Profiling, and Biological Evaluation of Dysphania ambrosioides: An In Silico and In Vitro Medicinal Investigation. J. Herbs Spices Med. Plants 2023, 30, 97–114. [Google Scholar] [CrossRef]
  41. Ponrasu, T.; Veerasubramanian, P.K.; Kannan, R.; Gopika, S.; Suguna, L.; Muthuvijayan, V. Morin incorporated polysaccharide–protein (psyllium–keratin) hydrogel scaffolds accelerate diabetic wound healing in Wistar rats. RSC Adv. 2018, 8, 2305–2314. [Google Scholar] [CrossRef]
  42. Hernandez-Hernandez, A.; Alarcon-Aguilar, F.; Almanza-Perez, J.; Nieto-Yañez, O.; Olivares-Sanchez, J.; Duran-Diaz, A.; Rodriguez-Monroy, M.; Canales-Martinez, M. Antimicrobial and anti-inflammatory activities, wound-healing effectiveness and chemical characterization of the latex of Jatropha neopauciflora Pax. J. Ethnopharmacol. 2017, 204, 1–7. [Google Scholar] [CrossRef]
  43. Tan, W.S.; Arulselvan, P.; Ng, S.-F.; Taib, C.N.M.; Sarian, M.N.; Fakurazi, S. Improvement of diabetic wound healing by topical application of Vicenin-2 hydrocolloid film on Sprague Dawley rats. BMC Complement. Altern. Med. 2019, 19, 20. [Google Scholar] [CrossRef]
  44. Mssillou, I.; Bakour, M.; Slighoua, M.; Laaroussi, H.; Saghrouchni, H.; Amrati, F.E.-Z.; Lyoussi, B.; Derwich, E. Investigation on wound healing effect of Mediterranean medicinal plants and some related phenolic compounds: A review. J. Ethnopharmacol. 2022, 298, 115663. [Google Scholar] [CrossRef]
  45. Trinh, X.-T.; Long, N.-V.; Van Anh, L.T.; Nga, P.T.; Giang, N.N.; Chien, P.N.; Nam, S.-Y.; Heo, C.-Y. A Comprehensive Review of Natural Compounds for Wound Healing: Targeting Bioactivity Perspective. Int. J. Mol. Sci. 2022, 23, 9573. [Google Scholar] [CrossRef]
  46. Natural Products and/or Isolated Compounds on Wound Healing. Available online: https://pubmed.ncbi.nlm.nih.gov/31186659/ (accessed on 22 February 2024).
  47. Wound Healing Properties of Selected Natural Products. Available online: https://pubmed.ncbi.nlm.nih.gov/30366427/ (accessed on 29 January 2024).
  48. Premarathna, A.D.; Ranahewa, T.; Wijesekera, S.; Harishchandra, D.; Karunathilake, K.; Waduge, R.N.; Wijesundara, R.; Jayasooriya, A.P.; Wijewardana, V.; Rajapakse, R. Preliminary screening of the aqueous extracts of twenty-three different seaweed species in Sri Lanka with in-vitro and in-vivo assays. Heliyon 2020, 6, e03918. [Google Scholar] [CrossRef]
  49. Hormozi, M.; Baharvand, P. Achillea biebersteinni Afan may inhibit scar formation: In vitro study. Mol. Genet. Genom. Med. 2019, 7, e640. [Google Scholar] [CrossRef] [PubMed]
  50. Zhang, Q.; Fong, C.C.; Yu, W.K.; Chen, Y.; Wei, F.; Koon, C.M.; Lau, K.M.; Leung, P.C.; Lau, C.B.S.; Fung, K.P.; et al. Herbal formula Astragali Radix and Rehmanniae Radix exerted wound healing effect on human skin fibroblast cell line Hs27 via the activation of transformation growth factor (TGF-β) pathway and promoting extracellular matrix (ECM) deposition. Phytomedicine 2012, 20, 9–16. [Google Scholar] [CrossRef] [PubMed]
  51. Ratcliffe, N.; Azambuja, P.; Mello, C.B. Recent advances in developing insect natural products as potential modern day medicines. Evid. Based Complement. Altern. Med. 2014, 2014, 904958. [Google Scholar] [CrossRef] [PubMed]
  52. Wu, F.; Bian, D.; Xia, Y.; Gong, Z.; Tan, Q.; Chen, J.; Dai, Y. Identification of Major Active Ingredients Responsible for Burn Wound Healing of Centella asiatica Herbs. Evid. Based Complement. Altern. Med. 2012, 2012, 848093. [Google Scholar] [CrossRef] [PubMed]
  53. El-Sherbeni, S.A.; Negm, W.A. The Wound Healing Effect of Botanicals and Pure Natural Substances Used in in Vivo Models. Inflammopharmacology 2023, 31, 755–772. [Google Scholar] [CrossRef] [PubMed]
  54. Singh, W.R.; Devi, H.S.; Kumawat, S.; Sadam, A.; Appukuttan, A.V.; Patel, M.R.; Lingaraju, M.C.; Singh, T.U.; Kumar, D. Angiogenic and MMPs Modulatory Effects of Icariin Improved Cutaneous Wound Healing in Rats. Eur. J. Pharmacol. 2019, 858, 172466. [Google Scholar] [CrossRef]
  55. Oryan, A.; Tabatabaei Naeini, A.; Moshiri, A.; Mohammadalipour, A.; Tabandeh, M.R. Modulation of Cutaneous Wound Healing by Silymarin in Rats. J. Wound Care 2012, 21, 457–464. [Google Scholar] [CrossRef]
  56. Li, W.; Kandhare, A.D.; Mukherjee, A.A.; Bodhankar, S.L. Hesperidin, a Plant Flavonoid Accelerated the Cutaneous Wound Healing in Streptozotocin-Induced Diabetic Rats: Role of TGF-ß/Smads and Ang-1/Tie-2 Signaling Pathways. EXCLI J. 2018, 17, 399–419. [Google Scholar] [CrossRef]
  57. Li, K.; Diao, Y.; Zhang, H.; Wang, S.; Zhang, Z.; Yu, B.; Huang, S.; Yang, H. Tannin Extracts from Immature Fruits of Terminalia Chebula Fructus Retz. Promote Cutaneous Wound Healing in Rats. BMC Complement. Altern. Med. 2011, 11, 86. [Google Scholar] [CrossRef]
  58. Porras-Reyes, B.H.; Lewis, W.H.; Roman, J.; Simchowitz, L.; Mustoe, T.A. Enhancement of wound healing by the alkaloid taspine defining mechanism of action. Exp. Biol. Med. 1993, 203, 18–25. [Google Scholar] [CrossRef]
  59. Salas-Oropeza, J.; Jimenez-Estrada, M.; Perez-Torres, A.; Castell-Rodriguez, A.E.; Becerril-Millan, R.; Rodriguez-Monroy, M.A.; Canales-Martinez, M.M. Wound Healing Activity of the Essential Oil of Bursera Morelensis, in Mice. Molecules 2020, 25, 1795. [Google Scholar] [CrossRef]
  60. Salas-Oropeza, J.; Jimenez-Estrada, M.; Perez-Torres, A.; Castell-Rodriguez, A.E.; Becerril-Millan, R.; Rodriguez-Monroy, M.A.; Jarquin-Yañez, K.; Canales-Martinez, M.M. Wound Healing Activity of α-Pinene and α-Phellandrene. Molecules 2021, 26, 2488. [Google Scholar] [CrossRef]
  61. Marchese, A.; Orhan, I.E.; Daglia, M.; Barbieri, R.; Di Lorenzo, A.; Nabavi, S.F.; Gortzi, O.; Izadi, M.; Nabavi, S.M. Antibacterial and Antifungal Activities of Thymol: A Brief Review of the Literature. Food Chem. 2016, 210, 402–414. [Google Scholar] [CrossRef] [PubMed]
  62. Riella, K.R.; Marinho, R.R.; Santos, J.S.; Pereira-Filho, R.N.; Cardoso, J.C.; Albuquerque-Junior, R.L.C.; Thomazzi, S.M. Anti-Inflammatory and Cicatrizing Activities of Thymol, a Monoterpene of the Essential Oil from Lippia Gracilis, in Rodents. J. Ethnopharmacol. 2012, 143, 656–663. [Google Scholar] [CrossRef] [PubMed]
  63. Wang, X.; Gao, Y.; Sun, X. Effect of Taspine hydrochloride on the repair of rat skin wounds by regulating keratinocyte growth factor signal. Bioengineered 2022, 13, 789–799. [Google Scholar] [CrossRef] [PubMed]
  64. Vaisberg, A.J.; Milla, M.; Planas, M.d.C.; Cordova, J.L.; de Agusti, E.R.; Ferreyra, R.; Mustiga, M.d.C.; Carlin, L.; Hammond, G.B. Taspine is the cicatrizant principle in Sangre de Grado Extracted from Croton lechleri*. Plant. Med. 1989, 55, 140–143. [Google Scholar] [CrossRef]
  65. Wu, S.; Tian, C.; Tu, Z.; Guo, J.; Xu, F.; Qin, W.; Chang, H.; Wang, Z.; Hu, T.; Sun, X.; et al. Protective effect of total flavonoids of Engelhardia roxburghiana Wall. leaves against radiation-induced intestinal injury in mice and its mechanism. J. Ethnopharmacol. 2023, 311, 116428. [Google Scholar] [CrossRef]
  66. Algahtani, M.S.; Ahmad, M.Z.; Shaikh, I.A.; Abdel-Wahab, B.A.; Nourein, I.H.; Ahmad, J. Thymoquinone Loaded Topical Nanoemulgel for Wound Healing: Formulation Design and In-Vivo Evaluation. Molecules 2021, 26, 3863. [Google Scholar] [CrossRef]
  67. Bonnard, M.; Martin, E.; Parrot, I. Wound Healing Potential of an Oleoresin Essential Oil Chemotype from Canarium schweinfurthii Engl. Molecules 2022, 27, 7966. [Google Scholar] [CrossRef]
  68. Yang, D.; Chen, H.; Wei, H.; Liu, A.; Wei, D.-X.; Chen, J. Hydrogel Wound Dressings Containing Bioactive Compounds Originated from Traditional Chinese Herbs: A Review. Smart Mater. Med. 2024, 5, 153–165. [Google Scholar] [CrossRef]
  69. Xu, N.; Wang, L.; Guan, J.; Tang, C.; He, N.; Zhang, W.; Fu, S. Wound Healing Effects of a Curcuma Zedoaria Polysaccharide with Platelet-Rich Plasma Exosomes Assembled on Chitosan/Silk Hydrogel Sponge in a Diabetic Rat Model. Int. J. Biol. Macromol. 2018, 117, 102–107. [Google Scholar] [CrossRef] [PubMed]
  70. Zhao, B.; Zhang, X.; Han, W.; Cheng, J.; Qin, Y. Wound healing effect of an Astragalus membranaceus polysaccharide and its mechanism. Mol. Med. Rep. 2017, 15, 4077–4083. [Google Scholar] [CrossRef] [PubMed]
  71. Prete, P.E. Growth effects of Phaenicia sericata larval extracts on fibroblasts: Mechanism for wound healing by maggot therapy. Life Sci. 1997, 60, 505–510. [Google Scholar] [CrossRef] [PubMed]
  72. Baptista-Silva, S.; Borges, S.; Costa-Pinto, A.R.; Costa, R.; Amorim, M.; Dias, J.R.; Ramos, Ó.; Alves, P.; Granja, P.L.; Soares, R.; et al. In Situ Forming Silk Sericin-Based Hydrogel: A Novel Wound Healing Biomaterial. ACS Biomater. Sci. Eng. 2021, 7, 1573–1586. [Google Scholar] [CrossRef] [PubMed]
  73. Zancanela, D.C.; Funari, C.S.; Herculano, R.D.; Mello, V.M.; Rodrigues, C.M.; Borges, F.A.; de Barros, N.R.; Marcos, C.M.; Almeida, A.M.F.; Guastaldi, A.C. Natural rubber latex membranes incorporated with three different types of propolis: Physical-chemistry and antimicrobial behaviours. Mater. Sci. Eng. C 2019, 97, 576–582. [Google Scholar] [CrossRef]
  74. El-Ashram, S.; El-Samad, L.M.; Basha, A.A.; El Wakil, A. Naturally-derived targeted therapy for wound healing: Beyond classical strategies. Pharmacol. Res. 2021, 170, 105749. [Google Scholar] [CrossRef]
  75. Pierce, G.F.; Mustoe, T.A.; Senior, R.M.; Reed, J.; Griffin, G.L.; Thomason, A.; Deuel, T.F. In vivo incisional wound healing augmented by platelet-derived growth factor and recombinant c-sis gene homodimeric proteins. J. Exp. Med. 1988, 167, 974–987. [Google Scholar] [CrossRef]
  76. Dinh, T.; Braunagel, S.; Rosenblum, B.I. Growth factors in wound healing. Clin. Podiatr. Med. Surg. 2015, 32, 109–119. [Google Scholar] [CrossRef] [PubMed]
  77. Larouche, J.; Sheoran, S.; Maruyama, K.; Martino, M.M. Immune Regulation of Skin Wound Healing: Mechanisms and Novel Therapeutic Targets. Adv. Wound Care 2018, 7, 209–231. [Google Scholar] [CrossRef] [PubMed]
  78. Sheikh, E.S.; Sheikh, E.S.; Fetterolf, D.E. Use of dehydrated human amniotic membrane allografts to promote healing in patients with refractory non healing wounds. Int. Wound J. 2014, 11, 711–717. [Google Scholar] [CrossRef] [PubMed]
  79. Todorović, V.; Peško, P.; Micev, M.; Bjelović, M.; Budeč, M.; Mićić, M.; Brašanac, D.; Ilić-Stojanović, O. Insulin-like growth factor-I in wound healing of rat skin. Regul. Pept. 2008, 150, 7–13. [Google Scholar] [CrossRef]
  80. Fotouhi, A.; Maleki, A.; Dolati, S.; Aghebati-Maleki, A.; Aghebati-Maleki, L. Platelet rich plasma, stromal vascular fraction and autologous conditioned serum in treatment of knee osteoarthritis. Biomed. Pharmacother. 2018, 104, 652–660. [Google Scholar] [CrossRef] [PubMed]
  81. Yoon, D.; Yoon, D.; Cha, H.-J.; Lee, J.-S.; Chun, W. Enhancement of wound healing efficiency mediated by artificial dermis functionalized with EGF or NRG1. Biomed. Mater. 2018, 13, 045007. [Google Scholar] [CrossRef]
  82. Smiell, J.M.; Treadwell, T.; Hahn, H.D.; Hermans, M.H. Real-world Experience with a Decellularized Dehydrated Human Amniotic Membrane Allograft. Wounds 2015, 27, 158–169. [Google Scholar]
  83. Li, M.; Huang, X.; Tang, T.-Y.D.; Mann, S. Synthetic cellularity based on non-lipid micro-compartments and protocell models. Curr. Opin. Chem. Biol. 2014, 22, 1–11. [Google Scholar] [CrossRef]
  84. Lin, J.; Sclafani, A.P. Platelet-Rich Plasma for Skin Rejuvenation and Tissue Fill. Facial Plast. Surg. Clin. N. Am. 2018, 26, 439–446. [Google Scholar] [CrossRef]
  85. Hwang, J.; Jeong, Y.; Park, J.M.; Lee, K.H.; Hong, J.W.; Choi, J. Biomimetics: Forecasting the future of science, engineering, and medicine. Int. J. Nanomed. 2015, 10, 5701–5713. [Google Scholar] [CrossRef]
  86. Rodrigues, F.B.; Duarte, G.S.; Marques, R.E.; Castelão, M.; Ferreira, J.; Sampaio, C.; Moore, A.P.; Costa, J. Botulinum toxin type A therapy for cervical dystonia. Emergencias 2020, 11, CD003633. [Google Scholar] [CrossRef]
  87. Carrero, L.M.K.; Ma, W.; Liu, H.; Yin, X.; Zhou, B. Botulinum toxin type A for the treatment and prevention of hypertrophic scars and keloids: Updated review. J. Cosmet. Dermatol. 2018, 18, 10–15. [Google Scholar] [CrossRef]
  88. A Systematic Review and Meta-Analysis: Botulinum Toxin A Effect on Postoperative Facial Scar Prevention. Available online: https://pubmed.ncbi.nlm.nih.gov/34609526/ (accessed on 29 January 2024).
  89. Gassner, H.G.; Brissett, A.E.; Otley, C.C.; Boahene, D.K.; Boggust, A.J.; Weaver, A.L.; Sherris, D.A. Botulinum toxin to improve facial wound healing: A prospective, blinded, placebo-controlled study. Mayo Clin. Proc. 2006, 81, 1023–1028. [Google Scholar] [CrossRef] [PubMed]
  90. An, M.K.; Cho, E.B.; Park, E.J.; Kim, L.S.; Kim, K.J. Appropriate Timing of Early Postoperative Botulinum Toxin Type A Injection for Thyroidectomy Scar Management: A Split-Scar Study. Plast. Reconstr. Surg. 2019, 144, 659e–668e. [Google Scholar] [CrossRef]
  91. Liu, D.-Q.; Li, X.-J.; Weng, X.-J. Effect of BTXA on Inhibiting Hypertrophic Scar Formation in a Rabbit Ear Model. Aesthetic Plast. Surg. 2017, 41, 721–728. [Google Scholar] [CrossRef]
  92. Jeong, H.S.; Lee, B.H.; Sung, H.M.; Park, S.Y.; Ahn, D.K.; Jung, M.S.; Suh, I.S. Effect of Botulinum Toxin Type A on Differentiation of Fibroblasts Derived from Scar Tissue. Plast. Reconstr. Surg. 2015, 136, 171e–178e. [Google Scholar] [CrossRef]
  93. Austin, E.; Koo, E.; Jagdeo, J. The Cellular Response of Keloids and Hypertrophic Scars to Botulinum Toxin A: A Comprehensive Literature Review. Dermatol. Surg. 2018, 44, 149–157. [Google Scholar] [CrossRef] [PubMed]
  94. Zhou, N.; Li, D.; Luo, Y.; Li, J.; Wang, Y. Effects of Botulinum Toxin Type A on Microvessels in Hypertrophic Scar Models on Rabbit Ears. BioMed Res. Int. 2020, 2020, 2170750. [Google Scholar] [CrossRef] [PubMed]
  95. Chandika, P.; Ko, S.-C.; Jung, W.-K. Marine-derived biological macromolecule-based biomaterials for wound healing and skin tissue regeneration. Int. J. Biol. Macromol. 2015, 77, 24–35. [Google Scholar] [CrossRef] [PubMed]
  96. Barrientos, S.; Stojadinovic, O.; Golinko, M.S.; Brem, H.; Tomic-Canic, M. Growth factors and cytokines in wound healing. Wound Repair Regen. 2008, 16, 585–601. [Google Scholar] [CrossRef] [PubMed]
  97. Dolati, S.; Yousefi, M.; Pishgahi, A.; Nourbakhsh, S.; Pourabbas, B.; Shakouri, S.K. Prospects for the application of growth factors in wound healing. Growth Factors 2020, 38, 25–34. [Google Scholar] [CrossRef]
  98. Walraven, M.; Akershoek, J.J.; Beelen, R.H.J.; Ulrich, M.M.W. In vitro cultured fetal fibroblasts have myofibroblast-associated characteristics and produce a fibrotic-like environment upon stimulation with TGF-β1: Is there a thin line between fetal scarless healing and fibrosis? Arch. Dermatol. Res. 2017, 309, 111–121. [Google Scholar] [CrossRef]
  99. Borena, B.M.; Martens, A.; Broeckx, S.Y.; Meyer, E.; Chiers, K.; Duchateau, L.; Spaas, J.H. Regenerative Skin Wound Healing in Mammals: State-of-the-Art on Growth Factor and Stem Cell Based Treatments. Cell. Physiol. Biochem. 2015, 36, 1–23. [Google Scholar] [CrossRef] [PubMed]
  100. Nabipour, M.; Mellati, A.; Abasi, M.; Barough, S.E.; Karimizade, A.; Banikarimi, P.; Hasanzadeh, E. Preparation of bilayer tissue-engineered polyurethane/poly-L-lactic acid nerve conduits and their in vitro characterization for use in peripheral nerve regeneration. J. Biol. Eng. 2024, 18, 16. [Google Scholar] [CrossRef]
  101. Tolmacheva, N.; Bhattacharyya, A.; Noh, I. Calcium Phosphate Biomaterials for 3D Bioprinting in Bone Tissue Engineering. Biomimetics 2024, 9, 95. [Google Scholar] [CrossRef]
  102. Alfarano, M.; Pastore, D.; Fogliano, V.; Schalkwijk, C.G.; Oliviero, T. The Effect of Sulforaphane on Glyoxalase I Expression and Activity in Peripheral Blood Mononuclear Cells. Nutrients 2018, 10, 1773. [Google Scholar] [CrossRef]
  103. Ma, J.; Qin, C.; Wu, J.; Zhuang, H.; Du, L.; Xu, J.; Wu, C. 3D multicellular micropatterning biomaterials for hair regeneration and vascularization. Mater. Horiz. 2023, 10, 3773–3784. [Google Scholar] [CrossRef]
  104. Singh, V.; Marimuthu, T.; Makatini, M.M.; Choonara, Y.E. Biopolymer-Based Wound Dressings with Biochemical Cues for Cell-Instructive Wound Repair. Polymers 2022, 14, 5371. [Google Scholar] [CrossRef]
  105. Hama, R.; Reinhardt, J.W.; Ulziibayar, A.; Watanabe, T.; Kelly, J.; Shinoka, T. Recent Tissue Engineering Approaches to Mimicking the Extracellular Matrix Structure for Skin Regeneration. Biomimetics 2023, 8, 130. [Google Scholar] [CrossRef]
  106. Salber, J.; Gräter, S.; Harwardt, M.; Hofmann, M.; Klee, D.; Dujic, J.; Jinghuan, H.; Ding, J.; Kippenberger, S.; Bernd, A.; et al. Influence of different ECM mimetic peptide sequences embedded in a nonfouling environment on the specific adhesion of human-skin keratinocytes and fibroblasts on deformable substrates. Small 2007, 3, 1023–1031. [Google Scholar] [CrossRef] [PubMed]
  107. Aghmiuni, A.I.; Keshel, S.H.; Sefat, F.; Khiyavi, A.A. Quince seed mucilage-based scaffold as a smart biological substrate to mimic mechanobiological behavior of skin and promote fibroblasts proliferation and h-ASCs differentiation into keratinocytes. Int. J. Biol. Macromol. 2020, 142, 668–679. [Google Scholar] [CrossRef] [PubMed]
  108. Oria, R.; Wiegand, T.; Escribano, J.; Elosegui-Artola, A.; Uriarte, J.J.; Moreno-Pulido, C.; Platzman, I.; Delcanale, P.; Albertazzi, L.; Navajas, D.; et al. Force loading explains spatial sensing of ligands by cells. Nature 2017, 552, 219–224. [Google Scholar] [CrossRef] [PubMed]
  109. Mobasseri, S.A.; Zijl, S.; Salameti, V.; Walko, G.; Stannard, A.; Garcia-Manyes, S.; Watt, F.M. Patterning of human epidermal stem cells on undulating elastomer substrates reflects differences in cell stiffness. Acta Biomater. 2019, 87, 256–264. [Google Scholar] [CrossRef] [PubMed]
  110. Hong, C.; Chung, H.; Lee, G.; Kim, C.; Kim, D.; Oh, S.J.; Kim, S.-H.; Lee, K. Hydrogel/Nanofiber Composite Wound Dressing Optimized for Skin Layer Regeneration through the Mechanotransduction-Based Microcellular Environment. ACS Appl. Bio Mater. 2023, 6, 1774–1786. [Google Scholar] [CrossRef] [PubMed]
  111. Chen, Z.; Xiao, L.; Hu, C.; Shen, Z.; Zhou, E.; Zhang, S.; Wang, Y. Aligned lovastatin-loaded electrospun nanofibers regulate collagen organization and reduce scar formation. Acta Biomater. 2023, 164, 240–252. [Google Scholar] [CrossRef]
  112. Tracy, L.E.; Minasian, R.A.; Caterson, E. Extracellular Matrix and Dermal Fibroblast Function in the Healing Wound. Adv. Wound Care 2016, 5, 119–136. [Google Scholar] [CrossRef]
  113. Xue, J.; Wu, T.; Dai, Y.; Xia, Y. Electrospinning and Electrospun Nanofibers: Methods, Materials, and Applications. Chem. Rev. 2019, 119, 5298–5415. [Google Scholar] [CrossRef]
  114. Sainio, A.; Järveläinen, H. Extracellular matrix-cell interactions: Focus on therapeutic applications. Cell. Signal. 2020, 66, 109487. [Google Scholar] [CrossRef]
  115. Majno, G.; Gabbiani, G.; Hirschel, B.J.; Ryan, G.B.; Statkov, P.R. Contraction of granulation tissue in vitro: Similarity to smooth muscle. Science 1971, 173, 548–550. [Google Scholar] [CrossRef]
  116. Zahedi, G.; Amraei, S.; Biglari, M. Simulation and optimization of ethanol amine production plant. Korean J. Chem. Eng. 2009, 26, 1504–1511. [Google Scholar] [CrossRef]
  117. Kim, H.S.; Chen, J.; Wu, L.-P.; Wu, J.; Xiang, H.; Leong, K.W.; Han, J. Prevention of excessive scar formation using nanofibrous meshes made of biodegradable elastomer poly(3-hydroxybutyrate-co-3-hydroxyvalerate). J. Tissue Eng. 2020, 11, 2041731420949332. [Google Scholar] [CrossRef]
  118. Gholipourmalekabadi, M.; Sapru, S.; Samadikuchaksaraei, A.; Reis, R.L.; Kaplan, D.L.; Kundu, S.C. Silk fibroin for skin injury repair: Where do things stand? Adv. Drug Deliv. Rev. 2020, 153, 28–53. [Google Scholar] [CrossRef]
  119. Ghasemi-Mobarakeh, L.; Prabhakaran, M.P.; Morshed, M.; Nasr-Esfahani, M.-H.; Ramakrishna, S. Electrospun poly(ɛ-caprolactone)/gelatin nanofibrous scaffolds for nerve tissue engineering. Biomaterials 2008, 29, 4532–4539. [Google Scholar] [CrossRef]
  120. Yang, M.; Yu, S.; Zhao, P.; Xie, L.; Lyu, G.; Yu, J. Fabrication of homogeneously-aligned nano-fillers encapsulated silk fibroin electrospun nanofibers for improved fibroblast attachment, epithelialization, and collagen depositions: In vitro and in vivo wound healing evaluation. J. Biomater. Sci. Polym. Ed. 2022, 33, 878–899. [Google Scholar] [CrossRef] [PubMed]
  121. Bugg, D.; Bretherton, R.C.; Kim, P.; Olszewski, E.; Nagle, A.; Schumacher, A.E.; Chu, N.; Gunaje, J.; DeForest, C.A.; Stevens, K.; et al. Infarct Collagen Topography Regulates Fibroblast Fate via p38-Yes-Associated Protein Transcriptional Enhanced Associate Domain Signals. Circ. Res. 2020, 127, 1306–1322. [Google Scholar] [CrossRef] [PubMed]
  122. Walraven, M.; Hinz, B. Therapeutic approaches to control tissue repair and fibrosis: Extracellular matrix as a game changer. Matrix Biol. 2018, 71–72, 205–224. [Google Scholar] [CrossRef] [PubMed]
  123. Gao, X.; Hou, T.; Wang, L.; Liu, Y.; Guo, J.; Zhang, L.; Yang, T.; Tang, W.; An, M.; Wen, M. Aligned electrospun fibers of different diameters for improving cell migration capacity. Colloids Surf. B Biointerfaces 2023, 234, 113674. [Google Scholar] [CrossRef] [PubMed]
  124. Rippon, M.; Ousey, K.; Cutting, K. Wound healing and hyper-hydration: A counterintuitive model. J. Wound Care 2016, 25, 68–75. [Google Scholar] [CrossRef] [PubMed]
  125. Barua, S.; Mitragotri, S. Challenges associated with Penetration of Nanoparticles across Cell and Tissue Barriers: A Review of Current Status and Future Prospects. Nano Today 2014, 9, 223–243. [Google Scholar] [CrossRef] [PubMed]
  126. Nguyen, H.X.; Nguyen, C.N. Microneedle-Mediated Transdermal Delivery of Biopharmaceuticals. Pharmaceutics 2023, 15, 277. [Google Scholar] [CrossRef]
  127. Peng, S.; Cheng, L.; Wu, Q.; Li, Y.; Ran, L.; Wang, W.; Huang, K.; Zhu, R.; Xue, S.; Zhou, C.; et al. A Modified Hyaluronic Acid–Based Dissolving Microneedle Loaded With Daphnetin Improved the Treatment of Psoriasis. Front. Bioeng. Biotechnol. 2022, 10, 900274. [Google Scholar] [CrossRef]
  128. Vasudevan, D.T.; Rajan, R. Effect of permeation enhancers on the penetration mechanism of transfersomal gel of ketoconazole. J. Adv. Pharm. Technol. Res. 2012, 3, 112–116. [Google Scholar] [CrossRef]
  129. Zhong, S.P.; Zhang, Y.Z.; Lim, C.T. Tissue scaffolds for skin wound healing and dermal reconstruction. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2010, 2, 510–525. [Google Scholar] [CrossRef]
  130. Negut, I.; Dorcioman, G.; Grumezescu, V. Scaffolds for Wound Healing Applications. Polymers 2020, 12, 2010. [Google Scholar] [CrossRef]
  131. Kianvash, N.; Bahador, A.; Pourhajibagher, M.; Ghafari, H.; Nikoui, V.; Rezayat, S.M.; Dehpour, A.R.; Partoazar, A. Evaluation of Propylene Glycol Nanoliposomes Containing Curcumin on Burn Wound Model in Rat: Biocompatibility, Wound Healing, and Anti-Bacterial Effects. Drug Deliv. Transl. Res. 2017, 7, 654–663. [Google Scholar] [CrossRef]
  132. Li, Z.; Liu, M.; Wang, H.; Du, S. Increased Cutaneous Wound Healing Effect of Biodegradable Liposomes Containing Madecassoside: Preparation Optimization, in Vitro Dermal Permeation, and in Vivo Bioevaluation. Int. J. Nanomed. 2016, 11, 2995–3007. [Google Scholar] [CrossRef]
  133. Gainza, G.; Pastor, M.; Aguirre, J.J.; Villullas, S.; Pedraz, J.L.; Hernandez, R.M.; Igartua, M. A novel strategy for the treatment of chronic wounds based on the topical administration of rhEGF-loaded lipid nanoparticles: In vitro bioactivity and in vivo effectiveness in healing-impaired db/db mice. J. Control. Release 2014, 185, 51–61. [Google Scholar] [CrossRef] [PubMed]
  134. Xu, H.; Chen, P.; ZhuGe, D.; Zhu, Q.; Jin, B.; Shen, B.; Xiao, J.; Zhao, Y. Liposomes with Silk Fibroin Hydrogel Core to Stabilize bFGF and Promote the Wound Healing of Mice with Deep Second-Degree Scald. Adv. Health Mater. 2017, 6, 1700344. [Google Scholar] [CrossRef] [PubMed]
  135. Nunes, P.S.; Rabelo, A.S.; de Souza, J.C.C.; Santana, B.V.; da Silva, T.M.M.; Serafini, M.R.; Menezes, P.d.P.; Lima, B.d.S.; Cardoso, J.C.; Alves, J.C.S.; et al. Gelatin-based membrane containing usnic acid-loaded liposome improves dermal burn healing in a porcine model. Int. J. Pharm. 2016, 513, 473–482. [Google Scholar] [CrossRef] [PubMed]
  136. Yang, F.; Cho, S.-W.; Son, S.M.; Bogatyrev, S.R.; Singh, D.; Green, J.J.; Mei, Y.; Park, S.; Bhang, S.H.; Kim, B.-S.; et al. Genetic engineering of human stem cells for enhanced angiogenesis using biodegradable polymeric nanoparticles. Proc. Natl. Acad. Sci. USA 2010, 107, 3317–3322. [Google Scholar] [CrossRef] [PubMed]
  137. Choi, J.U.; Lee, S.W.; Pangeni, R.; Byun, Y.; Yoon, I.-S.; Park, J.W. Preparation and in Vivo Evaluation of Cationic Elastic Liposomes Comprising Highly Skin-Permeable Growth Factors Combined with Hyaluronic Acid for Enhanced Diabetic Wound-Healing Therapy. Acta Biomater. 2017, 57, 197–215. [Google Scholar] [CrossRef]
  138. Chereddy, K.K.; Her, C.-H.; Comune, M.; Moia, C.; Lopes, A.; Porporato, P.E.; Vanacker, J.; Lam, M.C.; Steinstraesser, L.; Sonveaux, P.; et al. PLGA nanoparticles loaded with host defense peptide LL37 promote wound healing. J. Control. Release 2014, 194, 138–147. [Google Scholar] [CrossRef] [PubMed]
  139. Ziv-Polat, O.; Topaz, M.; Brosh, T.; Margel, S. Enhancement of incisional wound healing by thrombin conjugated iron oxide nanoparticles. Biomaterials 2010, 31, 741–747. [Google Scholar] [CrossRef] [PubMed]
  140. Gagliardi, A.; Paolino, D.; Costa, N.; Fresta, M.; Cosco, D. Zein- vs PLGA-based nanoparticles containing rutin: A comparative investigation. Mater. Sci. Eng. C 2020, 118, 111538. [Google Scholar] [CrossRef] [PubMed]
  141. Jia, Y.; Zhang, H.; Yang, S.; Xi, Z.; Tang, T.; Yin, R.; Zhang, W.; Pignatello, R.; Cenni, E.; Micieli, D.; et al. Electrospun PLGA membrane incorporated with andrographolide-loaded mesoporous silica nanoparticles for sustained antibacterial wound dressing. Nanomedicine 2018, 13, 2881–2899. [Google Scholar] [CrossRef]
  142. Ali, S.S.; Morsy, R.; El-Zawawy, N.A.; Fareed, M.F.; Bedaiwy, M.Y. Synthesized Zinc Peroxide Nanoparticles (ZnO2-NPs): A Novel Antimicrobial, Anti-Elastase, Anti-Keratinase, and Anti-Inflammatory Approach toward Polymicrobial Burn Wounds. Int. J. Nanomed. 2017, 12, 6059–6073. [Google Scholar] [CrossRef]
  143. Shan, Y.-H.; Peng, L.-H.; Liu, X.; Chen, X.; Xiong, J.; Gao, J.-Q. Silk fibroin/gelatin electrospun nanofibrous dressing functionalized with astragaloside IV induces healing and anti-scar effects on burn wound. Int. J. Pharm. 2015, 479, 291–301. [Google Scholar] [CrossRef]
  144. Sanchez, D.A.; Schairer, D.; Tuckman-Vernon, C.; Chouake, J.; Kutner, A.; Makdisi, J.; Friedman, J.M.; Nosanchuk, J.D.; Friedman, A.J. Amphotericin B Releasing Nanoparticle Topical Treatment of Candida Spp. in the Setting of a Burn Wound. Nanomed. Nanotechnol. Biol. Med. 2014, 10, 269–277. [Google Scholar] [CrossRef]
  145. Aly, U.F.; Aboutaleb, H.A.; Abdellatif, A.A.; Tolba, N.S. Formulation and evaluation of simvastatin polymeric nanoparticles loaded in hydrogel for optimum wound healing purpose. Drug Des. Dev. Ther. 2019, 13, 1567–1580. [Google Scholar] [CrossRef]
  146. Nunes, D.; Andrade, S.; Ramalho, M.J.; Loureiro, J.A.; Pereira, M.C. Polymeric Nanoparticles-Loaded Hydrogels for Biomedical Applications: A Systematic Review on In Vivo Findings. Polymers 2022, 14, 1010. [Google Scholar] [CrossRef]
  147. Cardoso, A.M.; de Oliveira, E.G.; Coradini, K.; Bruinsmann, F.A.; Aguirre, T.; Lorenzoni, R.; Barcelos, R.C.S.; Roversi, K.; Rossato, D.R.; Pohlmann, A.R.; et al. Chitosan hydrogels containing nanoencapsulated phenytoin for cutaneous use: Skin permeation/penetration and efficacy in wound healing. Mater. Sci. Eng. C 2019, 96, 205–217. [Google Scholar] [CrossRef]
  148. Xie, Z.; Paras, C.B.; Weng, H.; Punnakitikashem, P.; Su, L.-C.; Vu, K.; Tang, L.; Yang, J.; Nguyen, K.T. Dual Growth Factor Releasing Multi-Functional Nanofibers for Wound Healing. Acta Biomater. 2013, 9, 9351–9359. [Google Scholar] [CrossRef] [PubMed]
  149. Abadehie, F.S.; Dehkordi, A.H.; Zafari, M.; Bagheri, M.; Yousefiasl, S.; Pourmotabed, S.; Mahmoodnia, L.; Validi, M.; Ashrafizadeh, M.; Zare, E.N.; et al. Lawsone-Encapsulated Chitosan/Polyethylene Oxide Nanofibrous Mat as a Potential Antibacterial Biobased Wound Dressing. Eng. Regen. 2021, 2, 219–226. [Google Scholar] [CrossRef]
  150. Liu, J.; Jia, B.; Li, Z.; Li, W. Reactive oxygen species-responsive polymer drug delivery systems. Front. Bioeng. Biotechnol. 2023, 11, 1115603. [Google Scholar] [CrossRef]
  151. Lokhande, G.; Carrow, J.K.; Thakur, T.; Xavier, J.R.; Parani, M.; Bayless, K.J.; Gaharwar, A.K. Nanoengineered Injectable Hydrogels for Wound Healing Application. Acta Biomater. 2018, 70, 35–47. [Google Scholar] [CrossRef]
  152. Xi Loh, E.Y.; Fauzi, M.B.; Ng, M.H.; Ng, P.Y.; Ng, S.F.; Ariffin, H.; Mohd Amin, M.C.I. Cellular and Molecular Interaction of Human Dermal Fibroblasts with Bacterial Nanocellulose Composite Hydrogel for Tissue Regeneration. ACS Appl. Mater. Interfaces 2018, 10, 39532–39543. [Google Scholar] [CrossRef] [PubMed]
  153. Dehkordi, N.K.; Minaiyan, M.; Talebi, A.; Akbari, V.; Taheri, A. Nanocrystalline cellulose–hyaluronic acid composite enriched with GM-CSF loaded chitosan nanoparticles for enhanced wound healing. Biomed. Mater. 2019, 14, 035003. [Google Scholar] [CrossRef] [PubMed]
  154. Li, H.; Li, B.; Lv, D.; Li, W.; Lu, Y.; Luo, G. Biomaterials releasing drug responsively to promote wound healing via regulation of pathological microenvironment. Adv. Drug Deliv. Rev. 2023, 196, 114778. [Google Scholar] [CrossRef]
  155. Karavana, S.Y.; Gökçe, E.H.; Rençber, S.; Özbal, S.; Pekçetin, Ç.; Güneri, P.; Ertan, G. A New Approach to the Treatment of Recurrent Aphthous Stomatitis with Bioadhesive Gels Containing Cyclosporine A Solid Lipid Nanoparticles: In Vivo/in Vitro Examinations. Int. J. Nanomed. 2012, 7, 5693–5704. [Google Scholar] [CrossRef]
  156. Thanusha, A.V.; Dinda, A.K.; Koul, V. Evaluation of nano hydrogel composite based on gelatin/HA/CS suffused with Asiatic acid/ZnO and CuO nanoparticles for second degree burns. Mater. Sci. Eng. C 2018, 89, 378–386. [Google Scholar] [CrossRef]
  157. Manukumar, H.M.; Chandrasekhar, B.; Rakesh, K.P.; Ananda, A.P.; Nandhini, M.; Lalitha, P.; Sumathi, S.; Qin, H.-L.; Umesha, S. Novel T-C@AgNPs mediated biocidal mechanism against biofilm associated methicillin-resistant Staphylococcus aureus (Bap-MRSA) 090, cytotoxicity and its molecular docking studies. MedChemComm 2017, 8, 2181–2194. [Google Scholar] [CrossRef]
  158. Maheen, S.; Younis, H.; Khan, H.U.; Ali, S.; Rehman, A.U.; Ilyas, S.; Zafar, M.N.; Shafqat, S.R.; Kalam, A.; Al-Ghamdi, A.A. Enhanced Antifungal and Wound Healing Efficacy of Statistically Optimized, Physicochemically Evaluated Econazole-Triamcinolone Loaded Silica Nanoparticles. Front. Chem. 2022, 10, 836678. [Google Scholar] [CrossRef] [PubMed]
  159. Kim, H.S.; Sun, X.; Lee, J.-H.; Kim, H.-W.; Fu, X.; Leong, K.W. Advanced drug delivery systems and artificial skin grafts for skin wound healing. Adv. Drug Deliv. Rev. 2019, 146, 209–239. [Google Scholar] [CrossRef]
  160. Alexander, A.; Ajazuddin; Patel, R.J.; Saraf, S. Recent expansion of pharmaceutical nanotechnologies and targeting strategies in the field of phytopharmaceuticals for the delivery of herbal extracts and bioactives. J. Control. Release 2016, 241, 110–124. [Google Scholar] [CrossRef]
  161. Adhirajan, N.; Shanmugasundaram, N.; Shanmuganathan, S.; Babu, M. Functionally modified gelatin microspheres impregnated collagen scaffold as novel wound dressing to attenuate the proteases and bacterial growth. Eur. J. Pharm. Sci. 2009, 36, 235–245. [Google Scholar] [CrossRef]
  162. Ribeiro, M.P.; Morgado, P.I.; Miguel, S.A.P.; Coutinho, P.; Correia, I.J. Dextran-based hydrogel containing chitosan microparticles loaded with growth factors to be used in wound healing. Mater. Sci. Eng. C 2013, 33, 2958–2966. [Google Scholar] [CrossRef] [PubMed]
  163. Das, S.; Baker, A.B. Biomaterials and Nanotherapeutics for Enhancing Skin Wound Healing. Front. Bioeng. Biotechnol. 2016, 4, 82. [Google Scholar] [CrossRef] [PubMed]
  164. Garcia-Orue, I.; Gainza, G.; Garcia-Garcia, P.; Gutierrez, F.B.; Aguirre, J.J.; Hernandez, R.M.; Delgado, A.; Igartua, M. Composite nanofibrous membranes of PLGA/Aloe vera containing lipid nanoparticles for wound dressing applications. Int. J. Pharm. 2019, 556, 320–329. [Google Scholar] [CrossRef]
  165. Wang, W.; Lu, K.-J.; Yu, C.-H.; Huang, Q.-L.; Du, Y.-Z. Nano-drug delivery systems in wound treatment and skin regeneration. J. Nanobiotechnol. 2019, 17, 82. [Google Scholar] [CrossRef] [PubMed]
  166. Losi, P.; Briganti, E.; Magera, A.; Spiller, D.; Ristori, C.; Battolla, B.; Balderi, M.; Kull, S.; Balbarini, A.; Di Stefano, R.; et al. Tissue response to poly(ether)urethane-polydimethylsiloxane-fibrin composite scaffolds for controlled delivery of pro-angiogenic growth factors. Biomaterials 2010, 31, 5336–5344. [Google Scholar] [CrossRef] [PubMed]
  167. Chakrabarti, S.; Chattopadhyay, P.; Islam, J.; Ray, S.; Raju, P.S.; Mazumder, B. Aspects of Nanomaterials in Wound Healing. Curr. Drug Deliv. 2019, 16, 26–41. [Google Scholar] [CrossRef] [PubMed]
  168. Sandri, G.; Miele, D.; Faccendini, A.; Bonferoni, M.C.; Rossi, S.; Grisoli, P.; Taglietti, A.; Ruggeri, M.; Bruni, G.; Vigani, B.; et al. Chitosan/Glycosaminoglycan Scaffolds: The Role of Silver Nanoparticles to Control Microbial Infections in Wound Healing. Polymers 2019, 11, 1207. [Google Scholar] [CrossRef]
  169. Mitragotri, S.; Burke, P.A.; Langer, R. Overcoming the challenges in administering biopharmaceuticals: Formulation and delivery strategies. Nat. Rev. Drug Discov. 2014, 13, 655–672. [Google Scholar] [CrossRef] [PubMed]
  170. Dawoud, M.H.S.; Yassin, G.E.; Ghorab, D.M.; Morsi, N.M. Insulin Mucoadhesive Liposomal Gel for Wound Healing: A Formulation with Sustained Release and Extended Stability Using Quality by Design Approach. AAPS PharmSciTech 2019, 20, 158. [Google Scholar] [CrossRef]
  171. Manconi, M.; Manca, M.L.; Caddeo, C.; Valenti, D.; Cencetti, C.; Diez-Sales, O.; Nacher, A.; Mir-Palomo, S.; Terencio, M.C.; Demurtas, D.; et al. Nanodesign of new self-assembling core-shell gellan-transfersomes loading baicalin and in vivo evaluation of repair response in skin. Nanomedicine 2018, 14, 569–579. [Google Scholar] [CrossRef]
  172. Shalaby, M.; Hamouda, D.; Khedr, S.M.; Mostafa, H.M.; Saeed, H.; Ghareeb, A.Z. Nanoparticles fabricated from the bioactive tilapia scale collagen for wound healing: Experimental approach. PLoS ONE 2023, 18, e0282557. [Google Scholar] [CrossRef]
  173. Chigurupati, S.; Mughal, M.R.; Okun, E.; Das, S.; Kumar, A.; McCaffery, M.; Seal, S.; Mattson, M.P. Effects of cerium oxide nanoparticles on the growth of keratinocytes, fibroblasts and vascular endothelial cells in cutaneous wound healing. Biomaterials 2013, 34, 2194–2201. [Google Scholar] [CrossRef] [PubMed]
  174. Kim, J.E.; Lee, J.; Jang, M.; Kwak, M.H.; Go, J.; Kho, E.K.; Song, S.H.; Sung, J.E.; Lee, J.; Hwang, D.Y. Accelerated healing of cutaneous wounds using phytochemically stabilized gold nanoparticle deposited hydrocolloid membranes. Biomater. Sci. 2015, 3, 509–519. [Google Scholar] [CrossRef]
  175. Mahmoud, N.N.; Hikmat, S.; Abu Ghith, D.; Hajeer, M.; Hamadneh, L.; Qattan, D.; Khalil, E.A. Gold nanoparticles loaded into polymeric hydrogel for wound healing in rats: Effect of nanoparticles’ shape and surface modification. Int. J. Pharm. 2019, 565, 174–186. [Google Scholar] [CrossRef]
  176. Wallace, L.A.; Gwynne, L.; Jenkins, T. Challenges and opportunities of pH in chronic wounds. Ther. Deliv. 2019, 10, 719–735. [Google Scholar] [CrossRef]
  177. Miranda-Calderon, L.; Yus, C.; Landa, G.; Mendoza, G.; Arruebo, M.; Irusta, S. Pharmacokinetic control on the release of antimicrobial drugs from pH-responsive electrospun wound dressings. Int. J. Pharm. 2022, 624, 122003. [Google Scholar] [CrossRef]
  178. Pan, F.; Giovannini, G.; Zhang, S.; Altenried, S.; Zuber, F.; Chen, Q.; Boesel, L.F.; Ren, Q. pH-responsive silica nanoparticles for the treatment of skin wound infections. Acta Biomater. 2022, 145, 172–184. [Google Scholar] [CrossRef]
  179. Sobczak, M. Enzyme-Responsive Hydrogels as Potential Drug Delivery Systems—State of Knowledge and Future Prospects. Int. J. Mol. Sci. 2022, 23, 4421. [Google Scholar] [CrossRef]
  180. Mohanty, A.R.; Ravikumar, A.; Peppas, N.A. Recent advances in glucose-responsive insulin delivery systems: Novel hydrogels and future applications. Regen. Biomater. 2022, 9, rbac056. [Google Scholar] [CrossRef]
  181. Li, Z.; Huang, J.; Jiang, Y.; Liu, Y.; Qu, G.; Chen, K.; Zhao, Y.; Wang, P.; Wu, X.; Ren, J. Novel Temperature-Sensitive Hydrogel Promotes Wound Healing through YAP and MEK-Mediated Mechanosensitivity. Adv. Health Mater. 2022, 11, e2201878. [Google Scholar] [CrossRef] [PubMed]
  182. Li, D.; Fei, X.; Xu, L.; Wang, Y.; Tian, J.; Li, Y. Pressure-sensitive antibacterial hydrogel dressing for wound monitoring in bed ridden patients. J. Colloid Interface Sci. 2022, 627, 942–955. [Google Scholar] [CrossRef] [PubMed]
  183. Dong, L.; Zhang, W.; Ren, M.; Li, Y.; Wang, Y.; Zhou, Y.; Wu, Y.; Zhang, Z.; Di, J. Moisture-Adaptive Contractile Biopolymer-Derived Fibers for Wound Healing Promotion. Small 2023, 19, e2300589. [Google Scholar] [CrossRef] [PubMed]
  184. Zhang, Y.; Li, T.; Zhao, C.; Li, J.; Huang, R.; Zhang, Q.; Li, Y.; Li, X. An Integrated Smart Sensor Dressing for Real-Time Wound Microenvironment Monitoring and Promoting Angiogenesis and Wound Healing. Front. Cell Dev. Biol. 2021, 9, 701525. [Google Scholar] [CrossRef]
  185. Liang, Y.; Li, M.; Yang, Y.; Qiao, L.; Xu, H.; Guo, B. pH/Glucose Dual Responsive Metformin Release Hydrogel Dressings with Adhesion and Self-Healing via Dual-Dynamic Bonding for Athletic Diabetic Foot Wound Healing. ACS Nano 2022, 16, 3194–3207. [Google Scholar] [CrossRef]
  186. Wu, Y.; Chen, L.; Scott, P.G.; Tredget, E.E. Mesenchymal stem cells enhance wound healing through differentiation and angiogenesis. Stem Cells 2007, 25, 2648–2659. [Google Scholar] [CrossRef]
  187. Kim, W.-S.; Park, B.-S.; Sung, J.-H.; Yang, J.-M.; Park, S.-B.; Kwak, S.-J.; Park, J.-S. Wound healing effect of adipose-derived stem cells: A critical role of secretory factors on human dermal fibroblasts. J. Dermatol. Sci. 2007, 48, 15–24. [Google Scholar] [CrossRef]
  188. François, S.; Mouiseddine, M.; Mathieu, N.; Semont, A.; Monti, P.; Dudoignon, N.; Saché, A.; Boutarfa, A.; Thierry, D.; Gourmelon, P.; et al. Human mesenchymal stem cells favour healing of the cutaneous radiation syndrome in a xenogenic transplant model. Ann. Hematol. 2007, 86, 1–8. [Google Scholar] [CrossRef]
  189. Vojtaššák, J.; Danišovič, Ľ.; Kubeš, M.; Bakoš, D.; Jarábek, L.; Uličná, M.; Blaško, M. Autologous biograft and mesenchymal stem cells in treatment of the diabetic foot. Neuro Endocrinol. Lett. 2006, 27 (Suppl. S2), 134–137. [Google Scholar]
  190. Yates, C.C.; Rodrigues, M.; Nuschke, A.; Johnson, Z.I.; Whaley, D.; Stolz, D.; Newsome, J.; Wells, A. Multipotent stromal cells/mesenchymal stem cells and fibroblasts combine to minimize skin hypertrophic scarring. Stem Cell Res. Ther. 2017, 8, 193. [Google Scholar] [CrossRef] [PubMed]
  191. Qi, Y.; Jiang, D.; Sindrilaru, A.; Stegemann, A.; Schatz, S.; Treiber, N.; Rojewski, M.; Schrezenmeier, H.; Vander Beken, S.; Wlaschek, M.; et al. TSG-6 released from intradermally injected mesenchymal stem cells accelerates wound healing and reduces tissue fibrosis in murine full-thickness skin wounds. J. Investig. Dermatol. 2014, 134, 526–537. [Google Scholar] [CrossRef] [PubMed]
  192. Pereira, R.F.; Bártolo, P.J. Traditional Therapies for Skin Wound Healing. Adv. Wound Care 2016, 5, 208–229. [Google Scholar] [CrossRef] [PubMed]
  193. Dorai, A.A. Wound care with traditional, complementary and alternative medicine. Indian J. Plast. Surg. 2012, 45, 418–424. [Google Scholar] [CrossRef] [PubMed]
  194. Piraino, F.; Selimović, Š. A Current View of Functional Biomaterials for Wound Care, Molecular and Cellular Therapies. BioMed Res. Int. 2015, 2015, 403801. [Google Scholar] [CrossRef] [PubMed]
  195. Schneider, A.; Garlick, J.A.; Egles, C. Self-assembling peptide nanofiber scaffolds accelerate wound healing. PLoS ONE 2008, 3, e1410. [Google Scholar] [CrossRef] [PubMed]
  196. Jiang, Z.; Zhao, L.; He, F.; Tan, H.; Li, Y.; Tang, Y.; Duan, X.; Li, Y. Palmatine-loaded electrospun poly(ε-caprolactone)/gelatin nanofibrous scaffolds accelerate wound healing and inhibit hypertrophic scar formation in a rabbit ear model. J. Biomater. Appl. 2021, 35, 869–886. [Google Scholar] [CrossRef] [PubMed]
  197. Chen, L.; Cheng, L.; Wang, Z.; Zhang, J.; Mao, X.; Liu, Z.; Zhang, Y.; Cui, W.; Sun, X. Conditioned medium-electrospun fiber biomaterials for skin regeneration. Bioact. Mater. 2021, 6, 361–374. [Google Scholar] [CrossRef]
  198. Fadilah, N.I.M.; Maarof, M.; Motta, A.; Tabata, Y.; Fauzi, M.B. The Discovery and Development of Natural-Based Biomaterials with Demonstrated Wound Healing Properties: A Reliable Approach in Clinical Trials. Biomedicines 2022, 10, 2226. [Google Scholar] [CrossRef]
  199. Al-Gharibi, K.A.; Sharstha, S.; Al-Faras, M.A. Cost-Effectiveness of Wound Care: A concept analysis. Sultan Qaboos Univ. Med. J. 2018, 18, e433–e439. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Summary of wound healing stages and fibroblast involvement.
Figure 1. Summary of wound healing stages and fibroblast involvement.
Ijms 25 03274 g001
Figure 2. Summary of compounds/delivery systems and their influence on fibroblast behavior.
Figure 2. Summary of compounds/delivery systems and their influence on fibroblast behavior.
Ijms 25 03274 g002
Figure 3. Relative scale of natural compound sources and delivery systems.
Figure 3. Relative scale of natural compound sources and delivery systems.
Ijms 25 03274 g003
Table 1. Natural compounds.
Table 1. Natural compounds.
Compound CategoryCompound NameImprovement of Wound HealingModification of Fibroblast BehaviorSources
Plant-derived productsSarıçiçek (Achillea biebersteinni Afan)Anti-microbial, anti-oxidant, and anti-inflammatory propertiesDownregulates TGF-β1 and upregulates bFGF expression at the gene and protein level in murine embryonic fibroblastsHormozi, 2019 [50]
Mssillou, 2022 [45]
Astragalus propinquusPromotes re-epithelialization, revascularization, and immune functionPromotes cytokine secretion of TGF-β1, bFGF and EGF; promotes re-epithelialization; promotes proliferation, migration, and cell cycle progression of human skin fibroblastsEl-Ashram, 2021 [51]
Astragali Radix and Rehmanniae RadixPro-angiogenic and anti-inflammatory propertiesActivate genes in TGF-β1 pathway, regulate gene transcription for ECM synthesis via Smad pathway and cell motility via Ras/MAPK (non-Smad) pathway, and enhance skin fibroblast migrationZhang, 2012 [52]
QuercetinAnti-bacterial, pro-angiogenic, and anti-oxidant propertiesModulates fibroblast activity, upregulate TGF-β1El-Sherbeni, 2023 [53]
Mssillou, 2022 [45]
Falbo, 2023 [31]
CurcuminAnti-bacterial, anti-oxidant, anti-inflammatory, and pro-angiogenic properties.Induces fibroblast proliferation and collagen depositionEl-Ashram, 2021 [51]
Mssillou, 2022 [45]
Falbo, 2023 [31]
LuteolinAnti-bacterial, anti-oxidant, anti-inflammatory, and pro-angiogenic propertiesModulates IGF, PDGF, and FGFMssillou, 2022 [45]
Falbo, 2023 [31]
El-Sherbeni, 2023 [53]
KaempferolAnti-neoplastic, anti-inflammatory, anti-bacterial, and anti-oxidant propertiesIncreases hydroxyproline and collagen in wound tissueMssillou, 2022 [45]
El-Sherbeni, 2023 [53]
IcariinAnti-oxidant, anti-inflammatory, and anti-apoptotic propertiesAccelerates collagen depositionSingh, 2019 [54]
Mssillou, 2022 [45]
MorinAnti-oxidant, anti-inflammatory, and anti-apoptotic propertiesAccelerates collagen synthesisPonrasu, 2018 [42]
Mssillou, 2022 [45]
NaringinAnti-oxidant, anti-inflammatory, anti-microbial, and astringent propertiesAccelerates collagen synthesisMssillou, 2022 [45]
Catechin and Epigallocatechin-3-Gallate (EGCG)Anti-bacterial, anti-oxidant, anti-inflammatory, and pro-angiogenic propertiesEnhance wound contraction and modulate growth factorsHernandez-Hernandez, 2017 [43]
Mssillou, 2022 [45]
Falbo, 2023 [31]
SilymarinAnti-oxidant and anti-inflammatory propertiesIncreases number of fibrocytes, improves alignment of healing tissues, and enhances collagen fibers and fibroblastsOryan, 2012 [55]
Mssillou, 2022 [45]
HesperidinAnti-inflammatory, anti-microbial, anti-fungal, anti-oxidant, anti-neoplastic, anti-hypertensive, pro-angiogenic, and anti-atherogenic propertiesUpregulates TGF-β and Smad 2/3 mRNA expressionLi, 2018 [56]
Mssillou, 2022 [45]
Vicenin-2Anti-oxidant, anti-inflammatory, and pro-angiogenic propertiesInduces TGF-β to enhance fibroblast proliferation, migration and wound contractionTan, 2019 [43]
TanninsAnti-oxidant, pro-angiogenic, and antibacterial propertiesImproves fibroblast proliferation and promote wound contractionLi, 2011 [57]
Falbo, 2023 [31]
Terpinolene and α-phellandrene/α-pinene (PIN) and α-phellandreneAnti-oxidant, anti-bacterial, anti-fungal, and anti-inflammatory propertiesImprove migration and proliferation of fibroblastsBonnard, 2022 [58]
Salas-Oropeza, 2020 [59]
Salas-Oropeza, 2021 [60]
Falbo, 2023 [31]
ThymolAnti-oxidant, anti-inflammatory, cicazitrant, anti-septic, anti-bacterial, and anti-fungal propertiesInduces denser, thick, and parallel-arranged collagen fibersMarchese, 2016 [61]
Riella, 2012 [62]
Falbo, 2023 [31]
TaspineAnti-bacterial, anti-inflammatory, anti-viral, and anti-neoplastic propertiesStimulates fibroblast chemotaxis, and induces hydroxyproline and KGFPorras-Reyes, 1993 [63]
Wang, 2022 [64]
Vaisberg 1989 [65]
Falbo, 2023 [31]
ThymoquinoneAnti-microbial, anti-inflammatory, anti-oxidant, and anti-neoplastic propertiesEnhances fibroblast formation and augments wound contractionAlgahtani, 2021 [66]
El-Sherbeni, 2023 [53]
APS2-1 from AtragalusAnti-inflammatory, anti-oxidative, and immune-regulatory propertiesPromotes fibroblast propagation and accelerate cell cycle progression, and promotes expression of TGF-β1, bFGF, and EGFZhao, 2017 [67]
Yang, 2024 [68]
ZWP from Curcuma zedoariaPro-angiogenic propertiesEnhance collagen synthesis and depositionXu, 2018 [69]
El-Sherbeni, 2023 [53]
Extract of Sargasum Ilicifolium seaweed speciesIncreased speed of wound closureIncreases myofibroblast activity, promote TGF-β1 expressionPremarathna, 2020 [49]
Madecassoside and Asiaticoside from C. asiaticaImproved speed and quality of wound healingActivate the TGF-β/Smad pathway, enhancing type I and III collagen expressionWu, 2012 [70]
Animal-derived productsHoneyAnti-bacterial, anti-oxidant, anti-inflammatory, and pro-angiogenic propertiesInduces fibroblast proliferation and migration, and induces collagen matrix developmentRatcliffe, 2014 [71]
Ibrahim, 2018 [72]
El-Ashram, 2021 [51]
SericinAnti-oxidant, pro-angiogenic, and anti-inflammatory propertiesRegulates TGF-β1 and TGF-β3 expression, and activates collagen productionEl-Ashram, 2021 [51]
Maggot excretions/secretions of Phaenicia sericataAnti-inflammatory, pro-angiogenic, anti-viral, and anti-neoplastic propertiesActivate and enhance the growth rate of fibroblastsPrete, 1997 [73]
Ratcliffe, 2014 [71]
Marine collagenPro-angiogenic, and anti-aging propertiesPromote fibroblast migrationGeahchan, 2022 [33]
Chandika, 2015 [74]
Sea cucumbersAnti-bacterial, anti-inflammatory, anti-oxidant, and immune regulatory propertiesStimulate fibroblast chemotaxis and proliferation, and breakdown of ECM proteinsEl-Ashram, 2021 [51]
Ibrahim, 2018 [72]
Exogenous growth factorsTGF-βPro-angiogenic, and immune regulatory propertiesPrompts differentiation of fibroblasts into myofibroblasts and ECM formation/ deposition via TGF-β1 and TGF-β2 chemotaxis of fibroblasts via TGF-β3Walraven, 2017 [75]
Barrientos, 2008 [76]
Dolati, 2020 [77]
EGF and NRG1Anti-inflammatory and pro-angiogenic properties, and enhanced kerotinocyte recruitment and cell motilityStimulate recruitment of fibroblastsYoon, 2018 [78]
Dolati, 2020 [77]
FGF-2 (bFGF)Anti-inflammatory and pro-angiogenic, properties and enhanced kerotinocyte recruitment and wound contractionModulates ECM formation and inhibits TGF-β1/Smad-dependent pathwaysBorena, 2015 [79]
Dolati, 2020 [77]
IGF-1Anti-inflammatory, anti-apoptotic, and pro-angiogenic properties, and stimulator of keratinocyte proliferationStimulates proliferation of fibroblasts and collagenTodorović, 2008 [80]
Dolati, 2020 [77]
hPDGF and rPDGF-BPro-angiogenic properties; stimulates chemotaxis of polymorphonuclear leukocytes and monocytes; induce MMPs and tissue inhibitors of metalloproteinases (TIMPs)Stimulate fibroblast mitogenesis and chemotaxis; promote procollagen type I synthesisPierce, 1988 [81]
Dolati, 2020 [77]
Growth Factor-rich productsPlatelet-rich plasma (PRP)Anti-microbial, anti-inflammatory, and pro-angiogenic propertiesConsists of high levels of PDGF and TGF-β1Fotouhi, 2018 [82]
Dolati, 2020 [77]
Platelet-rich fibrin matrix (PRFM)Anti-inflammatory and pro-angiogenic propertiesStimulates release of PDGF, TGF-β1, EGF, FGF-2, and IGFLin, 2018 [83]
Dolati, 2020 [77]
Decellularized and dehydrated human amniotic membrane (DDHAM)Anti-bacterial, anti-inflammatory, and pro-angiogenic propertiesStimulate release of PDGF, TGF-β1, EGF, FGF-2, TGF-a, placental GF, G-CSF, Interleukin (IL)-4,
IL-10, and various TIMPs
Sheikh, 2014 [84]
Smiell, 2015 [85]
Dolati, 2020 [77]
Table 2. Delivery systems.
Table 2. Delivery systems.
Nano DDSNatural CompoundModification of Fibroblast BehaviorSources
Liposomes
Ability to improve bioavailability, cause sustained transdermal delivery of different medicinal compounds, and overcome possible drug overdose and toxicity.
CurcuminShortens inflammatory process, inhibits bacterial growth, promotes fibrosis, angiogenesis, re-epithelialization, and wound contractionKianvash, 2017 [131]
MadecassosideSignificant burn wound healing effectLi, 2016 [132]
Usnic acidInhibits the secretion of pro-inflammatory cytokines, TNF-α, IL-6, IL-1B
Induces nitric oxide and cyclooxygenase-2 (COX-2)
Increases IL-10 and HO-1 in a dose-dependent relation
Anti-bacterial activity
Enhances maturation of granulation tissue and better collagen deposition
Nunes, 2016 [133]
bFGFPromotes fibroblast proliferation, migration, differentiation
Expedites regeneration of vascular vessels and the synthesis of procollagen and collagen matrix
Xu, 2017 [134]
Insulin/chitosanIncreases re-epithelialization collagen content, granulation tissue, wound tensile strength, and local production of insulin-like growth factors by fibroblasts. Increases proliferation and migration of human keratinocytes, which stimulates cell growth and enhances wound healingDawoud, 2019 [135]
Transfersomes
Deformable liposomes with an edge activator.
Gellan cholesterol nanohydrogels/baicalinInhibits TNF-αInhibits IL-1β Visually improves wound healingManconi, 2018 [136]
Polymeric nanoparticles
Protect the degradation of drugs and release them in a controlled manner.
PDGF-A, IGF-1, EGFAdvanced granulation tissue formation, significantly enhances healing of chronic woundsChoi, 2017 [137]
PLGA/LL-37Increases collagen deposition, and organization, enhancement of epithelialization, and neovascularizationChereddy, 2014 [138]
hVEGF gene/stem cellsEnhances angiogenesis, and reduces tissue degeneration and fibrosis in ischemic limbsYang, 2010 [136]
Thymol/chitosan/AgNPsExcellent anti-bacterial propertiesManukumar, 2017 [139]
Inorganic nanoparticles
Deprived from inorganic materials, including metallic nanoparticles, carbon-based nanoparticles, and ceramic nanoparticles. Benefiting from the intrinsic nature of materials, inorganic nanoparticles exhibit both similar merits in wound healing treatment and a strong anti-bacterial effect.
Iron oxide/thrombinIncreases tensile strength of wounds, decreases inflammationZiv-Polat, 2010 [140]
Cerium oxideEnhances fibroblast proliferation, myofibroblast differentiation
Accelerates migration and tube-forming ability of vascular endothelial cells
Chigurupati, 2013 [141]
Zn02Shows anti-bacterial activity and enhances wound healingAli, 2017 [142]
GoldEnhances of wound healing, increases collagen expression, decreases MMP-1 expression and TGF-B1
Enhances VEGF, angiopoietin 1, and 2
Kim, 2015 [143]
Silane/amphotericin BShows efficacy in controlling Candida infectionSanchez, 2014 [144]
Lipid nanoparticles
Introduced to overcome the limitation of liposomes.
Controlled release of drugs due to their nontoxic colloidal dimensions.
rhEGFEnhances proliferation and migration of fibroblasts, wound contraction, and epidermal regenerationGainza, 2013 [145]
Nanofibrous structures
Mimic the ECM, provide favorable conditions for cell attachment and contact with drugs. Enhance variety of therapeutics agents due to their high area-to-volume ratio.
Andrographolide/silicaAccelerates wound healing, increases collagen deposition in the wound site, decreases inflammationJia, 2018 [146]
Astragaloside IVAccelerates wound healing and inhibits scar formation, increasing angiogenesis, regulating newly formed types of collagen, and improving collagen organizationShan, 2015 [147]
PDGF-BB/VEGFAccelerates wound healing, promotes fibroblast growth and inhibits bacteria in vitroXie, 2013 [148]
LawsoneSignificantly increases TGF-β1 and collagen gene expression in vitro and promotes re-epithelialization of the wound in vivoAbadehie, 2021 [149]
Nanohydrogel
High flexibility, high hydrophilicity, high mechanical strength, tunable structure, and the ability to absorb wound exudates as well as permeate oxygen and prevent wound dehydration.
Cellulose nanocrystal and hyaluronic acid/chitosan NPs/GM-CSFEnhances proliferation and differentiation of fibroblasts, lowers inflammation, and increases collagen depositionDehkordi, 2019 [150]
Carrageenan/nano silicatesEnhances cell adhesion and spreading, reduces blood clotting time, facilitates in vitro tissue regeneration and wound healingLokhande, 2018 [151]
Nanocellulose/acrylic acid hydrogelsMaintains the activity and morphology of human dermal fibroblasts, promotes rapid cell proliferation, and affects 9 genes’ expressions related to wound healingLoh, 2018 [152]
Hydrogels loaded with nanoparticles
Synergistic effect between hydrogels and nanoparticles encapsulated.
Chitosan hydrogels/phenytoinIncreases the content of collagen fibers and fibroblasts in the wound tissueCardoso, 2019 [153]
Thermosensitive hydrogel/gold NPsEnhances skin re-epithelialization, granulation tissue, vascularization, and collagen deposition
Modulates gene expression of inflammatory and anti-inflammatory mediators
Mahmoud, 2019 [154]
Hydrogels/cyclosporine A solid lipid NPsSignificantly increases rate of mucosal repairKaravana, 2012 [155]
Hyaluronic acid and chondroitin sulfate/asiatic acid/ZnO NPs/CuO NPsRaises DNA, total protein, hexosamine and hydroxyproline content, and leads to superior re-epithelization, collagen fiber arrangement and angiogenesisThanusha 2018 [156]
Hydrogel/Simvastatin polymeric NPsEnhances of epithelialization and wound healing
Decreases inflammatory cell infiltration
Aly, 2019 [157]
PRP/collagen NPsEnhances epithelialization and wound closureShalaby, 2023 [158]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Berry, C.E.; Brenac, C.; Gonzalez, C.E.; Kendig, C.B.; Le, T.; An, N.; Griffin, M.F. Natural Compounds and Biomimetic Engineering to Influence Fibroblast Behavior in Wound Healing. Int. J. Mol. Sci. 2024, 25, 3274. https://doi.org/10.3390/ijms25063274

AMA Style

Berry CE, Brenac C, Gonzalez CE, Kendig CB, Le T, An N, Griffin MF. Natural Compounds and Biomimetic Engineering to Influence Fibroblast Behavior in Wound Healing. International Journal of Molecular Sciences. 2024; 25(6):3274. https://doi.org/10.3390/ijms25063274

Chicago/Turabian Style

Berry, Charlotte E., Camille Brenac, Caroline E. Gonzalez, Carter B. Kendig, Thalia Le, Nicholas An, and Michelle F. Griffin. 2024. "Natural Compounds and Biomimetic Engineering to Influence Fibroblast Behavior in Wound Healing" International Journal of Molecular Sciences 25, no. 6: 3274. https://doi.org/10.3390/ijms25063274

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