Next Article in Journal
Sustainable Collagen Composites with Graphene Oxide for Bending Resistive Sensing
Next Article in Special Issue
Topographically and Chemically Enhanced Textile Polycaprolactone Scaffolds for Tendon and Ligament Tissue Engineering
Previous Article in Journal
Synthesis, Optical Properties and Cellular Toxicity of Water-Soluble near Infrared-II Fluorescent Assemblies Based on Pillar[5]arene
Previous Article in Special Issue
Functionalised Hybrid Collagen-Elastin for Acellular Cutaneous Substitute Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

An Overview of Scaffolds and Biomaterials for Skin Expansion and Soft Tissue Regeneration: Insights on Zinc and Magnesium as New Potential Key Elements

1
Department of Plastic, Reconstructive and Aesthetic Surgery, Faculty of Medicine, University Hospital Cologne, Kerpener Str. 62, 50937 Cologne, Germany
2
Biotechnology Department, Faculty of Science, Cairo University, Giza 12613, Egypt
3
Translational Matrix Biology, Medical Faculty, University of Cologne, 50923 Cologne, Germany
4
Cologne Excellence Cluster on Cellular Stress Responses in Aging-Associated Diseases (CECAD), University of Cologne, 50923 Cologne, Germany
5
Center for Molecular Medicine (CMMC), University of Cologne, 50923 Cologne, Germany
6
Department for Oral and Craniomaxillofacial and Plastic Surgery, University of Cologne, Kerpener Strasse 62, 50931 Cologne, Germany
*
Authors to whom correspondence should be addressed.
Polymers 2023, 15(19), 3854; https://doi.org/10.3390/polym15193854
Submission received: 29 June 2023 / Revised: 13 September 2023 / Accepted: 18 September 2023 / Published: 22 September 2023
(This article belongs to the Special Issue Biomaterials for Tissue Engineering and Regeneration II)

Abstract

:
The utilization of materials in medical implants, serving as substitutes for non-functional biological structures, supporting damaged tissues, or reinforcing active organs, holds significant importance in modern healthcare, positively impacting the quality of life for millions of individuals worldwide. However, certain implants may only be required temporarily to aid in the healing process of diseased or injured tissues and tissue expansion. Biodegradable metals, including zinc (Zn), magnesium (Mg), iron, and others, present a new paradigm in the realm of implant materials. Ongoing research focuses on developing optimized materials that meet medical standards, encompassing controllable corrosion rates, sustained mechanical stability, and favorable biocompatibility. Achieving these objectives involves refining alloy compositions and tailoring processing techniques to carefully control microstructures and mechanical properties. Among the materials under investigation, Mg- and Zn-based biodegradable materials and their alloys demonstrate the ability to provide necessary support during tissue regeneration while gradually degrading over time. Furthermore, as essential elements in the human body, Mg and Zn offer additional benefits, including promoting wound healing, facilitating cell growth, and participating in gene generation while interacting with various vital biological functions. This review provides an overview of the physiological function and significance for human health of Mg and Zn and their usage as implants in tissue regeneration using tissue scaffolds. The scaffold qualities, such as biodegradation, mechanical characteristics, and biocompatibility, are also discussed.

1. Introduction

The human body is made up of various organ systems that are composed of distinct types of cells within an extracellular matrix (ECM), forming different tissues. These cells are highly complex biochemical, molecular, and electrical reaction chambers that are interconnected within the ECM by cellular receptors and cytoskeletal structures [1]. The human body and its constituent cells are subject to diverse mechanical forces on a daily basis, such as tension, compression, shear, gravity, osmotic pressure, and hydrostatic pressure. The impact of microgravity on cellular biology is also acknowledged, particularly in the context of space exploration [2,3,4].
The skin, being the body’s largest organ, acts as a protective barrier separating the internal and external environment and interacts with a variety of forces and deformations caused by the environment [5]. Cutaneous cell populations must sense mechanical cues and respond appropriately to maintain homeostasis and proper mechanical function [6]. Thereby, ECM plays a crucial role in transmitting applied forces to the cells. These forces trigger mechanosignaling pathways in the cells and elicit various biological responses [7,8,9]. The ECM is produced by fibroblasts, which reside in the dermis and are mainly responsible for the tissue’s mechanical properties. The mechanisms of mechanotransduction are similar among many cell types in the body’s various connective tissues [10]. At the molecular level, these responses may involve changes in the configuration of cell membrane channels or receptor sensitivity, enzymatic and protein synthesis in the cytoplasm, and gene expression in the nucleus. In response to these molecular and biochemical reactions, cells can differentiate, proliferate, migrate, or undergo apoptosis [11]. Mechanical forces are increasingly being leveraged to shape cellular and tissue responses in ways that promote tissue regeneration, scar modulation, and wound healing. In the case of the skin, it possesses the remarkable ability for self-regeneration through the presence of stem cells within the epidermis. However, when faced with deep injuries and severe burns, the natural healing process may prove insufficient, resulting in the development of severe scars, wound contraction, and chronic wounds [12]. Therefore, to overcome extensive mechanical forces, sophisticated surgical techniques have to be used to close the wounds with satisfactory scar formation. These also include the application of skin crafts, skin substitutes, and tissue expansion. In the late 1980s, tissue engineering emerged as a distinct field in response to the pressing surgical challenges that needed to be addressed [13,14]. There are several limitations associated with translational applications in soft tissue engineering strategies, including issues related to cell survival, mechanical challenges such as scaffold collapse and availability, considerations of the microenvironment’s composition, potential induction of malignant behavior, cell migration, and cell exhaustion [15]. Tissue expansion is a widely used surgical technique aimed at growing additional skin to address various reconstructive needs such as birth defects, burn injuries, or cancerous breasts [16]. The technique of tissue expansion has been in practice for over three decades and has proven to be a valuable tool in reconstructive surgery across various anatomical regions [17]. One of the major hurdles encountered by reconstructive surgeons is the scarcity of viable soft tissues for such procedures [18]. As a result, there is a growing interest in tissue-engineered skin substitutes as alternative approaches to traditional wound healing, skin expansion strategies, and tissue regeneration [14,19]. In this regard, we highlight in this review several types of implants used in translational applications in soft tissue engineering strategies, including magnesium and zinc, as remarkable promising materials to be used in the field of soft tissue engineering and skin expansion.

2. Mechanotransduction in Skin

Mechanically sensitive cells, especially fibroblasts, have three types of mechanical sensors at the cell membrane: integrins, G protein-coupled receptors, and stretch-activated ion channels. Furthermore, the cytoskeleton, which provides overall structural support to the cell, can sense deformations through conformational changes, resulting in an additional sensing mechanism [20]. Activation of mechanical sensors directly triggers intracellular signaling pathways, which often activate secondary messengers, such as growth factors [21]. Growth factor receptors located at the cell membrane represent another important sensing mechanism. In response to mechanical stimuli, various cytokines are expressed in connective tissues, including transforming growth factors (TGF-beta), interleukins, fibroblast growth factors (FGF), vascular endothelial growth factors (VEGF), platelet-derived growth factors (PDGF), and tumor necrosis growth factors (TNF-alpha) (Figure 1). These cytokines are particularly important in connective tissues and contribute to various physiological processes [10].
The maintenance of the ECM by dermal fibroblasts involves a constant cycle of collagen and proteoglycan deposition and degradation of the collagen network through matrix metalloproteinases (MMP). Among the various mechanical cues that fibroblasts experience, tension is the most physiologically relevant. Thus, in vitro studies have primarily focused on examining the response of fibroblasts to uniaxial and biaxial strain loading conditions using flexible two-dimensional constructs [22]. In order to better simulate the in vivo environment, mechanical strain has also been applied to fibroblasts embedded within three-dimensional collagen gels [23]. The application of tensile strain to the ECM induces conformational changes in the cytoplasmic tails of the main ECM receptors and the integrins, which activate kinases such as focal adhesion kinase (FAK). FAK activation is then linked to mitogen-activated protein kinase (MAPK) pathways inside the cell [24]. The downstream effects of FAK activation include pro-inflammatory signaling, collagen production, and reduced apoptosis (Figure 1) [25]. Aside from direct mechanical sensing, additional signaling pathways, particularly TGF-beta, play a critical role in controlling how the ECM is remodeled by fibroblasts [26]. TGF-beta exposure results in the up-regulation of collagen genes and the downregulation of the Bax apoptotic gene in fibroblasts [23]. Other mechanical stimuli, including the microstructure and composition of the ECM, also influence fibroblast behavior. For instance, fibroblasts have been found to migrate preferentially along fiber directions, demonstrating the importance of ECM factors in regulating fibroblast behavior [27,28].
Figure 1. In response to mechanical force, a number of intracellular signaling pathways are initiated in mechanotransduction. Membrane-bound mechanosensory complexes such as stretch-activated ion channels, growth factor receptors, integrins, and G-protein-coupled receptors play a crucial role in sensing mechanical strain. In fibroblasts and keratinocytes, where matrix-integrin activation takes place in focal adhesion complexes, FAK is crucial. The mechanical force that is transferred across the cell membrane activates downstream biochemical pathways, such as calcium-dependent targets, nitric oxide (NO) signaling, MAPKs, Rho GTPases, and phosphoinositol-3-kinase (PI3K). When these signals come together, transcription factors are induced to activate mechanoresponsive genes in the nucleus [29]. Parts of the figure were drawn using elements from Servier Medical Art. (https://creativecommons.org/licenses/by/3.0/) (accessed on 16 May 2022).
Figure 1. In response to mechanical force, a number of intracellular signaling pathways are initiated in mechanotransduction. Membrane-bound mechanosensory complexes such as stretch-activated ion channels, growth factor receptors, integrins, and G-protein-coupled receptors play a crucial role in sensing mechanical strain. In fibroblasts and keratinocytes, where matrix-integrin activation takes place in focal adhesion complexes, FAK is crucial. The mechanical force that is transferred across the cell membrane activates downstream biochemical pathways, such as calcium-dependent targets, nitric oxide (NO) signaling, MAPKs, Rho GTPases, and phosphoinositol-3-kinase (PI3K). When these signals come together, transcription factors are induced to activate mechanoresponsive genes in the nucleus [29]. Parts of the figure were drawn using elements from Servier Medical Art. (https://creativecommons.org/licenses/by/3.0/) (accessed on 16 May 2022).
Polymers 15 03854 g001
While the dermis is commonly considered the main load-bearing layer of the skin, it is important to note that keratinocytes in the epidermis also demonstrate mechanosensitivity. Deformations of the dermis are conveyed to the epidermis through hemidesmosome junctions at the basement membrane. Forces are then transmitted to the cytoskeleton inside the keratinocytes via adherens junctions between neighboring cells. Subsequently, intracellular signaling is induced by deformations of the overall cell shape that affect keratinocyte mitosis [30]. Stretch-activated ion channels represent another significant type of mechanoreceptor in the epidermis [31]. When activated downstream of a mechanical stimulus, growth factor receptors in keratinocytes, such as epidermal growth factor (EGF), play a crucial role in controlling cellular proliferation [6]. Experiments on cultured keratinocytes have demonstrated increased mitosis in response to strain [8]. More recently, in vitro studies have explored engineered skin, where both the dermal and epidermal layers respond to strain in a coordinated mechanobiological manner [32].
Mechanotransduction in vivo has mostly been investigated in relation to tissue expansion, which involves implanting a subcutaneous balloon that is persistently inflated over many weeks in order to grow skin [17,33]. By stretching the skin beyond its normal capacity, the expansion process leads to tissue growth (Figure 2) [34]. The ultimate aim of tissue expansion is to avoid donor site morbidity, using the newly created skin as a vascularized flap to reconstruct soft tissue defects during a second operation [35]. However, the procedure does have some limitations, such as potential failures and complications, as well as the need for training and skill development to effectively plan and execute tissue expansion [36]. The technique is also not appropriate for pre-existing open wounds, which represents its biggest drawback [37]. The analysis of expanded tissues has shown an increase in keratinocyte proliferation, activation of the MAPK pathway, and an increase in collagen deposition [38,39,40]. Interestingly, the grown tissue has similar properties to native, unexpanded skin.

3. The Influence of Mechanical Forces on the Structure and Function of the Skin

Mechanotransduction is the process of converting physical forces into biochemical signals that trigger cellular responses. The mechano-responsiveness of cellular complexes, such as TGFβ/Smad, integrin, and calcium ion pathways, has been demonstrated (Figure 1) [41]. These signals are transmitted into the cell, ultimately reaching the nucleus. In vitro models have shown that mechanical strain can upregulate matrix remodeling genes and downregulate normal cellular apoptosis through an Akt-dependent mechanism, leading to increased production of extracellular matrix [42,43]. The skin is exposed to stretching forces both under normal physiological conditions, such as pregnancy, and through external means, such as tissue expansion using soft tissue expanders, external skin stretching devices, and distraction osteogenesis using external devices in hard tissue.
Skin stretching devices and techniques are useful for treating open wounds in surgery and are referred to as external tissue expanders [44]. Typically, they use hooks, sutures, wires, or loops to engage the skin and apply a mechanical force of tension to promote the approximation of wound edges through mechanical creep over time. The stretched skin edges then heal spontaneously during a consolidation period while the stretching device remains engaged. In most cases, the mechanically stretched skin is surgically freshened at the edges and closed after the device is removed. The impact of mechanical skin stretching devices or techniques on healing wounds is increasingly gaining attention [45].
The process of wound healing engages different types of cells, including inflammatory cells, keratinocytes, fibroblasts, myofibroblasts, and endothelial cells. These cells play a pivotal role in cutaneous healing, and they react to mechanical forces by initiating a cascade of events and pathways at both the cellular and molecular levels. This process takes place in the context of the tensegrity model, which involves the cytoskeletal framework being anchored in ECM [39,46]. The structurally interconnected cells respond to mechanical stimuli. Mechanotherapies are wound healing treatments that use mechanical forces to enhance the healing process. Examples of these therapies include micro-deformational wound therapy (MWT), such as negative pressure wound therapy (NPWT) [47,48], shock wave therapy [49,50], ultrasound [51], and electrotherapy [52]. Each therapy uses a different form of mechanical force to stimulate the cells and tissues involved in wound healing. The effects of NPWT have been extensively studied [53]. When suction is applied, the sponge collapses, causing the wound to shrink. This results in macro- and micro-deformation at the wound-sponge interface, which triggers mechanosignaling in a closed wound-healing environment [54]. This therapy has various biological effects, such as increased gene expression of leucocyte chemoattractants, proliferation and migration of epithelial cells and dermal fibroblasts, decreased activity of matrix metalloproteinases, and increased micro-vessel density [55,56,57,58]. As a result, NPWT promotes moist wound healing, angiogenesis, collagen synthesis, and the breakdown of dead tissue and fibrin [59]. Currently, NPWT is extensively utilized to expedite the healing of different types of wounds across various anatomical locations.
Scar formation is a significant concern after skin injury both functionally and aesthetically [60]. In contrast to the scarring process, which occurs in extra-uterine life, fetal wound healing follows a regenerative process [61,62,63]. This has sparked considerable research interest in the conversion from regenerative healing to scarring, with the ultimate goal of achieving scarless wound healing. Currently, there is growing interest in mechanobiology and scar research, with the aim of utilizing mechanotherapy to prevent and treat abnormal scarring [64]. The relationship between tension and scar growth has been observed clinically, particularly in keloid formation, where different shapes and configurations are seen, often associated with tension [65,66,67,68]. Animal models have demonstrated that mechanosignaling plays a role in the fibroproliferative response to tension [64]. Mechanotherapy, which aims to alter stimuli, processing, or reception, offers strategies to deal with these tension forces, such as skin stabilization by paper and silicone tape, multilayered suturing and plication, flaps, z-plasty, and the addition of radiotherapy [69]. The goal is to reduce skin tension, which is the source of cyclically applied mechanical forces in daily locomotion, and ultimately decrease skin inflammation.
Skin tissue engineering encompasses a multidisciplinary approach, incorporating diverse fields such as biochemistry, polymer chemistry, and stem cell research. The objective of skin tissue engineering is to synergize the expertise from these disciplines in order to develop a substitute that can be efficiently produced and effectively restore the skin’s natural functional, mechanical, and aesthetic characteristics. This involves regenerating ECM to provide support and guidance, improving graft take by establishing a vascular network, and restoring skin appendages for functions such as thermoregulation and sensitivity, as well as the various cell types required for protection. The objective can be achieved through two primary methods. The first method involves creating a biodegradable scaffold that is sophisticated enough to release a specific set of signaling molecules in a controlled manner, which can facilitate the migration, adhesion, and, ultimately, regeneration of skin cells. The second method involves designing a basic, temporary scaffold that can serve as a carrier for stem cells or undifferentiated cells to encourage skin regeneration [70]. Both methods necessitate the creation of a 3D environment that can support cell interactions and foster wound healing.
Even though we have emphasized how important mechanical forces and load-induced events are in skin tissue engineering, it is still critical to provide a quantitative perspective, especially when considering various tissues [71]. Quantitative data can provide a more thorough understanding of how different tissues are impacted by mechanical forces, resulting in more accurate and successful tissue engineering techniques [72]. Depending on the tissue under examination, there are several ways to quantitatively present mechanical forces, including tensile strength and strain, strain distribution, shear forces, compression and stress relaxation, fluid flow and perfusion, frequency and magnitude of mechanical loading, microenvironmental stiffness, and biomechanical properties of biomaterials [73,74,75,76]. Thus, researchers are exploring numerous scaffold materials and techniques to meet the requirements of skin tissue engineering.

4. Skin Expansion in Reconstructive Surgery

Soft-tissue expanders have emerged as a pre-augmentation technique in implant surgery to circumvent the complications associated with bone grafting procedures [77,78]. The principle of soft tissue expansion is rooted in the biological response of various soft tissues, including skin and mucous membranes, to mechanical forces by producing true tissue growth (cell proliferation) [17]. This phenomenon is evident in various situations, including pregnancy, muscle growth, obesity, and specific cultural practices such as lip and neck expansion in African customs [79]. Tissue expansion offers a remarkable approach to cultivating skin that closely resembles the neighboring healthy skin in terms of texture, color, and hair-bearing characteristics, thereby minimizing scarring and the potential for rejection [80]. The technique of soft tissue expansion is clinically useful in several ways, including preoperative expansion of oral mucosa for large bone augmentations, as well as the intra-oral repair of clefts in the lip and/or palate. Its applications have been popular in plastic surgery since 1976 [81]. Moreover, they are well-established for a variety of indications, such as correcting skin burns after burn wounds, scars, alopecia, congenital nevi, and post-mastectomy breast reconstruction [82,83,84,85]. It has also evolved into one of the principal surgical procedures for creating skin flaps to resurface large congenital defects of the skin, such as giant nevi and vascular anomalies [86,87]. In recent years, this concept has also been introduced in orthopedics, where it was successfully used in a clinical report to achieve skin closure in open fractures using an “external” soft tissue expander. The expansion of soft tissues can reduce the need for periosteal incisions and promote passive flap closure while generating tissues with appropriate color match and texture similar to the original tissues [88].
In 1957, Neumann developed soft tissue expanders using a subcutaneous rubber balloon to repair an ear defect. However, it was not until the early 1980s that soft tissue expanders regained significant interest, particularly in breast reconstruction [89] and the treatment of burns [90]. Early expanders consisted of silicone rubber and featured an external valve that allowed for manual inflation through sequential injections. The extent of soft tissue expansion achieved with conventional expanders has been documented to be influenced by factors such as the specific tissue being expanded and the configuration of the expanders themselves [91,92]. Studies have shown that tissue gain is more pronounced with rectangular and crescent-shaped expanders than with round-based ones [79]. Although conventional soft tissue expanders have shown positive results, they have several disadvantages. The intermittent inflations required for conventional expanders can increase the treatment time by several months and cause pressure peaks, leading to a decrease in tissue vascularity [93] and a higher risk of expander perforation through the soft tissues [88]. The reduction in local oxygen partial pressure increases the risk of expansion failures, and serial injections can result in increased treatment costs, morbidity, and risks for adverse effects [82]. Despite these drawbacks, conventional expanders are still utilized in plastic surgical procedures. However, their use is limited in craniofacial defects due to the aforementioned shortcomings [94].
A self-inflating osmotic soft tissue expander was developed by Austad & Rose (1982) to overcome the drawbacks of conventional soft tissue expanders. It was designed without an external port, and repetitive inflations were not necessary [95]. The new expander was made of a semi-permeable silicone membrane that contained a hypertonic sodium chloride solution. The expansion of the expander and subsequent growth of soft tissue were facilitated by a continuous influx of body fluids driven by an osmotic gradient. This led to an increase in the volume of the expander and the growth of surrounding soft tissues. However, the device had several drawbacks, such as leaks from the expander shell to the surrounding tissues, causing tissue necrosis. Wiese introduced a unique self-inflating osmotic soft tissue expander composed of hydrogel, comprising a polymer network and a variable aqueous component [96,97,98]. This expander, known as Osmed® (Ilmenau, Germany), was developed in 1999 and became the first commercially accessible self-inflatable osmotic expander. It received FDA approval in 2001 and has been available in the market since then. Osmotic expanders eliminate the need for repeated injections and instead inflate continuously by osmotic gradients without requiring any additional interventions. This consistent expansion, in contrast to intermittent inflation, stimulates the generation of new cells, tissue growth [92], and a greater final tissue gain [81,88,99].
The biomaterials used in the hydrogel expanders are the same as those used in contact lenses, providing high biocompatibility and causing no adverse effects such as toxicity, immune reactions, infections, or systemic manifestations [98]. Moreover, they do not provoke any inflammatory responses in the surrounding soft tissues, which is a crucial feature. The presence of methacrylate in ionic hydrogels enhances their osmotic potential, leading to a greater swelling capacity compared to non-ionic hydrogels [96,97,98]. The incorporation of “methyl” methacrylate, specifically in the osmotic hydrogel expanders, results in a higher swelling ratio when compared to “hydroxyethyl” methacrylate [100].
The polymer network of the hydrogel expander is insoluble in water due to the presence of cross-links, making it able to retain large volumes produced by swelling without dissolving [101]. Varga et al. (2009) sought to investigate alternative biomaterials and introduced a hydrogel osmotic soft tissue expander composed of acrylic acid (AAc), acrylamide (AAm), or N-isopropylacrylamide (NIPAAm) [102]. Among these, NIPAAm hydrogels were found to be the most suitable for plastic and reconstructive surgeries in terms of their biological and mechanical properties, although they have only been tested in vivo and require further validation in clinical trials. The next section will delve into the techniques used in scaffold fabrication employed in tissue engineering, shedding light on the advancements, challenges, and future prospects in this exciting field.

5. Scaffold Fabrication Methods Used for Tissue Engineering

Numerous techniques have been devised for constructing and fabricating scaffolds in tissue engineering. The choice of technique depends on the specific properties of the materials employed and the desired characteristics of the scaffold. These methods can be classified into conventional and advanced techniques [103].
Conventional techniques encompass several methods, including solvent-casting and particulate-leaching techniques, which entail the utilization of a polymer solution blended with salt particles of precise dimensions. Subsequent to solvent evaporation and immersion in water, the salt particles dissolve, creating a porous structure [104]. However, gas foaming involves subjecting molded biodegradable polymers to high pressures with gas-foaming agents such as CO2, nitrogen, water, or fluoroform. The polymers become saturated, leading to the nucleation and expansion of gas bubbles within the polymer matrix, typically ranging in size from 100 to 500 μm [105,106]. Moreover, phase separation entails the rapid cooling of a polymer solution, leading to its separation into two separate phases: a polymer-rich phase and a polymer-poor phase. The polymer-rich phase solidifies, while the polymer-poor phase is removed, resulting in the creation of a porous polymer network with high permeability [107]. In melt molding, a combination of polymer powder and porogen components is introduced into a mold, which is then subjected to elevated temperatures beyond the glass-transition temperature of the polymer, accompanied by the application of pressure. This process causes the raw materials to fuse together, forming a scaffold with a predetermined external shape. After removing the mold, the porogen is washed away, leaving behind a porous scaffold that is subsequently dried [108]. Freeze drying, also known as lyophilization, offers a method for the production of polymeric porous scaffolds. The process involves two stages. Initially, the polymer solution is cooled to a specific temperature, causing all components to freeze. During this freezing stage, ice crystals form from the solvent, prompting the polymer molecules to aggregate within the interstitial spaces. In the subsequent phase, the solvent is eliminated by applying a pressure lower than the equilibrium vapor pressure of the frozen solvent. As the solvent undergoes sublimation, a dry polymer scaffold with a well-connected porous microstructure is left behind. The porosity of the scaffolds is contingent upon the concentration of the polymer solution, while the freezing temperatures affect the distribution of pore sizes. In addition to its use in fabricating porous scaffolds, this technique finds application in drying biological samples to safeguard their bioactivities [109,110].
On the other hand, electrospinning and 3D printing technologies are considered advanced techniques in scaffold fabrication. The former methodology is a fabrication technique that utilizes electrical charges to create ultrafine fibers on a nanometer scale. It has found extensive application in the production of porous scaffolds with nanofibrous structures, closely resembling the architecture and biological properties of the native extracellular matrix [111]. This versatile method enables the generation of fibers ranging from 2 nm to several micrometers in diameter, utilizing solutions composed of both natural and synthetic polymers. The resulting scaffolds exhibit small pore sizes and possess a high surface area-to-volume ratio, making them suitable for various biomedical applications [103,112]. Furthermore, 3D printing technologies encompass a range of methods employed for scaffold fabrication, utilizing CAD/CAM technology (computer-aided design/computer-aided manufacturing) [113]. These techniques serve as viable alternatives to address the drawbacks associated with conventional approaches, such as the utilization of cytotoxic solvents and limited control over porosity. By leveraging this technology, it becomes possible to create patient-specific scaffolds with precise shapes guided by computed tomography (CT) images [114]. Multiple 3D printing technologies exist, each distinguished by their unique construction methods and materials employed during the production process [115]. To achieve successful outcomes in the fabrication of skin substitutes, understanding the properties and characteristics of these materials is crucial, which will be highlighted in the next section.

6. Materials Used for Skin Tissue Engineering

Biomaterials are substances that have been specifically designed to assume a particular form, either independently or as part of a more complex system, to influence and guide therapeutic or diagnostic procedures in the field of human or veterinary medicine by controlling their interactions with living systems [49]. Recently, due to the increasing aging of the world’s population, there has been a rise in bone-related diseases and fractures, necessitating treatments that include implants with or without complementary functionality, such as biocompatibility, biodegradability, and antibacterial activity for infection control or growth hormones. To ensure an implant’s success, longevity, and desired function, it is essential to select a suitable biomaterial for the proposed application. These can be classified into natural and synthetic biopolymers, bimetals, bioceramics, and biocomposites.

6.1. Natural Materials

Natural materials, including silk, collagen, elastin, chitosan, and fibronectin, have garnered considerable interest in the development of skin substitutes. The utilization of these biologically-derived components offers a significant advantage, as they enable the creation of scaffolds that are both biocompatible and biodegradable. Moreover, the degradation products resulting from the breakdown of these natural polymers are non-toxic, further enhancing their suitability for biomedical applications [70]. Additionally, natural polymers contain peptides that have evolved over time to provide signals that promote wound healing. However, natural polymers have certain drawbacks, such as batch-to-batch variation, the potential for immune rejection, and the risk of pathogen transfer.

6.1.1. Silk

Despite being utilized as sutures in clinical practice for centuries, silk has only recently garnered significant attention as a natural biomaterial for tissue engineering. Silk exhibits unique properties as a lightweight polymer, possessing a tensile strength comparable to that of Kevlar 49, which is an aramid fiber used in composite materials with polymeric organic materials as reinforcement. Notably, silk is also highly elastic and requires a greater amount of energy to break compared to Kevlar 49 [116,117]. Moreover, silks exhibit thermal stability of up to approximately 250 °C, enabling processing at a broad range of temperatures [118]. Silk fibers commonly studied in the field include cocoon silk derived from the silkworm Bombyx mori and dragline silk obtained from the spider Nephila clavipes [119,120,121,122]. The process of silk fiber formation involves the coating of a filament core protein known as silk fibroin with a sericin protein-based adhesive substance [116]. Structurally, both B. mori and N. clavipes silks exhibit distinct blocks of hydrophobic and hydrophilic amino acid sequences [123,124]. The hydrophobic blocks form β-sheets or crystals through hydrophobic interactions or hydrogen bonding, which provide the silk fibroin’s tensile strength, while the less ordered hydrophilic blocks contribute to its elasticity and toughness [119,120,121]. The hydrophobic segments within silk fibroin-like proteins are utilized in genetic engineering approaches to modify host systems, including yeast, E. coli, plant, and mammalian cells. This genetic manipulation leads to the production of recombinant proteins that resemble silk fibroin and exhibit low water solubility, primarily attributed to their inherent hydrophobic nature [122,125,126]. Numerous studies have demonstrated that silkworm silk can foster the attachment and growth of human fibroblasts [127,128,129,130]. Nevertheless, sericin has been identified as a primary cause of adverse immune reactions [131]. The successful elimination of sericin and subsequent regeneration of silk fibroin have resulted in the development of biocompatible [131,132,133], hemocompatible [134], and materials that possess excellent oxygen and water permeability [135]. Furthermore, the utilization of silk fibroin films and composites in wound dressings has demonstrated enhanced healing capabilities in vivo [136,137].

6.1.2. Chitosan

Chitosan, a linear polysaccharide made up of glucosamine and N-acetyl glucosamine units linked by β (1–4) glycosidic bonds, is produced by the partial deacetylation of chitin, the second most abundant natural polymer found in the exoskeletons of crustaceans [138]. Its hydroxyl and amino groups can be modified to synthesize different derivatives of chitosan [139,140,141]. The extent of deacetylation, which can vary from 30% to 95% depending on the source and preparation method, also influences the molecular weight and the quantity of glucosamine present [142,143]. While chitosan is insoluble in aqueous solutions with pH above 7, its protonated free amino groups on glucosamine make it soluble in dilute acids (pH < 6.0) [144]. However, the presence of cationic groups on chitosan poses difficulties for techniques such as electrospinning, as it necessitates a solvent capable of forming a salt with chitosan to disrupt the interactions between adjacent chitosan molecules [145,146,147,148]. These cationic groups also enable pH-dependent electrostatic interactions with anionic glycosaminoglycans (GAG) and proteoglycans [149]. Despite its challenges, chitosan has been shown to support the attachment and growth of cells, making it a promising material for tissue engineering. Chitosan has garnered interest in skin tissue engineering due to its biocompatibility, biodegradability, and bioactivity [150,151]. Furthermore, its capacity to enhance hemostasis, expedite tissue regeneration, and stimulate collagen synthesis by fibroblasts has established it as a valuable polymer [152,153,154,155]. Furthermore, the properties of chitosan are not lost when used to create double-polymer scaffolds. For example, when chitosan and alginate were combined to form a polyelectrolyte complex membrane, it demonstrated enhanced stability against pH changes compared to each material alone, leading to the development of more efficient controlled-release membranes [156]. This capability to retain properties after blending is beneficial in skin tissue engineering because it enables the combination of various materials to generate a scaffold with improved capabilities.

6.1.3. Collagen Type I

Collagen refers to a group of proteins that have a characteristic triple helix structure consisting of three polypeptide chains. These proteins can be classified based on their structure and organization into various types such as fibril-forming, fibril-associated, network-forming, anchoring fibril, transmembrane, basement membrane, and others, each with unique functions. Collagens are known to provide functional properties that promote cell attachment and proliferation [157]. Collagens exhibit a distinct structural pattern consisting of a right-handed triple helix, where each alpha chain forms an elongated left-handed helix with a pitch of 18 amino acids per turn [158]. Collagen molecules can exist as homotrimers or heterotrimers, with the chains staggered by one residue compared to one another and coiled around a central axis [159,160]. The presence of glycine at every third residue facilitates close packing of the alpha chains around this axis, with the bulkier side chains of other amino acids located in the outer positions [161,162]. In tissues such as skin, bone, and articular cartilage, a dynamic 3D environment is created by networks of fibril-forming collagens, including collagen types I, II, IV, V, and XI [161]. These collagens are capable of assembling into highly oriented supramolecular aggregates and display a banding pattern with a periodicity of around 70 nm when observed under SEM [163].

6.1.4. Elastin

The field of skin tissue engineering has shown a growing interest in elastin, a natural polymer. The monomer form, tropoelastin, cross-links to produce the insoluble biopolymer elastin [164]. Tropoelastin is encoded by a single-copy gene located in the 7q12.2 region in humans [165] and is characterized by alternating hydrophobic and hydrophilic domains [166]. These domains have different functions, with the hydrophobic domains responsible for monomer association and elastic function, while the hydrophilic domains facilitate polymerization through cross-linking [167]. The presence of elastin is commonly observed in elastic tissues such as skin, lungs, large arteries, and tendons [168]. Elastin is synthesized by a variety of cell types, including endothelial cells, fibroblasts, and smooth muscle cells [169]. Elastogenesis mainly occurs during the late fetal and early neonatal stages, and there is a limited turnover of elastin in healthy adult tissues. The durability of elastin is noteworthy as it has a half-life of around 70 years [170]. Elastic fibers in the body comprise two primary components, with elastin forming the core, encased within a sheath of microfibrils measuring approximately 10–12 nm in width [171]. Elastin is renowned for conferring flexibility, elasticity, and durability to the skin while simultaneously controlling its texture and quality [170]. When skin is damaged, elastin levels have been found to decrease or be absent, resulting in the reduced suppleness of scar tissue [172].
Elastin exhibits structural properties and inherent cell signaling properties such as chemotaxis, cell attachment, proliferation, and differentiation [166,173]. Due to these unique characteristics, elastin is an attractive polymer for skin tissue engineering, as it may improve the elasticity and cell-scaffold interactions of a skin substitute. Additionally, the presence of enzymatic elastin products has been found to stimulate elastin and collagen production. For example, cultures of dermal fibroblasts with elastin products showed a significant increase in elastin and collagen fiber production, as well as an increase in elastic fiber deposition in skin explants. This effect was also observed in human dermal fibroblasts (HDFs) injected into athymic nude mice, resulting in increased elastic fiber production [174].

6.1.5. Silicon

Silicon plays a crucial function as a biomaterial in supporting tissue regeneration. It is a remarkable component for consideration in the field of tissue engineering due to its distinct qualities and interactions with biological systems [175,176]. Silicon is extremely biocompatible in a variety of forms, including silicon-based ceramics and nanoparticles. This indicates that it can be utilized in close contact with biological tissues without having negative effects [177]. When considering its application as an implant or scaffold material in tissue engineering, its biocompatibility is especially beneficial. Moreover, Silicon has been reported to increase collagen formation, which is an important component of the extracellular matrix (ECM) [178]. Since the ECM offers structural support and signals for cell development and differentiation, this characteristic is crucial for tissue engineering, and its results may be improved by increased collagen production [179]. Silicon has been incorporated into a variety of biomaterials, including hydroxyapatite and bioglass, according to previous studies [180,181]. A favorable microenvironment for tissue regeneration is created by silicon-doped biomaterials, which offer a controlled and continuous release of silicon ions [182,183]. This controlled release mechanism of silicon ions ensures that the beneficial effects are delivered over an extended period, aiding in tissue healing. According to recent studies, silicon-based materials may have natural antibacterial characteristics [184]. For successful regeneration in tissue engineering, infection control is crucial. The possible antibacterial properties of silicon could be a useful addition to biomaterials, lowering the danger of post-implantation infections [184].

6.2. Synthetic Bioresorbable Polymers

Synthetic polymers, which are manufactured and easily obtainable, have gained attention in skin tissue engineering. Biodegradable, biocompatible, and bioresorbable synthetic polymers are preferred as they can be naturally degraded and eliminated without surgical intervention. Their predictable mechanical properties, such as tensile strength, offer an advantage in producing reliable treatment outcomes. Nevertheless, synthetic polymers do not possess the inherent biological signals present in natural polymers. Numerous synthetic polymers, including poly(lactic acid) (PLA), poly(glycolic acid) (PGA), and polycaprolactone (PCL), are being researched for use in skin tissue engineering [185].

6.2.1. Polycaprolactone (PCL)

Polycaprolactone (PCL) is an aliphatic polyester that undergoes biodegradation through hydrolysis and has obtained approval from the Food and Drug Administration [186,187]. Despite being synthesized in the 1930s, it has recently regained attention in the field of tissue engineering due to its biocompatibility, high tensile strength, and controllable biodegradability [188]. The degradation rate of PCL can be adjusted and can range from several months to years based on factors such as molecular weight, degree of crystallinity, and degradation conditions of the polymer [189,190]. Furthermore, the degradation products of PCL are non-toxic, in contrast to other synthetic polymers, such as PLA, which may induce mild inflammation [163]. Clinical trials have been conducted on subcutaneously implanted PCL capsules, showing that PCL was well-tolerated for over 40 weeks, and other studies have used PCL as a drug release vehicle [191]. In one study, an ultrathin PCL film was developed as a wound dressing and tested in rat and pig models. The results showed that the PCL films had a lower level of fibrosis compared to non-dressed wounds. The PCL films did not induce inflammation, and the wound dressing supported normal wound healing in both partial and full-thickness wounds [187].

6.2.2. Poly(d,l-lactic-co-glycolic acid) (PLGA)

PLGA is a group of biodegradable and biocompatible polymers approved by the FDA, which has gained popularity due to its extended clinical use and ability to provide sustained drug delivery despite its mild inflammatory degradation products. The copolymer is made up of equal parts of PGA and PLA, which contain both the optically active d and l enantiomers of PLA, with PDLA and PLLA being asymmetrical α-carbons, respectively. PLLA can be highly crystalline, PDLA can be completely amorphous, and PGA is highly crystalline. The solubility of PLGA is widespread, encompassing commonly used solvents such as acetone and ethyl acetate [192,193]. Under aqueous conditions, PLGA degrades through hydrolysis at its ester linkages. Consequently, by including more of the less hydrophilic PLA, water absorption can be reduced, leading to slower degradation rates [194]. Moreover, factors such as molecular weight and storage temperature have been demonstrated to influence the physical properties of PLGA, including mechanical strength [195].

6.3. Absorbable Metallic Materials

It is possible to significantly enhance the effectiveness of tissue regeneration by introducing certain physiologically active substances, such as metal elements, growth factors, peptides, genes, and stem cells [196,197]. Metal elements play important roles in the structure or expression of several biomacromolecules, including proteins and enzymes, as vital parts of the human body [198,199,200]. Metals are highly desirable for load-bearing implants due to their excellent mechanical properties and biocompatibility. Numerous studies have demonstrated that the regulation of various metal elements, which are essential for cytokine regulation and immunological processes, is closely associated with the tissue regeneration process [201,202,203]. Furthermore, certain metal particles possess inherent antibacterial properties that can effectively combat invading pathogens [204,205]. As a result of advancing research into the mechanisms underlying metal elements in soft tissue regeneration, wound repair techniques incorporating metal elements have gained significant attention [206]. Ideal biomaterials must encompass considerations of biocompatibility, biomechanics, biodegradability, and biofunctionalization (Figure 3) [207]. Currently, stainless steels, titanium, and cobalt-chromium-based alloys are the most commonly used metallic biomaterials [208], and strontium (Sr), iron (Fe), zinc (Zn), and magnesium (Mg) are the most commonly utilized biodegradable metals in clinical practice [209]. Titanium alloys have gained popularity in orthopedic surgeries due to their superior biocompatibility, enhanced corrosion resistance, and lower modulus compared to stainless steels and cobalt-based alloys.
Iron is an indispensable chemical element in the human body and possesses favorable mechanical properties, high biocompatibility, and a slow degradation rate [209]. Its high elastic modulus is associated with high radial strength. However, the degradation rate of Fe is too slow for it to be widely employed in tissue engineering. Further investigations are needed to achieve a desirable corrosion rate, and Fe material properties must be adjusted for it to be suitable for biomedical purposes [210]. Strontium (Sr) is also considered a promising biomaterial with distinct properties that can influence tissue regeneration processes [211], where it was reported that Sr increases osteoblast activity and increases bioactivity when incorporated with HA lattice [212]. Zinc plays a crucial role in various biological functions, including nucleic acid metabolism, DNA synthesis, enzymatic reactions, and apoptosis regulation. It is present in different body parts, such as the skin, liver, bones, and muscles. Mg plays a crucial role in various bodily functions. The Mg ion (Mg2+) acts as a cofactor in over 300 enzymatic reactions, including protein and DNA/RNA synthesis, ion transportation, cell migration and function, and intracellular energy production through the ATP system [213,214]. The interaction between an absorbable metal and human body fluid may lead to the initiation of the anodic reaction, which is accompanied by the generation of electrons, which are subsequently consumed by the cathodic reaction. For mg-based alloys, the cathodic reaction involves water reduction, while for Zn-based alloys and Fe-based alloys, it involves the reduction of dissolved oxygen. In a physiological environment, the presence of high chloride ion concentrations results in the breakdown of the degradation layers and accelerates the degradation process. Depending on the size of the degradation particles, macrophages and/or fibrous tissue may encapsulate these particles until complete degradation of the metal occurs [215].
In the next sections, we will delve into the specific properties and applications of magnesium and zinc as biomaterials, building upon their essential roles in various biological functions in soft tissue regeneration and skin expansion.

6.3.1. Magnesium (Mg)

The total Mg2+ content in the normal adult body is estimated to be around 25 g, with approximately 53% found in the bone, and the extracellular Mg2+ accounts for about 1% of the total Mg2+ content [216]. Magnesium plays a vital role in numerous biological processes and is essential for sustaining life. Mg2+ is involved in enzymatic reactions through two key interactions: (1) binding to the substrate, forming a complex that interacts with the enzyme, such as Mg ATP in ATP-utilizing enzymes, and (2) binding to the enzyme itself, acting as an allosteric activator [217,218]. Magnesium ions possess a small size that enables them to permeate the skin, resulting in reduced inflammation, enhanced water binding to the skin, and accelerated repair of the skin barrier [219,220]. The improvement of skin barrier function relies on minimizing trans-epidermal water loss, maintaining a hydrated stratum corneum, and minimizing inflammation [221]. Notably, a previous study specifically examined the impact of magnesium salt on skin barrier recovery and concluded that it effectively expedites the healing process of the skin barrier [222,223,224]. Magnesium ions have garnered significant attention in the context of skin substitutes due to their ability to promote skin repair. Among the various forms of magnesium, magnesium oxide, an inorganic metal oxide, stands out as it releases soluble Mg2+ ions that play a crucial role in cellular processes [225]. These Mg2+ ions actively participate in wound healing by facilitating the recovery of damaged tissue cells, regulating cellular metabolism, and controlling enzyme activity. This mechanism enhances the healing process following traumatic injuries [226]. Additionally, magnesium influences the migration of keratinocytes, regulates epidermal differentiation and proliferation, and exhibits anti-inflammatory properties [227]. It also plays a role in maintaining the barrier function of the skin, thereby controlling the hydration of the stratum corneum [227]. Considering these factors, improving the skin barrier function of compromised skin holds paramount importance for magnesium. On the other hand, Mg deficiency can reduce parathyroid hormone (PTH) secretion, which, in turn, results in lower vitamin D levels [228]. Researchers have found that post-menopausal women who are vitamin D deficient have lower levels of PTH and Mg than those who are not deficient [229].
On the cellular level, magnesium has been shown to stimulate collagen synthesis in cultured fibroblasts, indicating its potential role in promoting connective tissue formation [230]. On the other hand, it also inhibits prolyl and lysyl hydroxylases [231]. Additionally, studies have indicated that Mg2+ is specifically associated with the elastin core within elastic fibers rather than with the associated microfibrils known as oxytalan fibers [232,233]. Elastin degradation is a significant process in various physiological events, including growth, wound healing, and tissue remodeling [234]. The association of Mg2+ with elastin core suggests that it plays a crucial role in safeguarding the extensibility of elastin, contributing to its mechanical properties. Therefore, it seems that Mg2+ is not only involved in maintaining the structure and mechanical integrity of elastic fibers but also actively participates in the elastolysis of these fibers [235]. In cartilage, magnesium-associated proteoglycans play a crucial role in preventing tissue swelling and degradation. According to a previous study [236], it has been observed that decorin proteoglycans, which are found in proximity to collagen fibers, have the ability to interfere with TGFβ/smad dependent transcriptional processes in human mesangial cells [237], where Mg2+ has the potential to exert its effects at the protein kinase II level. Furthermore, magnesium regulates the functional activity of integrins [238]. The modulation of integrins, which are involved in cell adhesion to extracellular matrix components, has been found to play a crucial role in facilitating cell migration [239,240]. Studies have reported that the presence of Mg2+ can enhance the adhesion of keratinocytes and fibroblasts to type I collagen and laminins (glycoproteins found in the basement membrane). Interestingly, this effect is dependent on the concentration of Mg2+, while the presence of Ca2+ counteracts this enhancement [241,242]. Cell migration may be facilitated by the interaction between integrins and MMPs if Mg2+ is able to modulate the activity of MMPs and cause changes in integrin conformation [243]. These diverse properties associated with magnesium make it serve as a key player in physiological and pathological situations involving connective tissue and matrix-associated cells. Magnesium is a highly performant biodegradable metal that possesses significant properties [244]. When ingested in quantities of 350 mg per day, 25 mg is deposited in the human body, with half of it in bones, while the remainder is excreted in urine [225]. Its density is around 1.7 g/cm3, while its Young’s modulus is approximately 42 GPa, making it quite similar to that of human bone (density of 1.95 g/cm3 and Young’s modulus ranging between 3 and 20 GPa) [245]. Magnesium biodegradation inside the human body leads to a rapid degradation process, which creates gas pockets and eliminates hydrogen. This results in a local alkalization process near the scaffold structure, leading to an increase in hydroxyl (OH-) ions [246]. The presence of these ions can worsen the physiological microenvironment and even cause alkaline poisoning effects at pH levels above 7.8. Moreover, if the evolution of hydrogen gas is generated in high quantities, it may lead to tissue damage and other complications [247]. The accumulation of hydrogen gas in the tissue has the potential to harm the tissue, cause inflammation, and make the patient uncomfortable. Although the body may safely absorb and tolerate tiny amounts of hydrogen gas, excessive accumulation can have negative effects [248]. According to a previous study, the Mg ions were not toxic to the kidneys or liver, and no major bone healing issues were reported. However, the emitted gas containing H2, CO, and CO2 can result in problems such as long-term osteolytic lesions and superficial skin necrosis [249]. On the other hand, excessive magnesium intake might cause health concerns. It is crucial to understand that magnesium overdose is rare and usually affects those who have kidney problems or who take supplements with exceptionally high dosages of the element [250]. Too much magnesium can cause gastrointestinal problems such as nausea, vomiting, and diarrhea [250]. These symptoms are usually the body’s way of getting rid of excess magnesium. High doses of magnesium can also lower blood pressure (hypotension), which can cause dizziness, nausea, or even tremors in extreme cases [251]. In severe cases, magnesium toxicity can affect the respiratory system, making it difficult to breathe [250]. Before the widespread application of magnesium as a biomaterial, it is crucial to obtain a comprehensive understanding of its corrosion behavior and point out the link between these health issues and its long-term, excessive use. Although the correlation between in vivo and in vitro corrosion behavior is not yet quantitatively established, corrosion tests are commonly conducted during the initial stage of alloy development to avoid costly in vivo animal trials [252]. Several published papers have concentrated on exploring the impact of individual components present in different media on the corrosion of magnesium [253,254]. Representative test media include SBFs (simulated body fluids), cell culture medium, and protein-containing media [255,256].
The impact of individual inorganic ions found in the simulated body fluids, such as Cl, carbonate, phosphates, sulfate, and Ca2+, on the rate of corrosion has been extensively studied. It has been found that the presence of carbonate and phosphates can slow down the corrosion rate of magnesium, while the effect of sulfate is not significant [252]. Cell culture medium, for example, SBFs, is a mild corrosive medium that tends to minimize differences in corrosion rates between various magnesium alloys.
Moreover, proteins have been shown to have various effects on the corrosion of magnesium, either accelerating or inhibiting it. Gu et al. reported that FBS (fetal bovine serum) had the opposite effect on the corrosion of Mg–Ca and AZ series alloys in DMEM [257]. Zhang et al. observed that the corrosion rate of Mg–Nd–Zn–Zr in the M199 cell culture medium was slowed down with the addition of 10% FBS under sterile cell culture conditions [258]. In DMEM, Johnson et al. found that the degradation of pure Mg was minimally affected by the addition of FBS, but the weight loss of Mg–Y increased [259].
Implantation in an animal body leads to alterations in the local physiological environment, such as inflammation, which affects the corrosive environment surrounding the implant. Moreover, mechanical stress during the service period also impacts the corrosion of implanted materials [260,261,262]. These factors cannot be easily represented comprehensively through corrosion tests. Therefore, it is impractical to entirely simulate the in vivo response through corrosion tests [263]. No single corrosion test method can provide complete answers to all the questions regarding Mg corrosion. The corrosion rate measured by such tests cannot fully represent the in vivo corrosion rate, which also varies [264]. Hence, we believe that the main objective of corrosion tests should be to identify the most effective metallic materials, organic and inorganic coatings, as well as various additives such as drugs or corrosion modulators and to understand their interactions. Of course, the criteria for evaluating performance should be based on agreement within the scientific community. Improving the Mg properties involves carefully monitoring and enhancing its degradation rate through surface treatment or alloying with chemical components that lead to a reduction in the amount of hydrogen gas and OH ions.

Magnesium and Its Alloys for Medical Applications

Mg and its alloys have emerged as promising biodegradable metals for medical implants, revolutionizing the use of metal implants in medical settings [265,266]. Currently, magnesium-based implants are widely used in two key applications: vascular interventions [267] and orthopedic procedures [268]. In 2016, BIOTRONIK (Berlin, Germany) introduced the first Mg-based stent that is commercially biodegradable [269]. The biodegradable nature of magnesium and its alloys, which are equivalent to those of polymers, the most widely used materials for such purposes, and their mechanical characteristics, such as density and elastic modulus akin to cortical bone, make them promising.
The degradation of Mg results in changes in the surface chemistry of the implant. This process creates degradation products that dissolve and elevate the pH in the peri-implant region while simultaneously protecting the implant from further degradation [270]. However, for the successful use of biodegradable Mg implants in tissue repair, it is necessary to ensure the proper maintenance of the implant during the healing period. If the degradation of the implant occurs too quickly, it can result in the formation of hydrogen bubbles that interfere with the attachment of proteins and cells to the implant surface, leading to premature implant failure. Therefore, the ideal degradation of Mg implants should start slowly and increase gradually over time once the damaged tissue has sufficiently healed [271].
It was demonstrated previously that the Mg2+-incorporated alginate hydrogel showed potential effects on the proliferation and differentiation of osteoblasts [272]. Another study by Roh et al. reported that the addition of MgO and Hap to the 3D PCL scaffold positively influenced various behaviors of pre-osteoblast cells, including initial adhesion, proliferation, and differentiation [273]. Additionally, Yuan et al. demonstrated that the bioresorbable microspheres made up of poly(lactide-co-glycolide) (PLGA) co-embedded with MgO and MgCO3 affect the efficiency in treating bone defects [274]. These findings suggest that the controlled delivery of Mg2+ ions through an appropriate scaffold could be a promising approach to enhance bone regeneration.
Alloying elements have been added to pure Mg to achieve a moderate and homogeneous degradation behavior. According to previous work, the alloy composition plays a crucial role in implant degradation, and in Guinea pig bones, the device should be present for at least 12 weeks to allow for healing [275]. The use of rare earth elements (REE) in Mg implants has yielded the most favorable response in in-vivo applications [275]. REEs, which consist of 17 elements, are commonly added to Mg to enhance its ductility, degradation resistance, and grain boundary strength. In vitro studies by Feyerabend et al. examined the impact of certain REEs on the viability, apoptosis, and expression of inflammatory cytokines in four types of cells. The findings indicated that lanthanum (La) and cerium (Ce) had the highest cytotoxicity, while gadolinium (Gd) and yttrium (Y) appeared to be conducive to promoting cell growth [276].
Some researchers have investigated the potential use of Mg alloys as open-porous scaffolds for load-bearing applications in tissue engineering. Witte et al. created two porous metallic scaffolds by casting an Mg–Al–Zn alloy and an Mg–Al–Zn–Mn alloy and tested them in vivo. Degradation occurred too rapidly, and the presence of Al and Zn caused inflammatory reactions [277]. A wide range of biodegradable Mg-based alloys with varying zinc contents was reported previously, such as Mg–Zn, Mg–Zn–Mn–Ca, Mg–Zn–Y, and Mg–Zn–Si [278,279,280]. In conclusion, the use of Mg and its alloys as biodegradable metals for medical implants has brought about a significant revolution in the field. Despite the promising properties of magnesium, research on its applications in skin expansion and soft tissue regeneration is still relatively limited. Further studies are needed to explore and understand the full potential of magnesium in these areas. The continued innovation and improvement in the field of magnesium and its alloy implants are paving the way for a wider integration of these devices across many therapeutic applications. These substances not only improve biodegradability but also display better mechanical qualities, opening the door to significant developments in the field of tissue regeneration and repair. As we explore the potential that magnesium-based materials can provide for the advancement of healthcare practices, the future does indeed seem hopeful.

6.3.2. Zinc (Zn)

The essential micronutrient zinc plays a vital role in supporting immune function. Despite the human body containing a total zinc amount of 2 to 4 g, there is no specialized storage system [281]. Consequently, a daily intake of zinc is necessary to maintain a steady state for optimal immune function. Zn obtained from dietary sources is absorbed in the small intestine and distributed through plasma. However, the concentration of zinc in the plasma is relatively low, measuring around 90 µg/dL, which accounts for less than 1% of the body’s total zinc content [282]. Although the plasma zinc pool is small, it holds significant immunological importance. Zinc primarily exists as an intracellular ion, with distribution occurring between various cellular compartments [283]. This distribution includes the cell nucleus (30–40%), cytoplasm, organelles, and vesicles (50%) [284]. Notably, there are specialized zinc-containing vesicles known as “zincosomes” that can store high levels of zinc upon stimulation [285,286].
Similar to plasma zinc, the majority of cellular zinc is tightly bound to proteins, resulting in a small portion of intracellular zinc remaining unbound or loosely bound, known as free zinc. Recent research has identified approximately 4000 proteins and a similar number of transcription factors that possess zinc-binding motifs [287,288]. The zinc tightly bound to proteins plays a crucial role in the catalytic, cocatalytic, and structural functions of enzymes [284]. It contributes to the stabilization of structural domains, such as zinc fingers and related structures, and facilitates protein–protein or protein–nucleic acid interactions, as observed in numerous transcription factors [289].
Throughout evolution, efficient mechanisms for maintaining zinc homeostasis have developed to prevent excessive accumulation of this essential micronutrient. These mechanisms involve two families of eukaryotic zinc transporters known as Zip and ZnT. The Zip family, also referred to as the Zrt-like, Irt-like Protein family, consists of 14 genes designated as solute carrier family 39 (SLC39) A1 to A14. Zips are responsible for transporting zinc into the cytosol. On the other hand, the ZnT family, comprising 10 genes (SLC30A1-10), facilitates the transport of zinc in the opposite direction. These transporter families not only regulate zinc levels within the cytosol but also play a role in controlling zinc distribution within cellular compartments such as the endoplasmic reticulum, mitochondria, and Golgi apparatus [290,291]. Additionally, aside from these transporter-mediated processes, zinc uptake can occur through diffusion involving amino acids, calcium-conducting channels, and various receptors [292,293]. Alterations in the concentration of free zinc within cells can impact signaling pathways, ultimately leading to modifications in cellular responses. This connection between intracellular zinc levels and signaling pathways has been demonstrated in various studies [286,294,295]. Additionally, cellular activation and stimulation can induce fluctuations in intracellular zinc levels [296]. This suggests a potential interaction between zinc homeostasis and signal transduction, implying that zinc may play analogous roles to calcium, which is a well-known second messenger [297].
Zinc is present both intracellularly and in the ECM of epidermal and dermal tissues, where it plays various crucial roles [298]. In human skin, the concentration of zinc is higher in the epidermis (50–70 μg/g dry weight) compared to the dermis (10–15 μg/g dry weight). This difference may reflect the involvement of zinc-dependent RNA and DNA polymerases in the mitotically active basal cells [299,300,301,302]. Immunohistochemical and in situ hybridization localization studies have revealed that in normal skin, high levels of metallothionein (MTs) are found in the basal epidermis, while their concentrations are reduced in postmitotic keratinocytes, reticuloendothelial cells, and fibroblasts [303,304,305]. The presence of MTs is associated with increased tissue zinc concentrations [305], and skin lacking MTs exhibits significantly lower zinc content compared to wild-type mice [303].
The interplay between zinc and calcium is crucial for basal cell mitosis and postmitotic maturation processes in normal skin, which involve keratohyalin synthesis and keratinization [306,307]. In thin hairy skin, where mitosis is inversely related to hair coverage, the levels of zinc and calcium are noticeably lower [308]. In sensory epithelia such as the nasal mucosa and tongue, higher levels of zinc are observed. These elevated zinc levels not only correlate with high mitotic activity, prolonged zones of keratinization, and abundant protein-bound phospholipids but also underscore the importance of zinc in taste and smell perception [309].
The inclusion of zinc ions can accelerate numerous biochemical and molecular processes involved in wound repair by promoting the up-regulation of MTs and zinc metalloenzymes [305,310]. Moreover, zinc overdose is a rare condition that often happens when people use excessive amounts of zinc supplements over an extended period [311]. High zinc dosages may result in nausea, vomiting, and diarrhea. These signs typically appear shortly after consuming too much zinc [311]. Chronic zinc overload may weaken the immune system, increasing the susceptibility to infections [312]. The risk of heart disease may also rise if zinc intake is too high because it lowers levels of “good” high-density lipoprotein (HDL) cholesterol [313]. On the other hand, any deficiency in the expression of zinc-finger transcription factors in the mRNA coding of growth factors is indicative of compromised wound healing [314,315]. The modulation of zinc metabolism through the regulation of MT genes is attributed to the influence of Interleukin-1 (IL-1) [316]. This mechanism provides a potential explanation for the significant rise in zinc levels during the initial inflammatory phase of experimental wounds [304,317,318]. In a rat wound model, it was observed that zinc levels in the wound margin escalated by 15–20% within 24 h and further increased to 30% during the peak formation of granulation tissue and epidermal proliferation [317]. The functional significance of zinc in repair systems is supported by the presence of zinc metalloenzymes such as RNA and DNA polymerases, alkaline phosphatase, and MMPs, indicating their involvement in various biological processes [319]. Additionally, the expression of integrins α2β1, α3β1, α6β4, and αvβ5 is influenced by zinc, thereby regulating keratinization and keratinocyte migration. In normal skin, these integrins are predominantly expressed in the basal layer and play a crucial role in intercellular and cell-basement membrane adhesion [320]. However, their expression is altered in response to inflammation or tissue injury. The addition of supplementary zinc promotes the induction of α2, α3, αv, and α6 integrin subunits, which in turn affect keratinocyte motility during the healing phase [320]. Accordingly, doses and releases are carefully optimized to ensure therapeutic benefit while avoiding any adverse health effects associated with overdose.
Zn shows promise as a material for manufacturing medical implants. The cytotoxicity of zinc has been investigated in previous studies on various human and animal cell types to confirm this biocompatibility [321,322,323,324,325]. Wu et al. [326] reported a 70% viability of human endometrial epithelial cells when exposed to 150 μmol/L Zn. Conversely, rat retinal cells showed a similar reduction in viability at a lower concentration of 50 μmol/l Zn [327]. Human proximal tubular cells exhibited a viability of less than 50% when exposed to a 100 μmol/L Zn solution [328]. In a recent study by Cheng et al. [329], it was found that 1 μg/mL (equivalent to 15 μM) of Zn did not exhibit toxicity to ECV304 cells, but it decreased the viability of L929 cells in the same environment.
Initial experiments have indicated that Zn-based medical implants pose minimal risk of toxic side effects, as their absorption rate remains well below the tolerated threshold of approximately 15 mg/day due to the moderate degradation rate of Zn [321,323,325]. Zn-based implants exhibit sufficient structural longevity in the body, primarily attributed to the formation of a protective layer of Zn oxide (Zn(OH2) and ZnO) on the implant surface [330]. This protective layer resists corrosion during the initial 3 months following implantation, while corrosion rates tend to accelerate around 4.5 to 6 months as the layer thickness increases [325]. The degradation behavior of Zn demonstrates favorable mechanical integrity during the early stages of implantation (4–6 months) and suitable absorption periods (1–2 years) after the healing process [325,331]. Moreover, zinc in vivo experiments were conducted by Bowen et al. [325] and Li et al. [332], demonstrating significant advancements in this area. Bowen et al. conducted a study where they implanted thin zinc wires into the abdominal aorta of adult rats to examine corrosion rates, corrosion products, and tissue adherence. The results revealed a corrosion rate ranging from 10 to 50 μm/year, which exhibited a progressive increase over a six-month period. Additionally, it was noteworthy that the implant remained structurally intact for a minimum of four months following implantation [325,332].
Beyond its role in nutrition and biocompatibility, zinc has drawn interest in the field of biomaterials because of its natural antibacterial characteristics [333]. There are numerous mechanisms through which zinc exerts its antimicrobial effects. One important process includes rupturing bacterial cell membranes, which can cause cell death and the release of cellular contents. Furthermore, zinc ions can enter bacterial cells and disrupt vital biological functions, including protein synthesis and DNA replication [334,335]. Zinc is effective against a broad spectrum of bacteria, including both Gram-positive and Gram-negative strains [336]. The antibacterial properties of zinc not only defend against infections but also help biomaterial biocompatibility [337]. Improved patient outcomes and lower medical expenses can result from decreased infection rates and related consequences.

Zinc and Its Alloys for Medical Applications

Zinc and its alloys possess specific properties and degradation characteristics that offer potential solutions to the challenges limiting the broader use of magnesium (Mg) and iron (Fe)-based alloys for biodegradable implants. The corrosion rate of pure Zn falls within the range between pure Mg and Fe while also avoiding the release of hydrogen gas during in-vivo degradation processes [268]. Zn-based implants are demonstrated to sustain their mechanical integrity for the initial six months post-implantation. Furthermore, stress accelerates localized corrosion, leading to faster degradation during the recovery phase [338]. Furthermore, the degradation products that arise from the corrosion of zinc exhibit a compact nature and are biocompatible. Additionally, Zn possesses a comparably low melting point of approximately 419.75 °C and exhibits low chemical reactivity. These characteristics simplify the manufacturing processes involved in producing Zn-based products, including casting and thermo-mechanical processing, in comparison to other biodegradable and conventional permanent implant metals [339]. Nonetheless, the widespread use of Zn-based biodegradable materials is hindered by their inadequate mechanical properties, which encompass insufficient mechanical strength and ductility [324,340,341]. Previous studies have explored the implementation of thermo-mechanical processing techniques to enhance the mechanical properties of Zn [323,340,341]. Pure zinc processed at room temperature (RT) has been found to fall short of meeting the necessary mechanical criteria for load-bearing implants [342]. To enhance the mechanical properties of zinc alloys, different alloying elements have been investigated, including Mg, aluminum (Al), Ti, copper (Cu), calcium (Ca), silver (Ag), strontium (Sr), and manganese (Mn). These alloying elements aim to improve the mechanical properties of Zn alloys through mechanisms such as grain boundary strengthening, solid solution strengthening, dispersion strengthening, and precipitation strengthening [324,342,343].
Copper is a favorable choice for alloying with zinc due to its relatively high solubility in Zn at the melting temperature, reaching up to 2 wt%. The improved mechanical properties of Zn–Cu alloys stem from the effects of solid solution strengthening and grain boundary strengthening exerted by Cu [344]. When the Cu content in Zn–Cu alloys is below 1%, they exhibit a nearly single-phase microstructure. As the Cu content increases, the ductility of Zn–Cu alloys is further enhanced, although the increase in mechanical strength becomes relatively minor [345]. Extensive research has been conducted to investigate the impact of Mg additions on the mechanical properties of Zn alloys [346]. The mechanical improvements observed in Zn–Mg alloys are primarily attributed to the presence of brittle phases, namely Mg2Zn11 and MgZn2 eutectic phases, as the solubility of Mg in Zn is low (0.1 wt% at temperatures above 300 ℃ and nearly zero at RT). Notably, Zn alloys containing small amounts of Mg (<1%) exhibit an exceptional balance of strength and plasticity compared to other investigated Zn alloys [325]. However, the conventional casting and processing methods used for Zn alloyed with low Mg content (<1%) are unable to achieve the mechanical properties required for load-bearing biodegradable implants [347,348]. Hence, the processing of Zn–Mg alloys at room temperature presents a promising approach for manufacturing Zn–Mg products with enhanced mechanical properties.

6.3.3. Zn–Mg Alloys

Zn alloyed with Mg has emerged as a promising material for biomedical applications. It is crucial to adjust the ratio between these ions to achieve targeted tissue responses and improve the healing process. There are various methods for adjusting the Mg and Zn ion ratios in biomaterials for various clinical uses [333,349]. Ion release rates must be carefully controlled in order to produce the intended therapeutic benefits while minimizing potential cytotoxicity or unfavorable tissue reactions [350]. Mg and Zn ion release kinetics can be effectively altered by adjusting various factors, including material composition, surface coatings, and fabrication processes [351]. For instance, a sustained release of both ions over time can be achieved by developing Mg-rich alloys with precisely controlled Zn content. Moreover, depending on the targeted tissue and its regenerating needs, the ideal Mg–Zn ratio can significantly change. A higher Mg content in the implant may be preferred in the field of bone tissue engineering due to the well-known function of Mg in bone growth and mineralization [352]. Zn-rich materials, on the other hand, can be preferred for soft tissue applications where wound-healing and antibacterial capabilities of Zn are useful [201,353]. Moreover, the optimal Mg–Zn ratio can also vary from patient to patient, depending on their age, general health, and underlying illnesses. This personalized method can have the promise of promoting improved tissue healing and clinical outcomes across a wide range of medical applications.
According to Yao et al. [354], Zn alloys containing Mg compositions below 10% exhibit significantly enhanced corrosion properties compared to pure Zn. The enhanced corrosion resistance of the implanted Zn–Mg alloy enables it to retain its mechanical integrity for a sufficient period, facilitating tissue repair. The favorable corrosion resistance is primarily attributed to the presence of intermetallic phases, namely Mg2Zn11 and MgZn2, which facilitate the formation of an electrochemically inert protective film, such as Mg2(OH)2CO3, on the surface of the Zn–Mg alloys [343]. Notably, as-cast Zn–3 wt%Mg alloys have been reported to exhibit a refined nanostructure and favorable corrosion resistance [279]. Additionally, studies have indicated that Zn alloyed with 1 wt% Mg achieves an optimal balance between strength and ductility [355]. Previous research has suggested that higher Mg contents (≤3 wt%) result in increased volume fractions of the eutectic phase, leading to improved strength and hardness but reduced ductility [343,356]. The limited ductility observed in Zn–Mg alloys is considered inadequate for medical implant applications, and the non-uniform breakdown due to preferential corrosion of Mg has hindered the widespread use of these alloys in biomedical applications [357,358]. Therefore, increasing research attention has been focused on Zn alloys containing low amounts of Mg (≤1 wt% Mg) in order to reduce the presence of intermetallic phases while incorporating well-designed processing techniques to enhance mechanical and corrosion properties [323]. Gong et al. [356] conducted a study showing that hot extrusion enhances the uniformity of biodegradation and mechanical properties in Zn–1 wt% Mg alloys. In vitro cytotoxicity tests also confirmed the excellent biocompatibility of the alloy. Additionally, several studies have investigated the mechanical properties, corrosion properties, and cytotoxicity of Zn–1 wt% Mg alloys alone or in combination with other elements such as Mn, Ca, and Sr, employing various fabrication techniques [343,356,359].
Li et al. [332] demonstrated the high viability of ECV304, VSCM, and MG63 cells when exposed to extracts from extruded Zn–1 Mg alloy. Additionally, they found that ECV304 and MG63 cells exhibited healthy behavior when directly cultured on the surface of Zn–1 Mg alloy for 24 h. In contrast, VSCM cells displayed an unhealthy and deceased morphology after 24 h on the surface of Zn–1 Mg alloys [332]. Furthermore, Gong et al. observed excellent viability of L-929 cells in diluted extracts (1:15) prepared with DMEM and 10% FBS after 24 and 72 h [356]. Initial in vitro cytotoxicity assessments were conducted on as-cast Zn alloys, and the results demonstrated that U-2 OS cells exposed to the extract from Zn–0.8 Mg alloys containing 70 μmol/L Zn exhibited a viability of 80% [360]. Murni et al. [361] conducted a study on the cytotoxicity of Zn–3 Mg alloy. Extracts were prepared by incubating 0.75 mg of Mg–3Zn powder in cell culture for 72 h, and the resulting extract from Zn–3 Mg contained 0.49 ppm of Zn ions. Li et al. [332] conducted a study affirming that the utilization of Zn–1 Mg alloys in mouse femora does not adversely impact the well-being of the mice. In fact, the study observed a strong development of new bone throughout the process [362].
In another study, the implantation of Mg and Zn ions was utilized to enhance the soft tissue sealing ability of Ti [363]. The introduction of these ions resulted in changes in the surface wettability of titanium, impacting the adsorption of proteins on the sample surfaces. Characterization of the physicochemical and biological properties of the ion-implanted samples revealed that both Mg and Zn ions facilitated the accumulation of ECM components such as collagen-I and fibronectin and improved cell adhesion, migration, and proliferation of HGFs [363]. Specifically, the release of Mg2+ from Mg ion-implanted samples promoted improved adhesion and motility of human gingival fibroblasts (HGFs), potentially through the regulation of ITGB1 expression and activation of the MAPK signaling pathway [363]. On the other hand, the release of Zn2+ resulting from Zn ion implantation primarily enhanced the proliferation of HGFs, which could be attributed to the upregulation of ZIP7 and ZIP13 expression and activation of the TGF-β signaling pathway [363].

7. Experimental Studies of Magnesium and Zinc in Soft Tissue Engineering

Magnesium and zinc exhibit distinctive characteristics that make them highly appealing for biomedical applications, including soft tissue engineering and skin expansion. The role of zinc in wound healing has been unequivocally demonstrated in several studies [201,364]. Topical zinc therapy has been shown to effectively reduce wound debris and promote epithelialization in surgical wounds in rat models [365]. Observations of reduced wound debris and necrotic material following topical zinc application in wounds of various origins have led researchers to investigate the action of zinc-dependent MMPs in cultured necrotic tissue from porcine wounds [366]. In vitro experiments using zinc oxide have shown that it enhances the enzymatic breakdown of collagen fragments through the activity of MMPs, which exhibit substrate specificity for various ECM molecules [367,368]. Additionally, locally applied zinc oxide has been found to enhance the repair of ulcerated skin [369]. On the other hand, blocking MMPs has been demonstrated to considerably prolong the wound-healing process [370]. These findings highlight the role of zinc-dependent MMPs in promoting the breakdown of collagen fragments and the repair of damaged skin [371]. In another study, the administration of zinc via intraperitoneal injection immediately after surgery and daily for 4 days (at a dose of 2 mg/kg/day) was found to increase the bursting pressure of colon anastomoses on the seventh day after surgery in both normal rabbits and rabbits treated with a chemotherapeutic agent. Furthermore, the zinc-treated rabbits exhibited increased infiltration of fibroblasts and enhanced epithelialization [298]. However, when these beneficial effects were applied using intraperitoneal zinc sulfate on colon anastomosis repair in a rat model, different results were observed either on the third or seventh day after surgery [372]. In contrast, the research on the role of magnesium in soft tissue engineering and wound healing has been limited. However, new targeted interventions can be investigated by gaining a deeper understanding of the mechanisms underlying the effects of both zinc and magnesium on wound healing and soft tissue regeneration. These interventions have the potential to harness the benefits of these metals, facilitating the healing process and enhancing clinical outcomes.

8. Conclusions

Medical implants are surgically inserted into the body to enhance human life by either preserving or restoring functionality in damaged tissues. Nevertheless, the long-term presence of foreign implant materials in the body can lead to persistent detrimental and inflammatory reactions, necessitating secondary surgical procedures for their removal. This introduces additional risks to patients and significantly increases costs. In order to address the limitations of permanent implants, there has been significant research focused on biodegradable implants that are designed to gradually degrade within the body. These biodegradable implants must exhibit excellent biocompatibility, meaning they should not trigger inflammatory responses and must be non-toxic. Furthermore, the implant must effectively fulfill its intended function throughout the entire recovery period. Nowadays, the majority of implants are permanent and comprised of metallic components. Therefore, the base metals, alloy elements, and corrosion products formed during degradation must be non-toxic.
In the realm of soft tissue engineering and skin expansion, two elements, magnesium, and zinc, have emerged as particularly promising candidates. These two elements exhibit distinctive characteristics that render them appealing for biomedical applications. However, despite their potential, they have not been extensively explored in these specific areas of research.
Magnesium exhibits excellent biocompatibility and biodegradability, making it an ideal candidate for medical implants. The biodegradable characteristic of Mg enables gradual decomposition, eliminating the necessity for surgical removal of the implant. Magnesium alloys have shown promising results in orthopedic procedures and vascular interventions. They possess mechanical properties similar to cortical bone, ensuring adequate support and stability during the healing process. Furthermore, magnesium has been found to stimulate collagen synthesis, inhibit fibrotic processes, maintain the extensibility of elastin, and regulate integrin activity. These properties indicate its potential role in promoting tissue regeneration and wound healing.
Similarly, zinc plays a crucial role in various biological processes and is essential for cellular functions such as proliferation, migration, and maturation. It acts as a stabilizer of cell membranes and serves as a cofactor for numerous enzymes. Zinc also influences the expression of zinc-finger transcription factors involved in the coding of growth factors essential for wound healing. Moreover, zinc is involved in the modulation of integrin expression, which affects keratinocyte migration and intercellular adhesion. These attributes highlight the potential of zinc in promoting cell behaviors relevant to tissue engineering and skin expansion.
Despite these promising properties, the specific application of magnesium and zinc in soft tissue engineering and skin expansion has not received extensive research attention. Comprehensive studies examining their effects on cell behavior, tissue regeneration, and the development of suitable scaffolds for controlled delivery are necessary. By addressing these research gaps, magnesium and zinc can be further harnessed as valuable resources for advancing the field of soft tissue engineering and achieving successful skin expansion. Several intriguing directions for future study and development emerge. This includes the potential to improve implant performance, and biocompatibility lies in the ongoing improvement of implant designs, including surface changes and composite materials. It will also be critical to fine-tune the degradation kinetics of magnesium and zinc implants. Additionally, investigating the incorporation of therapeutic substances, such as growth factors or antimicrobial coatings, into implant materials can result in multifunctional implants that can respond to certain clinical needs, such as infection prevention or accelerated wound healing. The field of biodegradable magnesium and zinc implants will advance with the inclusion of these perspectives in future research, ultimately helping patients through improved tissue regeneration, decreased problems, and improved overall quality of care.
In conclusion, this review on the use of magnesium and zinc in skin expansion and soft tissue engineering shows great potential. These substances display a variety of beneficial properties, including biocompatibility, biodegradability, and cellular stimulation. To fully realize their potential in these particular fields, additional comprehensive and in-depth studies are still needed. These initiatives have the potential to transform the field of soft tissue engineering and offer safer and better ways to expand the skin and regenerate tissue.

Author Contributions

Conceptualization, N.H. and N.K.; writing—original draft preparation, N.H. and N.K.; writing—review and editing, N.H., N.K., K.S., T.K. and M.Z.; visualization, N.H. and N.K.; supervision, N.K.; funding acquisition, N.K. and N.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the European Union’s Horizon Europe Research and Innovation program (project no. 101047008). HORIZON-EIC-2021-PATHFINDEROPEN-01-01—HORIZON-EIC HORIZON EIC Grants.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ingber, D. Integrins as mechanochemical transducers. Curr. Opin. Cell Biol. 1991, 3, 841–848. [Google Scholar] [CrossRef]
  2. Davidson, J.M.; Aquino, A.M.; Woodward, S.C.; Wilfinger, W.W. Sustained microgravity reduces intrinsic wound healing and growth factor responses in the rat. FASEB J. 1999, 13, 325–329. [Google Scholar] [CrossRef] [PubMed]
  3. Jhala, D.V.; Kale, R.K.; Singh, R.P. Microgravity Alters Cancer Growth and Progression. Curr. Cancer Drug Targets 2014, 14, 394–406. [Google Scholar] [CrossRef] [PubMed]
  4. Farahani, R.M.; A DiPietro, L. Microgravity and the implications for wound healing. Int. Wound J. 2008, 5, 552–561. [Google Scholar] [CrossRef]
  5. Evans, N.D.; Oreffo, R.; Healy, E.; Thurner, P.; Man, Y. Epithelial mechanobiology, skin wound healing, and the stem cell niche. J. Mech. Behav. Biomed. 2013, 28, 397–409. [Google Scholar] [CrossRef]
  6. Silver, F.H.; Siperko, L.M.; Seehra, G.P. Mechanobiology of force transduction in dermal tissue. Ski. Res. Technol. 2003, 9, 3–23. [Google Scholar] [CrossRef] [PubMed]
  7. Tranquillo, R.T.; Durrani, M.A.; Moon, A.G. Tissue engineering science: Consequences of cell traction force. Cytotechnology 1992, 10, 225–250. [Google Scholar] [CrossRef]
  8. Takei, T.; Mills, I.; Arai, K.; Sumpio, B.E. Molecular Basis for Tissue Expansion: Clinical Implications for the Surgeon. Plast. Reconstr. Surg. 1998, 102, 247–258. [Google Scholar] [CrossRef]
  9. Huang, S.; Ingber, D.E. The structural and mechanical complexity of cell-growth control. Nature 1999, 1, E131–E138. [Google Scholar] [CrossRef]
  10. Wang, J.H.-C.; Thampatty, B.P. An Introductory Review of Cell Mechanobiology. Biomech. Model. Mechanobiol. 2006, 5, 1–16. [Google Scholar] [CrossRef]
  11. Huang, S.; Ingber, D.E. Shape-Dependent Control of Cell Growth, Differentiation, and Apoptosis: Switching between Attractors in Cell Regulatory Networks. Exp. Cell Res. 2000, 261, 91–103. [Google Scholar] [CrossRef] [PubMed]
  12. Herndon, D.N.; Barrow, R.E.; Rutan, R.L.; Rutan, T.C.; Desai, M.H.; Abston, S. A Comparison of Conservative Versus Early Excision. Therapies in severely burned patients. Ann. Surg. 1989, 209, 547–553; discussion 552–553. [Google Scholar] [CrossRef] [PubMed]
  13. Langer, R.; Vacanti, J. Tissue engineering. Science 1993, 260, 920–926. [Google Scholar] [CrossRef] [PubMed]
  14. Karp, J.M.; Langer, R. Development and therapeutic applications of advanced biomaterials. Curr. Opin. Biotechnol. 2007, 18, 454–459. [Google Scholar] [CrossRef]
  15. Ikada, Y. Challenges in tissue engineering. J. R. Soc. Interface 2006, 3, 589–601. [Google Scholar] [CrossRef]
  16. Tepole, A.B.; Ploch, C.J.; Wong, J.; Gosain, A.K.; Kuhl, E. Growing skin: A computational model for skin expansion in reconstructive surgery. J. Mech. Phys. Solids 2011, 59, 2177–2190. [Google Scholar] [CrossRef]
  17. Neumann, C.G. The expansion of an area of skin by progressive distention of a subcutaneous balloon; use of the method for securing skin for subtotal reconstruction of the ear. Plast. Reconstr. Surg. 1957, 19, 124–130. [Google Scholar] [CrossRef]
  18. Wagh, M.; Dixit, V. Tissue expansion: Concepts, techniques and unfavourable results. Indian J. Plast. Surg. 2013, 46, 333–348. [Google Scholar] [CrossRef]
  19. Langer, R.; Tirrell, D.A. Designing materials for biology and medicine. Nature 2004, 428, 487–492. [Google Scholar] [CrossRef]
  20. Wong, V.W.; Longaker, M.T.; Gurtner, G.C. Soft tissue mechanotransduction in wound healing and fibrosis. Semin. Cell Dev. Biol. 2012, 23, 981–986. [Google Scholar] [CrossRef]
  21. Zöllner, A.M.; Tepole, A.B.; Kuhl, E. On the biomechanics and mechanobiology of growing skin. J. Theor. Biol. 2012, 297, 166–175. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, J.H.-C.; Thampatty, B.P.; Lin, J.-S.; Im, H.-J. Mechanoregulation of gene expression in fibroblasts. Gene 2007, 391, 1–15. [Google Scholar] [CrossRef]
  23. Derderian, C.A.; Bastidas, N.; Lerman, O.Z.; Bhatt, K.A.; Lin, S.-E.; Voss, J.; Holmes, J.W.; Levine, J.P.; Gurtner, G.C. Mechanical Strain Alters Gene Expression in an in Vitro Model of Hypertrophic Scarring. Ann. Plast. Surg. 2005, 55, 69–75; discussion 75. [Google Scholar] [CrossRef] [PubMed]
  24. Silver, F.H.; Siperko, L.M. Mechanosensing and Mechanochemical Transduction: How Is Mechanical Energy Sensed and Converted Into Chemical Energy in an Extracellular Matrix? Crit. Rev. Biomed. Eng. 2003, 31, 78. [Google Scholar] [CrossRef]
  25. Wong, V.; Levi, K.; Akaishi, S.; Schultz, G.; Dauskardt, R. Scar zones: Region-specific differences in skin tension may determine incisional scar formation. Plast. Reconstr. Surg. 2012, 129, 1272–1276. [Google Scholar] [CrossRef]
  26. Montesano, R.; Orci, L. Transforming growth factor beta stimulates collagen-matrix contraction by fibroblasts: Implications for wound healing. Proc. Natl. Acad. Sci. USA 1988, 85, 4894–4897. [Google Scholar] [CrossRef] [PubMed]
  27. Halfter, W.; Liverani, D.; Vigny, M.; Monard, D. Deposition of extracellular matrix along the pathways of migrating fibroblasts. Cell Tissue Res. 1990, 262, 467–481. [Google Scholar] [CrossRef] [PubMed]
  28. Sander, E.A.; Barocas, V.H.; Tranquillo, R.T. Initial Fiber Alignment Pattern Alters Extracellular Matrix Synthesis in Fibroblast-Populated Fibrin Gel Cruciforms and Correlates with Predicted Tension. Ann. Biomed. Eng. 2010, 39, 714–729. [Google Scholar] [CrossRef]
  29. Wong, V.W.; Akaishi, S.; Longaker, M.T.; Gurtner, G.C. Pushing Back: Wound Mechanotransduction in Repair and Regeneration. J. Investig. Dermatol. 2011, 131, 2186–2196. [Google Scholar] [CrossRef]
  30. Ko, K.S.; McCulloch, C.A. Intercellular Mechanotransduction: Cellular Circuits That Coordinate Tissue Responses to Mechanical Loading. Biochem. Biophys. Res. Commun. 2001, 285, 1077–1083. [Google Scholar] [CrossRef]
  31. Geffeney, S.L.; Goodman, M.B. How We Feel: Ion Channel Partnerships that Detect Mechanical Inputs and Give Rise to Touch and Pain Perception. Neuron 2012, 74, 609–619. [Google Scholar] [CrossRef] [PubMed]
  32. Powell, H.M.; McFarland, K.L.; Butler, D.L.; Supp, D.M.; Boyce, S.T. Uniaxial Strain Regulates Morphogenesis, Gene Expression, and Tissue Strength in Engineered Skin. Tissue Eng. Part A 2010, 16, 1083–1092. [Google Scholar] [CrossRef] [PubMed]
  33. De Filippo, R.E.; Atala, A. Stretch and Growth: The Molecular and Physiologic Influences of Tissue Expansion. Plast. Reconstr. Surg. 2002, 109, 2450–2462. [Google Scholar] [CrossRef] [PubMed]
  34. Tepole, A.B.; Gart, M.; Gosain, A.K.; Kuhl, E. Characterization of living skin using multi-view stereo and isogeometric analysis. Acta Biomater. 2014, 10, 4822–4831. [Google Scholar] [CrossRef] [PubMed]
  35. Wilhelmi, B.J.; Blackwell, S.J.; Mancoll, J.S.; Phillips, L.G. Creep vs. Stretch: A Review of the Viscoelastic Properties of Skin. Ann. Plast. Surg. 1998, 41, 215–219. [Google Scholar] [CrossRef]
  36. LoGiudice, J.; Gosain, A.K. Pediatric Tissue Expansion: Indications and Complications. J. Craniofacial Surg. 2003, 14, 866–872. [Google Scholar] [CrossRef] [PubMed]
  37. Filho, P.T.B.; Neves, R.I.; Gemperli, R.; Kaweski, S.; Kahler, S.H.; Banducci, D.R.; Manders, E.K. Soft-Tissue Expansion in Lower Extremity Reconstruction. Clin. Plast. Surg. 1991, 18, 593–599. [Google Scholar] [CrossRef]
  38. Beauchene, J.; Chambers, M.; Peterson, A.; Scott, P. Biochemical, biomechanical, and physical changes in the skin in an experimental animal model of therapeutic tissue expansion. J. Surg. Res. 1989, 47, 507–514. [Google Scholar] [CrossRef]
  39. Pietramaggiori, G.; Liu, P.; Scherer, S.S.; Kaipainen, A.; Prsa, M.J.; Mayer, H.; Newalder, J.; Alperovich, M.; Mentzer, S.J.; Konerding, M.A.; et al. Tensile Forces Stimulate Vascular Remodeling and Epidermal Cell Proliferation in Living Skin. Ann. Surg. 2007, 246, 896–902. [Google Scholar] [CrossRef]
  40. Purnell, C.A.; Gart, M.S.; Buganza-Tepole, A.; Tomaszewski, J.P.; Topczewska, J.M.; Kuhl, E.; Gosain, A.K. Determining the Differential Effects of Stretch and Growth in Tissue-Expanded Skin: Combining Isogeometric Analysis and Continuum Mechanics in a Porcine Model. Dermatol. Surg. 2018, 44, 48–52. [Google Scholar] [CrossRef]
  41. Huang, C.; Akaishi, S.; Ogawa, R. Mechanosignaling pathways in cutaneous scarring. Arch. Dermatol. Res. 2012, 304, 589–597. [Google Scholar] [CrossRef] [PubMed]
  42. Hinz, B.; Gabbiani, G. Mechanisms of force generation and transmission by myofibroblasts. Curr. Opin. Biotechnol. 2003, 14, 538–546. [Google Scholar] [CrossRef] [PubMed]
  43. 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]
  44. Ismavel, R.; Samuel, S.; Boopalan, P.R.J.V.C.; Chittaranjan, S.B. A Simple Solution for Wound Coverage by Skin Stretching. J. Orthop. Trauma 2011, 25, 127–132. [Google Scholar] [CrossRef]
  45. Lancerotto, L.; Chin, M.S.; Freniere, B.; Lujan-Hernandez, J.R.; Li, Q.; Vasquez, A.V.; Bassetto, F.; Del Vecchio, D.A.; Lalikos, J.F.; Orgill, D.P. Mechanisms of Action of External Volume Expansion Devices. Plast. Reconstr. Surg. 2013, 132, 569–578. [Google Scholar] [CrossRef]
  46. Ingber, D.E. Tensegrity: The architectural basis of cellular mechanotransduction. Annu. Rev. Physiol. 1997, 59, 575–599. [Google Scholar] [CrossRef]
  47. Orgill, D.P.; Bayer, L. Negative pressure wound therapy: Past, present and future. Int. Wound J. 2013, 10 (Suppl. S1), 15–19. [Google Scholar] [CrossRef]
  48. Lancerotto, L.; Bayer, L.R.; Orgill, D.P. Mechanisms of action of microdeformational wound therapy. Semin. Cell Dev. Biol. 2012, 23, 987–992. [Google Scholar] [CrossRef]
  49. Omar, M.T.; Alghadir, A.; Al-Wahhabi, K.K.; Al-Askar, A.B. Efficacy of shock wave therapy on chronic diabetic foot ulcer: A single-blinded randomized controlled clinical trial. Diabetes Res. Clin. Pract. 2014, 106, 548–554. [Google Scholar] [CrossRef]
  50. Ottomann, C.; Stojadinovic, A.; Lavin, P.T.; Gannon, F.H.; Heggeness, M.H.; Thiele, R.; Schaden, W.; Hartmann, B. Prospective Randomized Phase II Trial of Accelerated Reepithelialization of Superficial Second-Degree Burn Wounds Using Extracorporeal Shock Wave Therapy. Ann. Surg. 2012, 255, 23–29. [Google Scholar] [CrossRef]
  51. Ennis, W.J.; Foremann, P.; Mozen, N.; Massey, J.; Conner-Kerr, T.; Meneses, P. Ultrasound therapy for recalcitrant diabetic foot ulcers: Results of a randomized, double-blind, controlled, multicenter study. Ostomy Wound Manag. 2005, 51, 24–39. [Google Scholar]
  52. Kloth, L.C. Electrical Stimulation for Wound Healing: A Review of Evidence from In Vitro Studies, Animal Experiments, and Clinical Trials. Int. J. Low. Extrem. Wounds 2005, 4, 23–44. [Google Scholar] [CrossRef] [PubMed]
  53. El-Sabbagh, A.H. Negative pressure wound therapy: An update. Chin. J. Traumatol. 2017, 20, 103–107. [Google Scholar] [CrossRef]
  54. Wiegand, C.; White, R. Microdeformation in wound healing. Wound Repair Regen. 2013, 21, 793–799. [Google Scholar] [CrossRef]
  55. Nuutila, K.; Siltanen, A.; Peura, M.; Harjula, A.; Nieminen, T.; Vuola, J.; Kankuri, E.; Aarnio, P. Gene expression profiling of negative-pressure-treated skin graft donor site wounds. Burns 2013, 39, 687–693. [Google Scholar] [CrossRef] [PubMed]
  56. McNulty, A.K.; Schmidt, M.; Feeley, T.; Kieswetter, K. Effects of negative pressure wound therapy on fibroblast viability, chemotactic signaling, and proliferation in a provisional wound (fibrin) matrix. Wound Repair Regen. 2007, 15, 838–846. [Google Scholar] [CrossRef] [PubMed]
  57. Greene, A.K.; Puder, M.; Roy, R.; Arsenault, D.; Kwei, S.; Moses, M.; Orgill, D. Microdeformational wound therapy: Effects on angiogenesis and matrix metalloproteinases in chronic wounds of 3 debilitated patients. Ann. Plast. Surg. 2006, 56, 418–422. [Google Scholar] [CrossRef]
  58. Lu, F.; Ogawa, R.; Nguyen, D.T.; Chen, B.; Guo, D.; Helm, D.L.; Zhan, Q.; Murphy, G.F.; Orgill, D.P. Microdeformation of Three-Dimensional Cultured Fibroblasts Induces Gene Expression and Morphological Changes. Ann. Plast. Surg. 2011, 66, 296–300. [Google Scholar] [CrossRef]
  59. Junker, J.P.; Kamel, R.A.; Caterson, E.; Eriksson, E.; Nuutila, K.; Patil, P.S.; Fathollahipour, S.; Inmann, A.; Pant, A.; Amini, R.; et al. Clinical Impact Upon Wound Healing and Inflammation in Moist, Wet, and Dry Environments. Adv. Wound Care 2013, 2, 348–356. [Google Scholar] [CrossRef]
  60. Walmsley, G.G.; Maan, Z.N.; Wong, V.W.; Duscher, D.; Hu, M.S.; Zielins, E.R.; Wearda, T.; Muhonen, E.; McArdle, A.; Tevlin, R.; et al. Scarless wound healing: Chasing the holy grail. Plast. Reconstr. Surg. 2015, 135, 907–917. [Google Scholar] [CrossRef]
  61. Ud-Din, S.; Volk, S.W.; Bayat, A. Regenerative healing, scar-free healing and scar formation across the species: Current concepts and future perspectives. Exp. Dermatol. 2014, 23, 615–619. [Google Scholar] [CrossRef] [PubMed]
  62. Nauta, A.; Gurtner, G.; Longaker, M. Wound healing and regenerative strategies. Oral Dis. 2011, 17, 541–549. [Google Scholar] [CrossRef] [PubMed]
  63. Reinke, J.M.; Sorg, H. Wound repair and regeneration. Eur. Surg. Res. 2012, 49, 35–43. [Google Scholar] [CrossRef] [PubMed]
  64. Huang, C.; Holfeld, J.; Schaden, W.; Orgill, D.; Ogawa, R. Mechanotherapy: Revisiting physical therapy and recruiting mechanobiology for a new era in medicine. Trends Mol. Med. 2013, 19, 555–564. [Google Scholar] [CrossRef] [PubMed]
  65. Akaishi, S.; Akimoto, M.; Hyakusoku, H.; Ogawa, R. 142B: The relationship between keloid growth pattern and stretching tension-visual analysis using the finite element method. Ann Plast. Surg. 2008, 60, 445–451. [Google Scholar] [CrossRef] [PubMed]
  66. Akaishi, S.; Ogawa, R.; Hyakusoku, H. Keloid and hypertrophic scar: Neurogenic inflammation hypotheses. Med. Hypotheses 2008, 71, 32–38. [Google Scholar] [CrossRef]
  67. Ogawa, R.; Okai, K.; Tokumura, F.; Mori, K.; Ohmori, Y.; Huang, C.; Hyakusoku, H.; Akaishi, S. The relationship between skin stretching/contraction and pathologic scarring: The important role of mechanical forces in keloid generation. Wound Repair Regen. 2012, 20, 149–157. [Google Scholar] [CrossRef]
  68. Ogawa, R. Mechanobiology of scarring. Wound Repair Regen. 2011, 19 (Suppl. S1), s2–s9. [Google Scholar] [CrossRef]
  69. Ogawa, R.; Akaishi, S.; Kuribayashi, S.; Miyashita, T. Keloids and Hypertrophic Scars Can Now Be Cured Completely: Recent Progress in Our Understanding of the Pathogenesis of Keloids and Hypertrophic Scars and the Most Promising Current Therapeutic Strategy. J. Nippon. Med. Sch. 2016, 83, 46–53. [Google Scholar] [CrossRef]
  70. Metcalfe, A.D.; Ferguson, M.W. Tissue engineering of replacement skin: The crossroads of biomaterials, wound healing, embryonic development, stem cells and regeneration. J. R. Soc. Interface 2006, 4, 413–437. [Google Scholar] [CrossRef]
  71. Rajendran, A.K.; Sankar, D.; Amirthalingam, S.; Kim, H.D.; Rangasamy, J.; Hwang, N.S. Trends in mechanobiology guided tissue engineering and tools to study cell-substrate interactions: A brief review. Biomater. Res. 2023, 27, 1–24. [Google Scholar] [CrossRef] [PubMed]
  72. Butler, D.L.; Goldstein, S.A.; Guldberg, R.E.; Guo, X.E.; Kamm, R.; Laurencin, C.T.; McIntire, L.V.; Mow, V.C.; Nerem, R.M.; Sah, R.L.; et al. The Impact of Biomechanics in Tissue Engineering and Regenerative Medicine. Tissue Eng. Part B Rev. 2009, 15, 477–484. [Google Scholar] [CrossRef] [PubMed]
  73. Huang, C.-Y.; Mow, V.C.; Ateshian, G.A. The Role of Flow-Independent Viscoelasticity in the Biphasic Tensile and Compressive Responses of Articular Cartilage. J. Biomech. Eng. 2001, 123, 410–417. [Google Scholar] [CrossRef] [PubMed]
  74. Walker, M.; Godin, M.; Harden, J.L.; Pelling, A.E. Time dependent stress relaxation and recovery in mechanically strained 3D microtissues. APL Bioeng. 2020, 4, 036107. [Google Scholar] [CrossRef] [PubMed]
  75. Onal, S.; Alkaisi, M.M.; Nock, V. Microdevice-based mechanical compression on living cells. iScience 2022, 25, 105518. [Google Scholar] [CrossRef]
  76. Ning, L.; Gil, C.J.; Hwang, B.; Theus, A.S.; Perez, L.; Tomov, M.L.; Bauser-Heaton, H.; Serpooshan, V. Biomechanical factors in three-dimensional tissue bioprinting. Appl. Phys. Rev. 2020, 7, 041319. [Google Scholar] [CrossRef]
  77. Kaner, D.; Friedmann, A. Soft tissue expansion with self-filling osmotic tissue expanders before vertical ridge augmentation: A proof of principle study. J. Clin. Periodontol. 2010, 38, 95–101. [Google Scholar] [CrossRef]
  78. Mertens, C.; Thiele, O.; Engel, M.; Seeberger, R.; Hoffmann, J.; Freier, K. The Use of Self-Inflating Soft Tissue Expanders Prior to Bone Augmentation of Atrophied Alveolar Ridges. Clin. Implant. Dent. Relat. Res. 2013, 17, 44–51. [Google Scholar] [CrossRef]
  79. Johnson, T.M.; Lowe, L.; Brown, M.D.; Sullivan, M.J.; Nelson, B.R. Histology and Physiology of Tissue Expansion. J. Dermatol. Surg. Oncol. 1993, 19, 1074–1078. [Google Scholar] [CrossRef]
  80. Rivera, R.; LoGiudice, J.; Gosain, A.K. Tissue expansion in pediatric patients. Clin. Plast. Surg. 2005, 32, 35–44. [Google Scholar] [CrossRef]
  81. Uijlenbroek, H.J.J.; Liu, Y.; He, J.F.; Visscher, C.; van Waas, M.A.J.; Wismeyer, D. Expanding soft tissue with Osmed® tissue expanders in the goat maxilla. Clin. Oral Implant. Res. 2010, 22, 121–128. [Google Scholar] [CrossRef] [PubMed]
  82. Berge, S.J.; Wiese, K.G.; von Lindern, J.J.; Niederhagen, B.; Appel, T.; Reich, R.H. Tissue expansion using osmotically active hydrogel systems for direct closure of the donor defect of the radial forearm flap. Plast. Reconstr. Surg. 2001, 108, 1–5, discussion 6–7. [Google Scholar] [CrossRef] [PubMed]
  83. Ronert, M.A.M.; Hofheinz, H.M.; Manassa, E.M.; Asgarouladi, H.; Olbrisch, R.R.M. The Beginning of a New Era in Tissue Expansion: Self-Filling Osmotic Tissue Expander—Four-Year Clinical Experience. Plast. Reconstr. Surg. 2004, 114, 1025–1031. [Google Scholar] [CrossRef] [PubMed]
  84. Obdeijn, M.C.; Nicolai, J.-P.A.; Werker, P.M. The osmotic tissue expander: A three-year clinical experience. J. Plast. Reconstr. Aesthetic Surg. 2009, 62, 1219–1222. [Google Scholar] [CrossRef] [PubMed]
  85. Chummun, S.; Addison, P.; Stewart, K.J. The osmotic tissue expander: A 5-year experience. J. Plast. Reconstr. Aesthet. Surg. 2010, 63, 2128–2132. [Google Scholar] [CrossRef]
  86. Arneja, J.S.; Gosain, A.K. Giant congenital melanocytic nevi. Plast. Reconstr. Surg. 2009, 124 (Suppl. S1), 1e–13e. [Google Scholar] [CrossRef]
  87. Moustafa, D.; Blundell, A.R.; Hawryluk, E.B. Congenital melanocytic nevi. Curr. Opin. Pediatr. 2020, 32, 491–497. [Google Scholar] [CrossRef]
  88. Formby, P.; Flint, J.; Gordon, W.T.; Fleming, M.; Andersen, R.C. Use of a Continuous External Tissue Expander in the Conversion of a Type IIIB Fracture to a Type IIIA Fracture. Orthopedics 2013, 36, e249–e251. [Google Scholar] [CrossRef]
  89. Radovan, C. Breast Reconstruction after Mastectomy Using the Temporary Expander. Plast. Reconstr. Surg. 1982, 69, 207–208. [Google Scholar] [CrossRef]
  90. Argenta, L.C.M.; Marks, M.W.M.; Grabb, W.C.M. Selective Use of Serial Expansion in Breast Reconstruction. Ann. Plast. Surg. 1983, 11, 188–195. [Google Scholar] [CrossRef]
  91. Brobmann, G.F.; Huber, J. Effects of Different-Shaped Tissue Expanders on Transluminal Pressure, Oxygen Tension, Histopathologic Changes, and Skin Expansion in Pigs. Plast. Reconstr. Surg. 1985, 76, 731–745. [Google Scholar] [CrossRef] [PubMed]
  92. Van Rappard, J.H.; Molenaar, J.; Van Doorn, K.; Sonneveld, G.J.; Borghouts, J.M. Surface-area increase in tissue expansion. Plast. Reconstr. Surg. 1988, 82, 833–839. [Google Scholar] [CrossRef] [PubMed]
  93. Pietila, J.P. Tissue expansion and skin circulation. Simultaneous monitoring by laser Doppler flowmetry and transcutaneous oximetry. Scand. J. Plast. Reconstr. Surg. Hand Surg. 1990, 24, 135–140. [Google Scholar]
  94. Van Damme, P.A.; Heidbüchel, K.L.; Kuijpers-Jagtman, A.-M.; Maltha, J.C.; Freihofer, H.P.M. Cranio-maxillo-facial tissue expansion, experimentally based or clinically empiric? A review of the literature. J. Cranio Maxillofac. Surg. 1992, 20, 61–69. [Google Scholar] [CrossRef]
  95. Austad, E.D.; Rose, G.L. A self-inflating tissue expander. Plast. Reconstr. Surg. 1982, 70, 588–594. [Google Scholar] [CrossRef] [PubMed]
  96. Wiese, K. Osmotically induced tissue expansion with hydrogels: A new dimension in tissue expansion? A preliminary report. J. Cranio-Maxillofac. Surg. 1993, 21, 309–313. [Google Scholar] [CrossRef] [PubMed]
  97. Wiese, K.G.; Vogel, M.; Guthoff, R.; Gundlach, K.K. Treatment of congenital anophthalmos with self-inflating polymer expanders: A new method. J. Cranio-Maxillofac. Surg. 1999, 27, 72–76. [Google Scholar] [CrossRef]
  98. Wiese, K.G.; Heinemann, D.E.H.; Ostermeier, D.; Peters, J.H. Biomaterial properties and biocompatibility in cell culture of a novel self-inflating hydrogel tissue expander. J. Biomed. Mater. Res. 2000, 54, 179–188. [Google Scholar] [CrossRef]
  99. Hoffmann, J.F. Tissue expansion in the head and neck. Facial Plast. Surg. Clin. N. Am. 2005, 13, 315–324, vii. [Google Scholar] [CrossRef]
  100. Downes, R.; Lavin, M.; Collin, R. Hydrophilic expanders for the congenital anophthalmic socket. Adv. Ophthalmic Plast. Reconstr. Surg. 1992, 9, 57–61. [Google Scholar]
  101. Bell, C.L.; Peppas, N.A. Water, solute and protein diffusion in physiologically responsive hydrogels of poly(methacrylic acid-g-ethylene glycol). Biomaterials 1996, 17, 1203–1218. [Google Scholar] [CrossRef] [PubMed]
  102. Varga, J.; Janovak, L.; Varga, E.; Eros, G.; Dekany, I.; Kemeny, L. Acrylamide, Acrylic Acid and N-Isopropylacrylamide Hydrogels as Osmotic Tissue Expanders. Ski. Pharmacol. Physiol. 2009, 22, 305–312. [Google Scholar] [CrossRef] [PubMed]
  103. Freeman, F.; Browe, D.; Nulty, J.; Von Euw, S.; Grayson, W.; Kelly, D. Biofabrication of multiscale bone extracellular matrix scaffolds for bone tissue engineering. Eur. Cells Mater. 2019, 38, 168–187. [Google Scholar] [CrossRef] [PubMed]
  104. Mikos, A.G.; Thorsen, A.J.; Czerwonka, L.A.; Bao, Y.; Langer, R.; Winslow, D.N.; Vacanti, J.P. Preparation and characterization of poly(l-lactic acid) foams. Polymers 1994, 35, 1068–1077. [Google Scholar] [CrossRef]
  105. Haugen, H.; Ried, V.; Brunner, M.; Will, J.; Wintermantel, E. Water as foaming agent for open cell polyurethane structures. J. Mater. Sci. Mater. Med. 2004, 15, 343–346. [Google Scholar] [CrossRef]
  106. Parks, K.L.; Beckman, E.J. Generation of microcellular polyurethane foams via polymerization in carbon dioxide. II: Foam formation and characterization. Polym. Eng. Sci. 1996, 36, 2417–2431. [Google Scholar] [CrossRef]
  107. Lee, K.-W.D.; Chan, P.K.; Feng, X. Morphology development and characterization of the phase-separated structure resulting from the thermal-induced phase separation phenomenon in polymer solutions under a temperature gradient. Chem. Eng. Sci. 2004, 59, 1491–1504. [Google Scholar] [CrossRef]
  108. Thomson, R.C.; Wake, M.C.; Yaszemski, M.J.; Mikos, A.G. Biodegradable polymer scaffolds to regenerate organs. Adv. Polym. Sci. 1995, 122, 245–274. [Google Scholar] [CrossRef]
  109. Liapis, A.; Bruttini, R. A theory for the primary and secondary drying stages of the freeze-drying of pharmaceutical crystalline and amorphous solutes: Comparison between experimental data and theory. Sep. Technol. 1994, 4, 144–155. [Google Scholar] [CrossRef]
  110. Pikal, M.; Shah, S.; Roy, M.; Putman, R. The secondary drying stage of freeze drying: Drying kinetics as a function of temperature and chamber pressure. Int. J. Pharm. 1990, 60, 203–207. [Google Scholar] [CrossRef]
  111. Vergnol, G.; Ginsac, N.; Rivory, P.; Meille, S.; Chenal, J.M.; Balvay, S.; Chevalier, J.; Hartmann, D.J. In vitro and in vivo evaluation of a polylactic acid-bioactive glass composite for bone fixation devices. J. Biomed. Mater. Res. B 2016, 104, 180–191. [Google Scholar] [CrossRef] [PubMed]
  112. Del Bakhshayesh, A.R.; Annabi, N.; Khalilov, R.; Akbarzadeh, A.; Samiei, M.; Alizadeh, E.; Alizadeh-Ghodsi, M.; Davaran, S.; Montaseri, A. Recent advances on biomedical applications of scaffolds in wound healing and dermal tissue engineering. Artif. Cells Nanomed. Biotechnol. 2018, 46, 691–705. [Google Scholar] [CrossRef] [PubMed]
  113. Bracaglia, L.G.; Smith, B.T.; Watson, E.; Arumugasaamy, N.; Mikos, A.G.; Fisher, J.P. 3D printing for the design and fabrication of polymer-based gradient scaffolds. Acta Biomater. 2017, 56, 3–13. [Google Scholar] [CrossRef] [PubMed]
  114. Nyberg, E.L.; Farris, A.L.; Hung, B.P.; Dias, M.; Garcia, J.R.; Dorafshar, A.H.; Grayson, W.L. 3D-Printing Technologies for Craniofacial Rehabilitation, Reconstruction, and Regeneration. Ann. Biomed. Eng. 2016, 45, 45–57. [Google Scholar] [CrossRef]
  115. Lee, M.J.; Kim, S.E.; Park, J.; Ahn, G.Y.; Yun, T.H.; Choi, I.; Kim, H.; Choi, S. Curcumin-loaded biodegradable polyurethane scaffolds modified with gelatin using 3D printing technology for cartilage tissue engineering. Polym. Adv. Technol. 2019, 30, 3083–3090. [Google Scholar] [CrossRef]
  116. Wang, Y.; Kim, H.-J.; Vunjak-Novakovic, G.; Kaplan, D.L. Stem cell-based tissue engineering with silk biomaterials. Biomaterials 2006, 27, 6064–6082. [Google Scholar] [CrossRef]
  117. Rising, A.; Nimmervoll, H.; Grip, S.; Fernandez-Arias, A.; Storckenfeldt, E.; Knight, D.P.; Vollrath, F.; Engström, W. Spider Silk Proteins—Mechanical Property and Gene Sequence. Zool. Sci. 2005, 22, 273–281. [Google Scholar] [CrossRef]
  118. Foo, C.W.P.; Kaplan, D.L. Genetic engineering of fibrous proteins: Spider dragline silk and collagen. Adv. Drug Deliv. Rev. 2002, 54, 1131–1143. [Google Scholar] [CrossRef]
  119. Vollrath, F.; Knight, D.P. Liquid crystalline spinning of spider silk. Nature 2001, 410, 541–548. [Google Scholar] [CrossRef]
  120. Gosline, J.M.; Guerette, P.A.; Ortlepp, C.S.; Savage, K.N. The mechanical design of spider silks: From fibroin sequence to mechanical function. J. Exp. Biol. 1999, 202 Pt 23, 3295–3303. [Google Scholar] [CrossRef]
  121. Jin, H.-J.; Kaplan, D.L. Mechanism of silk processing in insects and spiders. Nature 2003, 424, 1057–1061. [Google Scholar] [CrossRef] [PubMed]
  122. Lazaris, A.; Arcidiacono, S.; Huang, Y.; Zhou, J.F.; Duguay, F.; Chretien, N.; Welhs, E.A.; Soares, J.W.; Karatzas, C.N. High-toughness Spider Silk Fibers Spun from Soluble Recombinant Silk Produced in Mammalian Cells. Science 2002, 295, 472–476. [Google Scholar] [CrossRef] [PubMed]
  123. Winkler, S.; Kaplan, D.L. Molecular biology of spider silk. Rev. Mol. Biotechnol. 2000, 74, 85–93. [Google Scholar] [CrossRef]
  124. Bini, E.; Knight, D.P.; Kaplan, D.L. Mapping Domain Structures in Silks from Insects and Spiders Related to Protein Assembly. J. Mol. Biol. 2003, 335, 27–40. [Google Scholar] [CrossRef]
  125. Simmons, A.H.; Michal, C.A.; Jelinski, L.W. Molecular Orientation and Two-Component Nature of the Crystalline Fraction of Spider Dragline Silk. Science 1996, 271, 84–87. [Google Scholar] [CrossRef] [PubMed]
  126. Prince, J.T.; McGrath, K.P.; DiGirolamo, C.M.; Kaplan, D.L. Construction, Cloning, and Expression of Synthetic Genes Encoding Spider Dragline Silk. Biochemistry 1995, 34, 10879–10885. [Google Scholar] [CrossRef] [PubMed]
  127. Gotoh, Y.; Tsukada, M.; Minoura, N. Effect of the chemical modification of the arginyl residue in Bombyx mori silk fibroin on the attachment and growth of fibroblast cells. J. Biomed. Mater. Res. 1998, 39, 351–357. [Google Scholar] [CrossRef]
  128. Inouye, K.; Kurokawa, M.; Nishikawa, S.; Tsukada, M. Use of Bombyx mori silk fibroin as a substratum for cultivation of animal cells. J. Biochem. Biophys. Methods 1998, 37, 159–164. [Google Scholar] [CrossRef]
  129. Sofia, S.; McCarthy, M.B.; Gronowicz, G.; Kaplan, D.L. Functionalized silk-based biomaterials for bone formation. J. Biomed. Mater. Res. 2000, 54, 139–148. [Google Scholar] [CrossRef]
  130. Chen, J.; Altman, G.H.; Karageorgiou, V.; Horan, R.; Collette, A.; Volloch, V.; Colabro, T.; Kaplan, D.L. Human bone marrow stromal cell and ligament fibroblast responses on RGD-modified silk fibers. J. Biomed. Mater. Res. A 2003, 67, 559–570. [Google Scholar] [CrossRef]
  131. Yang, M. Silk-based biomaterials. Microsc. Res. Tech. 2017, 80, 321–330. [Google Scholar] [CrossRef]
  132. Santin, M.; Motta, A.; Freddi, G.; Cannas, M. In vitro evaluation of the inflammatory potential of the silk fibroin. J. Biomed. Mater. Res. 1999, 46, 382–389. [Google Scholar] [CrossRef]
  133. Sugihara, A.; Sugiura, K.; Morita, H.; Ninagawa, T.; Tubouchi, K.; Tobe, R.; Izumiya, M.; Horio, T.; Abraham, N.G.; Ikehara, S. Promotive effects of a silk film on epidermal recovery from full-thickness skin wounds. Proc. Soc. Exp. Biol. Med. 2000, 225, 58–64. [Google Scholar] [CrossRef]
  134. Meinel, L.; Hofmann, S.; Karageorgiou, V.; Kirker-Head, C.; McCool, J.; Gronowicz, G.; Zichner, L.; Langer, R.; Vunjak-Novakovic, G.; Kaplan, D.L. The inflammatory responses to silk films in vitro and in vivo. Biomaterials 2005, 26, 147–155. [Google Scholar] [CrossRef] [PubMed]
  135. Minoura, N.; Tsukada, M.; Nagura, M. Physico-chemical properties of silk fibroin membrane as a biomaterial. Biomaterials 1990, 11, 430–434. [Google Scholar] [CrossRef] [PubMed]
  136. Chiarini, A.; Petrini, P.; Bozzini, S.; Dal Pra, I.; Armato, U. Silk fibroin/poly(carbonate)-urethane as a substrate for cell growth: In vitro interactions with human cells. Biomaterials 2003, 24, 789–799. [Google Scholar] [CrossRef] [PubMed]
  137. Yeo, J.H.; Lee, K.G.; Kim, H.C.; Oh, Y.L.; Kim, A.-J.; Kim, S.Y. The Effects of PVA/Chitosan/Fibroin (PCF)-Blended Spongy Sheets on Wound Healing in Rats. Biol. Pharm. Bull. 2000, 23, 1220–1223. [Google Scholar] [CrossRef] [PubMed]
  138. Katti, D.S.; Robinson, K.W.; Ko, F.K.; Laurencin, C.T. Bioresorbable nanofiber-based systems for wound healing and drug delivery: Optimization of fabrication parameters. J. Biomed. Mater. Res. B Appl. Biomater. 2004, 70, 286–296. [Google Scholar] [CrossRef]
  139. Chen, J.-P.; Chen, S.-H.; Lai, G.-J. Preparation and characterization of biomimetic silk fibroin/chitosan composite nanofibers by electrospinning for osteoblasts culture. Nanoscale Res. Lett. 2012, 7, 170. [Google Scholar] [CrossRef]
  140. Li, W.J.; Danielson, K.G.; Alexander, P.G.; Tuan, R.S. Biological response of chondrocytes cultured in three-dimensional nanofibrous poly(epsilon-caprolactone) scaffolds. J. Biomed. Mater. Res. A 2003, 67, 1105–1114. [Google Scholar] [CrossRef]
  141. Yoshimoto, H.; Shin, Y.; Terai, H.; Vacanti, J. A biodegradable nanofiber scaffold by electrospinning and its potential for bone tissue engineering. Biomaterials 2003, 24, 2077–2082. [Google Scholar] [CrossRef] [PubMed]
  142. Dornish, M.; Kaplan, D.; Skaugrud, O. Standards and guidelines for biopolymers in tissue-engineered medical products: ASTM alginate and chitosan standard guides. American Society for Testing and Materials. Ann. N. Y. Acad. Sci. 2001, 944, 388–397. [Google Scholar] [CrossRef] [PubMed]
  143. VandeVord, P.J.; Matthew, H.W.; DeSilva, S.P.; Mayton, L.; Wu, B.; Wooley, P.H. Evaluation of the biocompatibility of a chitosan scaffold in mice. J. Biomed. Mater. Res. 2002, 59, 585–590. [Google Scholar] [CrossRef]
  144. Madihally, S.V.; Flake, A.W.; Matthew, H.W.T. Maintenance of CD34 Expression During Proliferation of CD34+ Cord Blood Cells on Glycosaminoglycan Surfaces. Stem Cells 1999, 17, 295–305. [Google Scholar] [CrossRef]
  145. Mo, X.; Xu, C.; Kotaki, M.; Ramakrishna, S. Electrospun P(LLA-CL) nanofiber: A biomimetic extracellular matrix for smooth muscle cell and endothelial cell proliferation. Biomaterials 2004, 25, 1883–1890. [Google Scholar] [CrossRef] [PubMed]
  146. Xu, C.; Inai, R.; Kotaki, M.; Ramakrishna, S. Aligned biodegradable nanofibrous structure: A potential scaffold for blood vessel engineering. Biomaterials 2003, 25, 877–886. [Google Scholar] [CrossRef]
  147. Khil, M.-S.; Cha, D.-I.; Kim, H.-Y.; Kim, I.-S.; Bhattarai, N. Electrospun nanofibrous polyurethane membrane as wound dressing. J. Biomed. Mater. Res. B Appl. Biomater. 2003, 67, 675–679. [Google Scholar] [CrossRef]
  148. Şenel, S.; McClure, S.J. Potential applications of chitosan in veterinary medicine. Adv. Drug Deliv. Rev. 2004, 56, 1467–1480. [Google Scholar] [CrossRef]
  149. Chow, K.S.; Khor, E. Novel Fabrication of Open-Pore Chitin Matrixes. Biomacromolecules 2000, 1, 61–67. [Google Scholar] [CrossRef]
  150. Shin, M.; Yoshimoto, H.; Vacanti, J.P. In Vivo Bone Tissue Engineering Using Mesenchymal Stem Cells on a Novel Electrospun Nanofibrous Scaffold. Tissue Eng. 2004, 10, 33–41. [Google Scholar] [CrossRef]
  151. Shin, M.; Ishii, O.; Sueda, T.; Vacanti, J. Contractile cardiac grafts using a novel nanofibrous mesh. Biomaterials 2004, 25, 3717–3723. [Google Scholar] [CrossRef] [PubMed]
  152. Taravel, M.; Domard, A. Collagen and its interaction with chitosan: II. Influence of the physicochemical characteristics of collagen. Biomaterials 1995, 16, 865–871. [Google Scholar] [CrossRef] [PubMed]
  153. Taravel, M.N.; Domard, A. Collagen and its interactions with chitosan, III some biological and mechanical properties. Biomaterials 1996, 17, 451–455. [Google Scholar] [CrossRef] [PubMed]
  154. Cho, Y.-N.; Chung, S.-H.; Yoo, G.; Ko, S.-W. Water-soluble chitin as a wound healing accelerator. Biomaterials 1999, 20, 2139–2145. [Google Scholar] [CrossRef] [PubMed]
  155. Ma, J.; Wang, H.; He, B.; Chen, J. A preliminary in vitro study on the fabrication and tissue engineering applications of a novel chitosan bilayer material as a scaffold of human neofetal dermal fibroblasts. Biomaterials 2000, 22, 331–336. [Google Scholar] [CrossRef]
  156. Wang, L.; Khor, E.; Wee, A.; Lim, L.Y. Chitosan-alginate PEC membrane as a wound dressing: Assessment of incisional wound healing. J. Biomed. Mater. Res. 2002, 63, 610–618. [Google Scholar] [CrossRef]
  157. Fratzl, P.; Misof, K.; Zizak, I.; Rapp, G.; Amenitsch, H.; Bernstorff, S. Fibrillar structure and mechanical properties of collagen. J. Struct. Biol. 1998, 122, 119–122. [Google Scholar] [CrossRef]
  158. Franchi, M.; Trirè, A.; Quaranta, M.; Orsini, E.; Ottani, V. Collagen Structure of Tendon Relates to Function. Sci. World J. 2007, 7, 404–420. [Google Scholar] [CrossRef]
  159. Hofmann, H.; Fietzek, P.; Kühn, K. The role of polar and hydrophobic interactions for the molecular packing of type I collagen: A three-dimensional evaluation of the amino acid sequence. J. Mol. Biol. 1978, 125, 137–165. [Google Scholar] [CrossRef]
  160. Matthews, J.A.; Wnek, G.E.; Simpson, D.G.; Bowlin, G.L. Electrospinning of Collagen Nanofibers. Biomacromolecules 2002, 3, 232–238. [Google Scholar] [CrossRef]
  161. Gelse, K.; Poschl, E.; Aigner, T. Collagens—Structure, function, and biosynthesis. Adv. Drug Deliv. Rev. 2003, 55, 1531–1546. [Google Scholar] [CrossRef] [PubMed]
  162. Abdollahiyan, P.; Oroojalian, F.; Hejazi, M.; de la Guardia, M.; Mokhtarzadeh, A. Nanotechnology, and scaffold implantation for the effective repair of injured organs: An overview on hard tissue engineering. J. Control. Release 2021, 333, 391–417. [Google Scholar] [CrossRef] [PubMed]
  163. Hulmes, D.J.S.; Miller, A. Molecular packing in collagen. Nature 1981, 293, 239–240. [Google Scholar] [CrossRef] [PubMed]
  164. Mithieux, S.M.; Weiss, A.S. Elastin. Adv. Protein Chem. 2005, 70, 437–461. [Google Scholar]
  165. Fazio, M.J.; Mattei, M.G.; Passage, E.; Chu, M.L.; Black, D.; Solomon, E.; Davidson, J.M.; Uitto, J. Human elastin gene: New evidence for localization to the long arm of chromosome 7. Am. J. Hum. Genet. 1991, 48, 696–703. [Google Scholar]
  166. Rosenbloom, J.; Abrams, W.R.; Mecham, R. Extracellular matrix 4: The elastic fiber. FASEB J. 1993, 7, 1208–1218. [Google Scholar] [CrossRef]
  167. Muiznieks, L.D.; Jensen, S.A.; Weiss, A.S. Structural changes and facilitated association of tropoelastin. Arch. Biochem. Biophys. 2003, 410, 317–323. [Google Scholar] [CrossRef]
  168. Sandberg, L.B.; Soskel, N.T.; Leslie, J.G. Elastin structure, biosynthesis, and relation to disease states. N. Engl. J. Med. 1981, 304, 566–579. [Google Scholar] [CrossRef]
  169. Uitto, V.-J.; Larjava, H. Extracellular Matrix Molecules and their Receptors: An Overview with Special Emphasis on Periodontal Tissues. Crit. Rev. Oral Biol. Med. 1991, 2, 323–354. [Google Scholar] [CrossRef]
  170. Rnjak, J.; Wise, S.G.; Mithieux, S.M.; Weiss, A.S. Severe Burn Injuries and the Role of Elastin in the Design of Dermal Substitutes. Tissue Eng. Part B Rev. 2011, 17, 81–91. [Google Scholar] [CrossRef]
  171. Ramirez, F. Pathophysiology of the microfibril/elastic fiber system: Introduction. Matrix Biol. 2000, 19, 455–456. [Google Scholar] [CrossRef] [PubMed]
  172. Nouri, K.; Jimenez, G.P.; Harrison-Balestra, C.; Elgart, G.W. 585-nm pulsed dye laser in the treatment of surgical scars starting on the suture removal day. Dermatol. Surg. 2003, 29, 65–73; discussion 73. [Google Scholar] [PubMed]
  173. Powell, J.T.; Vine, N.; Crossman, M. On the accumulation of d-aspartate in elastin and other proteins of the ageing aorta. Atherosclerosis 1992, 97, 201–208. [Google Scholar] [CrossRef] [PubMed]
  174. Hinek, A.; Wang, Y.; Liu, K.; Mitts, T.F.; Jimenez, F. Proteolytic digest derived from bovine Ligamentum Nuchae stimulates deposition of new elastin-enriched matrix in cultures and transplants of human dermal fibroblasts. J. Dermatol. Sci. 2005, 39, 155–166. [Google Scholar] [CrossRef] [PubMed]
  175. Carlisle, E.M. Silicon as a trace nutrient. Sci. Total Environ. 1988, 73, 95–106. [Google Scholar] [CrossRef]
  176. Wei, S.; Ma, J.-X.; Xu, L.; Gu, X.-S.; Ma, X.-L. Biodegradable materials for bone defect repair. Mil. Med. Res. 2020, 7, 1–25. [Google Scholar] [CrossRef]
  177. Sarker, B.; Lyer, S.; Arkudas, A.; Boccaccini, A.R. Collagen/silica nanocomposites and hybrids for bone tissue engineering. Nanotechnol. Rev. 2013, 2, 427–447. [Google Scholar] [CrossRef]
  178. Dong, C.; Lv, Y. Application of Collagen Scaffold in Tissue Engineering: Recent Advances and New Perspectives. Polymers 2016, 8, 42. [Google Scholar] [CrossRef]
  179. Khanna, A.; Zamani, M.; Huang, N.F. Extracellular Matrix-Based Biomaterials for Cardiovascular Tissue Engineering. J. Cardiovasc. Dev. Dis. 2021, 8, 137. [Google Scholar] [CrossRef]
  180. Szurkowska, K.; Kolmas, J. Hydroxyapatites enriched in silicon—Bioceramic materials for biomedical and pharmaceutical applications. Prog. Nat. Sci. Mater. 2017, 27, 401–409. [Google Scholar] [CrossRef]
  181. Al-Harbi, N.; Mohammed, H.; Al-Hadeethi, Y.; Bakry, A.S.; Umar, A.; Hussein, M.A.; Abbassy, M.A.; Vaidya, K.G.; Al Berakdar, G.; Mkawi, E.M.; et al. Silica-Based Bioactive Glasses and Their Applications in Hard Tissue Regeneration: A Review. Pharmaceuticals 2021, 14, 75. [Google Scholar] [CrossRef] [PubMed]
  182. Hao, S.; Wang, M.; Yin, Z.; Jing, Y.; Bai, L.; Su, J. Microenvironment-targeted strategy steers advanced bone regeneration. Mater. Today Bio 2023, 22, 100741. [Google Scholar] [CrossRef]
  183. Gao, C.; Peng, S.; Feng, P.; Shuai, C. Bone biomaterials and interactions with stem cells. Bone Res. 2017, 5, 17059. [Google Scholar] [CrossRef] [PubMed]
  184. Tian, B.; Liu, Y. Antibacterial applications and safety issues of silica-based materials: A review. Int. J. Appl. Ceram. Technol. 2020, 18, 289–301. [Google Scholar] [CrossRef]
  185. Song, R.; Murphy, M.; Li, C.; Ting, K.; Soo, C.; Zheng, Z. Current development of biodegradable polymeric materials for biomedical applications. Drug Des. Dev. Ther. 2018, 12, 3117–3145. [Google Scholar] [CrossRef]
  186. Labet, M.; Thielemans, W. Synthesis of polycaprolactone: A review. Chem. Soc. Rev. 2009, 38, 3484–3504. [Google Scholar] [CrossRef]
  187. Kweon, H.; Yoo, M.K.; Park, I.K.; Kim, T.H.; Lee, H.C.; Lee, H.S.; Oh, J.S.; Akaike, T.; Cho, C.S. A novel degradable polycaprolactone networks for tissue engineering. Biomaterials 2003, 24, 801–808. [Google Scholar] [CrossRef]
  188. Woodward, S.C.; Brewer, P.S.; Moatamed, F.; Schindler, A.; Pitt, C.G. The intracellular degradation of poly(epsilon-caprolactone). J. Biomed. Mater. Res. 1985, 19, 437–444. [Google Scholar] [CrossRef]
  189. Pitt, G.G.; Gratzl, M.M.; Kimmel, G.L.; Surles, J.; Sohindler, A. Aliphatic polyesters II. The degradation of poly (DL-lactide), poly (ε-caprolactone), and their copolymers in vivo. Biomaterials 1981, 2, 215–220. [Google Scholar] [CrossRef]
  190. Sinha, V.R.; Bansal, K.; Kaushik, R.; Kumria, R.; Trehan, A. Poly-epsilon-caprolactone microspheres and nanospheres: An overview. Int. J. Pharm. 2004, 278, 1–23. [Google Scholar] [CrossRef]
  191. Coombes, A.; Rizzi, S.; Williamson, M.; Barralet, J.; Downes, S.; Wallace, W. Precipitation casting of polycaprolactone for applications in tissue engineering and drug delivery. Biomaterials 2003, 25, 315–325. [Google Scholar] [CrossRef] [PubMed]
  192. Uhrich, K.E.; Cannizzaro, S.M.; Langer, R.S.; Shakesheff, K.M. Polymeric Systems for Controlled Drug Release. Chem. Rev. 1999, 99, 3181–3198. [Google Scholar] [CrossRef] [PubMed]
  193. Wu, X.S. Synthesis, Characterization, Biodegradation, and Drug Delivery Application of Biodegradable Lactic/Glycolic Acid Polymers: Part III. Drug Delivery Application. Artif. Cells Blood Substit. Biotechnol. 2004, 32, 575–591. [Google Scholar] [CrossRef] [PubMed]
  194. Makadia, H.K.; Siegel, S.J. Poly lactic-co-glycolic acid (PLGA) As biodegradable controlled drug delivery carrier. Polymers 2011, 3, 1377–1397. [Google Scholar] [CrossRef]
  195. Shirazi, R.N.; Aldabbagh, F.; Erxleben, A.; Rochev, Y.; McHugh, P. Nanomechanical properties of poly(lactic-co-glycolic) acid film during degradation. Acta Biomater. 2014, 10, 4695–4703. [Google Scholar] [CrossRef]
  196. Dzobo, K.; Thomford, N.E.; Senthebane, D.A.; Shipanga, H.; Rowe, A.; Dandara, C.; Pillay, M.; Motaung, K.S.C.M. Advances in regenerative medicine and tissue engineering: Innovation and transformation of medicine. Stem Cells Int. 2018, 2018, 2495848. [Google Scholar] [CrossRef]
  197. Han, F.; Wang, J.; Ding, L.; Hu, Y.; Li, W.; Yuan, Z.; Guo, Q.; Zhu, C.; Yu, L.; Wang, H.; et al. Tissue Engineering and Regenerative Medicine: Achievements, Future, and Sustainability in Asia. Front. Bioeng. Biotechnol. 2020, 8, 83. [Google Scholar] [CrossRef]
  198. Witkowska, D.; Rowińska-Żyrek, M. Biophysical approaches for the study of metal-protein interactions. J. Inorg. Biochem. 2019, 199, 110783. [Google Scholar] [CrossRef]
  199. Churchfield, L.A.; Tezcan, F.A. Design and Construction of Functional Supramolecular Metalloprotein Assemblies. Acc. Chem. Res. 2019, 52, 345–355. [Google Scholar] [CrossRef]
  200. Nastri, F.; D’alonzo, D.; Leone, L.; Zambrano, G.; Pavone, V.; Lombardi, A. Engineering Metalloprotein Functions in Designed and Native Scaffolds. Trends Biochem. Sci. 2019, 44, 1022–1040. [Google Scholar] [CrossRef]
  201. Lin, P.-H.; Sermersheim, M.; Li, H.; Lee, P.H.U.; Steinberg, S.M.; Ma, J. Zinc in Wound Healing Modulation. Nutrients 2017, 10, 16. [Google Scholar] [CrossRef] [PubMed]
  202. Tenaud, I.; Sainte-Marie, I.; Jumbou, O.; Litoux, P.; Dreno, B. In vitro modulation of keratinocyte wound healing integrins by zinc, copper and manganese. Brit. J. Dermatol. 1999, 140, 26–34. [Google Scholar] [CrossRef] [PubMed]
  203. Huang, J.; Chen, L.; Yuan, Q.; Gu, Z.; Wu, J. Tofu-Based Hybrid Hydrogels with Antioxidant and Low Immunogenicity Activity for Enhanced Wound Healing. J. Biomed. Nanotechnol. 2019, 15, 1371–1383. [Google Scholar] [CrossRef] [PubMed]
  204. Coelho, T.S.; Halicki, P.C.B.; Silva, L.; Vicenti, J.R.d.M.; Gonçalves, B.L.; da Silva, P.E.A.; Ramos, D.F. Metal-based antimicrobial strategies against intramacrophage Mycobacterium tuberculosis. Lett. Appl. Microbiol. 2020, 71, 146–153. [Google Scholar] [CrossRef] [PubMed]
  205. Zhang, E.; Zhao, X.; Hu, J.; Wang, R.; Fu, S.; Qin, G. Antibacterial metals and alloys for potential biomedical implants. Bioact. Mater. 2021, 6, 2569–2612. [Google Scholar] [CrossRef]
  206. Eming, S.A.; Martin, P.; Tomic-Canic, M. Wound repair and regeneration: Mechanisms, signaling, and translation. Sci. Transl. Med. 2014, 6, 265sr6. [Google Scholar] [CrossRef]
  207. Ashtiani, R.E.; Alam, M.; Tavakolizadeh, S.; Abbasi, K. The Role of Biomaterials and Biocompatible Materials in Implant-Supported Dental Prosthesis. Evid. Based Complement. Altern. Med. 2021, 2021, 3349433. [Google Scholar] [CrossRef]
  208. Long, M.; Rack, H.J. Titanium alloys in total joint replacement—A materials science perspective. Biomaterials 1998, 19, 1621–1639. [Google Scholar] [CrossRef]
  209. Li, Y.; Jahr, H.; Lietaert, K.; Pavanram, P.; Yilmaz, A.; Fockaert, L.I.; Leeflang, M.A.; Pouran, B.; Gonzalez-Garcia, Y.; Weinans, H.; et al. Additively manufactured biodegradable porous iron. Acta Biomater. 2018, 77, 380–393. [Google Scholar] [CrossRef]
  210. Wang, X.; Xu, S.; Zhou, S.; Xu, W.; Leary, M.; Choong, P.; Qian, M.; Brandt, M.; Xie, Y.M. Topological design and additive manufacturing of porous metals for bone scaffolds and orthopaedic implants: A review. Biomaterials 2016, 83, 127–141. [Google Scholar] [CrossRef]
  211. You, J.; Zhang, Y.; Zhou, Y. Strontium Functionalized in Biomaterials for Bone Tissue Engineering: A Prominent Role in Osteoimmunomodulation. Front. Bioeng. Biotechnol. 2022, 10, 928799. [Google Scholar] [CrossRef]
  212. Bulina, N.V.; Vinokurova, O.B.; Prosanov, I.Y.; Vorobyev, A.M.; Gerasimov, K.B.; Borodulina, I.A.; Pryadko, A.; Botvin, V.V.; Surmeneva, M.A.; Surmenev, R.A. Mechanochemical synthesis of strontium- and magnesium-substituted and cosubstituted hydroxyapatite powders for a variety of biomedical applications. Ceram. Int. 2022, 48, 35217–35226. [Google Scholar] [CrossRef]
  213. Iseri, L.T.; French, J.H. Magnesium: Nature’s physiologic calcium blocker. Am. Heart J. 1984, 108, 188–193. [Google Scholar] [CrossRef] [PubMed]
  214. Yang, L.; Arora, K.; Beard, W.A.; Wilson, S.H.; Schlick, T. Critical role of magnesium ions in DNA polymerase beta’s closing and active site assembly. J. Am. Chem. Soc. 2004, 126, 8441–8453. [Google Scholar] [CrossRef] [PubMed]
  215. Panaghie, C.; Zegan, G.; Sodor, A.; Cimpoeșu, N.; Lohan, N.-M.; Istrate, B.; Roman, A.-M.; Ioanid, N. Analysis of Degradation Products of Biodegradable ZnMgY Alloy. Materials 2023, 16, 3092. [Google Scholar] [CrossRef] [PubMed]
  216. Wallach, S. Effects of magnesium on skeletal metabolism. Magnes. Trace Elem. 1989, 9, 1–14. [Google Scholar]
  217. Cowan, J. Structural and catalytic chemistry of magnesium-dependent enzymes. BioMetals 2002, 15, 225–235. [Google Scholar] [CrossRef]
  218. Sissi, C.; Palumbo, M. Effects of magnesium and related divalent metal ions in topoisomerase structure and function. Nucleic Acids Res. 2009, 37, 702–711. [Google Scholar] [CrossRef]
  219. Haftek, M.; Abdayem, R.; Guyonnet-Debersac, P. Skin Minerals: Key Roles of Inorganic Elements in Skin Physiological Functions. Int. J. Mol. Sci. 2022, 23, 6267. [Google Scholar] [CrossRef]
  220. Chandrasekaran, N.C.; Weir, C.; Alfraji, S.; Grice, J.; Roberts, M.S.; Barnard, R.T. Effects of magnesium deficiency—More than skin deep. Exp. Biol. Med. 2014, 239, 1280–1291. [Google Scholar] [CrossRef]
  221. Del Rosso, J.; Zeichner, J.; Alexis, A.; Cohen, D.; Berson, D. Understanding the Epidermal Barrier in Healthy and Compromised Skin: Clinically Relevant Information for the Dermatology Practitioner: Proceedings of an Expert Panel Roundtable Meeting. J. Clin. Aesthetic Dermatol. 2016, 9 (Suppl. S1), S2–S8. [Google Scholar]
  222. Denda, M.; Katagiri, C.; Hirao, T.; Maruyama, N.; Takahashi, M. Some magnesium salts and a mixture of magnesium and calcium salts accelerate skin barrier recovery. Arch. Dermatol. Res. 1999, 291, 560–563. [Google Scholar] [CrossRef] [PubMed]
  223. Denda, M. New strategies to improve skin barrier homeostasis. Adv. Drug Deliv. Rev. 2002, 54 (Suppl. S1), S123–S130. [Google Scholar] [CrossRef]
  224. Denda, M. Methodology to improve epidermal barrier homeostasis: How to accelerate the barrier recovery? Int. J. Cosmet. Sci. 2009, 31, 79–86. [Google Scholar] [CrossRef]
  225. Jahnen-Dechent, W.; Ketteler, M. Magnesium basics. Clin. Kidney J. 2012, 5 (Suppl. S1), i3–i14. [Google Scholar] [CrossRef]
  226. Njau, S.; Epivatianos, P.; Tsoukali-Papadopoulou, H.; Psaroulis, D.; Stratis, J. Magnesium, calcium and zinc fluctuations on skin induced injuries in correlation with time of induction. Forensic Sci. Int. 1991, 50, 67–73. [Google Scholar] [CrossRef] [PubMed]
  227. Proksch, E.; Nissen, H.-P.; Bremgartner, M.; Urquhart, C. Bathing in a magnesium-rich Dead Sea salt solution improves skin barrier function, enhances skin hydration, and reduces inflammation in atopic dry skin. Int. J. Dermatol. 2005, 44, 151–157. [Google Scholar] [CrossRef] [PubMed]
  228. Rude, R.K.; Adams, J.S.; Ryzen, E.; Endres, D.B.; Niimi, H.; Horst, R.L.; Haddad, J.G.; Singer, F.R. Low Serum Concentrations of 1,25-Dihydroxyvitamin D in Human Magnesium Deficiency. J. Clin. Endocrinol. Metab. 1985, 61, 933–940. [Google Scholar] [CrossRef] [PubMed]
  229. Sahota, O.; Mundey, M.K.; San, P.; Godber, I.M.; Hosking, D.J. Vitamin D insufficiency and the blunted PTH response in established osteoporosis: The role of magnesium deficiency. Osteoporos. Int. 2006, 17, 1013–1021. [Google Scholar] [CrossRef]
  230. Nie, X.; Sun, X.; Wang, C.; Yang, J. Effect of magnesium ions/Type I collagen promote the biological behavior of osteoblasts and its mechanism. Regen. Biomater. 2019, 7, 53–61. [Google Scholar] [CrossRef]
  231. Shigematsu, T.; Tajima, S.; Nishikawa, T.; Murad, S.; Pinnell, S.R.; Nishioka, I. Inhibition of collagen hydroxylation by lithospermic acid magnesium salt, a novel compound isolated from Salviae miltiorrhizae Radix. Biochim. Biophys. Acta (BBA) Gen. Subj. 1994, 1200, 79–83. [Google Scholar] [CrossRef]
  232. Müller, W.; Iffland, R.; Firsching, R. Relationship between magnesium and elastic fibres. Magnes. Res. 1993, 6, 215–222. [Google Scholar] [PubMed]
  233. Muller, W.; Firsching, R. Differentiation of Oxytalan Fibers from Elastic Fibers with Reagents for Detection of Magnesium. Ann. Anat. 1992, 174, 357–359. [Google Scholar] [CrossRef] [PubMed]
  234. Heinz, A. Elastases and elastokines: Elastin degradation and its significance in health and disease. Crit. Rev. Biochem. Mol. Biol. 2020, 55, 252–273. [Google Scholar] [CrossRef] [PubMed]
  235. Senni, K.; Foucault-Bertaud, A.; Godeau, G. Magnesium and connective tissue. Magnes. Res. 2003, 16, 70–74. [Google Scholar]
  236. Heinegård, D.; Wieslander, J.; Sheehan, J.; Paulsson, M.; Sommarin, Y. Separation and characterization of two populations of aggregating proteoglycans from cartilage. Biochem. J. 1985, 225, 95–106. [Google Scholar] [CrossRef]
  237. Yamaguchi, Y.; Mann, D.M.; Ruoslahti, E. Negative regulation of transforming growth factor-beta by the proteoglycan decorin. Nature 1990, 346, 281–284. [Google Scholar] [CrossRef]
  238. Nunes, A.M.; Minetti, C.A.; Remeta, D.P.; Baum, J. Magnesium Activates Microsecond Dynamics to Regulate Integrin-Collagen Recognition. Structure 2018, 26, 1080–1090.e5. [Google Scholar] [CrossRef]
  239. Harjunpää, H.; Asens, M.L.; Guenther, C.; Fagerholm, S.C. Cell Adhesion Molecules and Their Roles and Regulation in the Immune and Tumor Microenvironment. Front. Immunol. 2019, 10, 1078. [Google Scholar] [CrossRef]
  240. Danen, E.H.; Sonnenberg, A. Integrins in regulation of tissue development and function. J. Pathol. 2003, 200, 471–480. [Google Scholar] [CrossRef]
  241. Lange, T.S.; Bielinsky, A.K.; Kirchberg, K.; Bank, I.; Herrmann, K.; Krieg, T.; Scharffetter-Kochanek, K. Mg2+ and Ca2+ differentially regulate beta 1 integrin-mediated adhesion of dermal fibroblasts and keratinocytes to various extracellular matrix proteins. Exp. Cell Res. 1994, 214, 381–388. [Google Scholar] [CrossRef] [PubMed]
  242. Lange, T.S.; Kirchberg, K.; Bielinsky, A.K.; Leuker, A.; Bank, I.; Ruzicka, T.; Scharffetter-Kochanek, K. Divalent cations (Mg2+, Ca2+) differentially influence the beta1 integrin-mediated migration of human fibroblasts and keratinocytes to different extracellular matrix proteins. Exp. Dermatol. 1995, 4, 130–137. [Google Scholar] [CrossRef]
  243. Stelling, M.P.; Motta, J.M.; Mashid, M.; Johnson, W.E.; Pavão, M.S.; Farrell, N.P. Metal ions and the extracellular matrix in tumor migration. FEBS J. 2019, 286, 2950–2964. [Google Scholar] [CrossRef]
  244. Antoniac, I.; Miculescu, M.; Păltânea, V.M.; Stere, A.; Quan, P.H.; Păltânea, G.; Robu, A.; Earar, K. Magnesium-Based Alloys Used in Orthopedic Surgery. Materials 2022, 15, 1148. [Google Scholar] [CrossRef]
  245. Brandt-Wunderlich, C.; Ruppelt, P.; Zumstein, P.; Schmidt, W.; Arbeiter, D.; Schmitz, K.-P.; Grabow, N. Mechanical behavior of in vivo degraded second generation resorbable magnesium scaffolds (RMS). J. Mech. Behav. Biomed. Mater. 2018, 91, 174–181. [Google Scholar] [CrossRef]
  246. Qin, Y.; Wen, P.; Guo, H.; Xia, D.; Zheng, Y.; Jauer, L.; Poprawe, R.; Voshage, M.; Schleifenbaum, J.H. Additive manufacturing of biodegradable metals: Current research status and future perspectives. Acta Biomater. 2019, 98, 3–22. [Google Scholar] [CrossRef] [PubMed]
  247. Amukarimi, S.; Mozafari, M. Biodegradable magnesium-based biomaterials: An overview of challenges and opportunities. Medcomm 2021, 2, 123–144. [Google Scholar] [CrossRef] [PubMed]
  248. Seetharaman, S.; Sankaranarayanan, D.; Gupta, M. Magnesium-Based Temporary Implants: Potential, Current Status, Applications, and Challenges. J. Funct. Biomater. 2023, 14, 324. [Google Scholar] [CrossRef]
  249. Kim, Y.-K.; Lee, K.-B.; Kim, S.-Y.; Bode, K.; Jang, Y.-S.; Kwon, T.-Y.; Jeon, M.H.; Lee, M.-H. Gas formation and biological effects of biodegradable magnesium in a preclinical and clinical observation. Sci. Technol. Adv. Mater. 2018, 19, 324–335. [Google Scholar] [CrossRef]
  250. Fiorentini, D.; Cappadone, C.; Farruggia, G.; Prata, C. Magnesium: Biochemistry, Nutrition, Detection, and Social Impact of Diseases Linked to Its Deficiency. Nutrients 2021, 13, 1136. [Google Scholar] [CrossRef]
  251. Swaminathan, R. Magnesium metabolism and its disorders. Clin. Biochem. Rev. 2003, 24, 47–66. [Google Scholar] [PubMed]
  252. Gonzalez, J.; Hou, R.Q.; Nidadavolu, E.P.; Willumeit-Römer, R.; Feyerabend, F. Magnesium degradation under physiological conditions—Best practice. Bioact. Mater. 2018, 3, 174–185. [Google Scholar] [CrossRef] [PubMed]
  253. Xin, Y.; Huo, K.; Tao, H.; Tang, G.; Chu, P.K. Influence of aggressive ions on the degradation behavior of biomedical magnesium alloy in physiological environment. Acta Biomater. 2008, 4, 2008–2015. [Google Scholar] [CrossRef] [PubMed]
  254. Törne, K.; Örnberg, A.; Weissenrieder, J. The influence of buffer system and biological fluids on the degradation of magnesium. J. Biomed. Mater. Res. Part B Appl. Biomater. 2016, 105, 1490–1502. [Google Scholar] [CrossRef]
  255. Kokubo, T.; Takadama, H. How useful is SBF in predicting in vivo bone bioactivity? Biomaterials 2006, 27, 2907–2915. [Google Scholar] [CrossRef]
  256. Oyane, A.; Kim, H.-M.; Furuya, T.; Kokubo, T.; Miyazaki, T.; Nakamura, T. Preparation and assessment of revised simulated body fluids. J. Biomed. Mater. Res. A 2003, 65, 188–195. [Google Scholar] [CrossRef]
  257. Gu, X.N.; Zheng, Y.F.; Chen, L.J. Influence of artificial biological fluid composition on the biocorrosion of potential orthopedic Mg-Ca, AZ31, AZ91 alloys. Biomed. Mater. 2009, 4, 065011. [Google Scholar] [CrossRef]
  258. Zhang, J.; Kong, N.; Shi, Y.; Niu, J.; Mao, L.; Li, H.; Xiong, M.; Yuan, G. Influence of proteins and cells on in vitro corrosion of Mg-Nd-Zn-Zr alloy. Corros. Sci. 2014, 85, 477–481. [Google Scholar] [CrossRef]
  259. Johnson, I.; Jiang, W.; Liu, H. The Effects of Serum Proteins on Magnesium Alloy Degradation in Vitro. Sci. Rep. 2017, 7, 14335. [Google Scholar] [CrossRef]
  260. Gao, Y.; Wang, L.; Li, L.; Gu, X.; Zhang, K.; Xia, J.; Fan, Y. Effect of stress on corrosion of high-purity magnesium in vitro and in vivo. Acta Biomater. 2018, 83, 477–486. [Google Scholar] [CrossRef]
  261. Liu, M.; Wang, J.; Zhu, S.; Zhang, Y.; Sun, Y.; Wang, L.; Guan, S. Corrosion fatigue of the extruded Mg-Zn-Y-Nd alloy in simulated body fluid. J. Magnes. Alloy 2020, 8, 231–240. [Google Scholar] [CrossRef]
  262. Bian, D.; Zhou, W.; Liu, Y.; Li, N.; Zheng, Y.; Sun, Z. Fatigue behaviors of HP-Mg, Mg-Ca and Mg-Zn-Ca biodegradable metals in air and simulated body fluid. Acta Biomater. 2016, 41, 351–360. [Google Scholar] [CrossRef]
  263. Zhen, Z.; Xi, T.-F.; Zheng, Y.-F. A review on in vitro corrosion performance test of biodegradable metallic materials. Trans. Nonferrous Met. Soc. China 2013, 23, 2283–2293. [Google Scholar] [CrossRef]
  264. Mei, D.; Lamaka, S.V.; Lu, X.; Zheludkevich, M.L. Selecting medium for corrosion testing of bioabsorbable magnesium and other metals—A critical review. Corros. Sci. 2020, 171, 108722. [Google Scholar] [CrossRef]
  265. Claes, L. Mechanical characterization of biodegradable implants. Clin. Mater. 1992, 10, 41–46. [Google Scholar] [CrossRef]
  266. Fard, M.G.; Sharifianjazi, F.; Kazemi, S.S.; Rostamani, H.; Bathaei, M.S. Laser-Based Additive Manufacturing of Magnesium Alloys for Bone Tissue Engineering Applications: From Chemistry to Clinic. J. Manuf. Mater. Process. 2022, 6, 158. [Google Scholar] [CrossRef]
  267. Sotomi, Y.; Onuma, Y.; Collet, C.; Tenekecioglu, E.; Virmani, R.; Kleiman, N.S.; Serruys, P.W. Bioresorbable Scaffold the Emerging Reality and Future Directions. Circ. Res. 2017, 120, 1341–1352. [Google Scholar] [CrossRef]
  268. Wang, J.L.; Xu, J.K.; Hopkins, C.; Chow, D.H.K.; Qin, L. Biodegradable Magnesium-Based Implants in Orthopedics—A General Review and Perspectives. Adv. Sci. 2020, 7, 1902443. [Google Scholar] [CrossRef]
  269. Witte, F. The history of biodegradable magnesium implants: A review. Acta Biomater. 2010, 6, 1680–1692. [Google Scholar] [CrossRef]
  270. Willumeit, R.; Fischer, J.; Feyerabend, F.; Hort, N.; Bismayer, U.; Heidrich, S.; Mihailova, B. Chemical surface alteration of biodegradable magnesium exposed to corrosion media. Acta Biomater. 2011, 7, 2704–2715. [Google Scholar] [CrossRef]
  271. Banerjee, P.C.; Al-Saadi, S.; Choudhary, L.; Harandi, S.E.; Singh, R. Magnesium Implants: Prospects and Challenges. Materials 2019, 12, 136. [Google Scholar] [CrossRef]
  272. Yin, M.; Xu, F.; Ding, H.; Tan, F.; Song, F.; Wang, J. Incorporation of magnesium ions into photo-crosslinked alginate hydrogel enhanced cell adhesion ability. J. Tissue Eng. Regen. Med. 2015, 9, 1088–1092. [Google Scholar] [CrossRef]
  273. Roh, H.-S.; Lee, C.-M.; Hwang, Y.-H.; Kook, M.-S.; Yang, S.-W.; Lee, D.; Kim, B.-H. Addition of MgO nanoparticles and plasma surface treatment of three-dimensional printed polycaprolactone/hydroxyapatite scaffolds for improving bone regeneration. Mater. Sci. Eng. C Mater. Biol. Appl. 2017, 74, 525–535. [Google Scholar] [CrossRef]
  274. Yuan, Z.; Wei, P.; Huang, Y.; Zhang, W.; Chen, F.; Zhang, X.; Mao, J.; Chen, D.; Cai, Q.; Yang, X. Injectable PLGA microspheres with tunable magnesium ion release for promoting bone regeneration. Acta Biomater. 2018, 85, 294–309. [Google Scholar] [CrossRef]
  275. Witte, F.; Kaese, V.; Haferkamp, H.; Switzer, E.; Meyer-Lindenberg, A.; Wirth, C.J.; Windhagen, H. In vivo corrosion of four magnesium alloys and the associated bone response. Biomaterials 2005, 26, 3557–3563. [Google Scholar] [CrossRef] [PubMed]
  276. Feyerabend, F.; Fischer, J.; Holtz, J.; Witte, F.; Willumeit, R.; Drücker, H.; Vogt, C.; Hort, N. Evaluation of short-term effects of rare earth and other elements used in magnesium alloys on primary cells and cell lines. Acta Biomater. 2010, 6, 1834–1842. [Google Scholar] [CrossRef] [PubMed]
  277. Antoniac, I.; Paltanea, V.M.; Paltanea, G.; Antoniac, A.; Nemoianu, I.V.; Petrescu, M.I.; Dura, H.; Bodog, A.D. Additive Manufactured Magnesium-Based Scaffolds for Tissue Engineering. Materials 2022, 15, 8693. [Google Scholar] [CrossRef] [PubMed]
  278. Zhang, E.; Yang, L.; Xu, J.; Chen, H. Microstructure, mechanical properties and bio-corrosion properties of Mg-Si(-Ca, Zn) alloy for biomedical application. Acta Biomater. 2010, 6, 1756–1762. [Google Scholar] [CrossRef]
  279. Istrate, B.; Munteanu, C.; Antoniac, I.V.; Lupescu, S.C. Current Research Studies of Mg-Ca-Zn Biodegradable Alloys Used as Orthopedic Implants-Review. Crystals 2022, 12, 1468. [Google Scholar] [CrossRef]
  280. Ben-Hamu, G.; Eliezer, D.; Shin, K.S. The role of Si and Ca on new wrought Mg-Zn-Mn based alloy. Mat. Sci. Eng. A 2007, 447, 35–43. [Google Scholar] [CrossRef]
  281. Turan, B. A Brief Overview from the Physiological and Detrimental Roles of Zinc Homeostasis via Zinc Transporters in the Heart. Biol. Trace Elem. Res. 2018, 188, 160–176. [Google Scholar] [CrossRef] [PubMed]
  282. Jackson, M.J.; Lowe, N.M. Physiological-Role of Zinc. Food Chem. 1992, 43, 233–238. [Google Scholar] [CrossRef]
  283. Maret, W. Zinc in Cellular Regulation: The Nature and Significance of “Zinc Signals”. Int. J. Mol. Sci. 2017, 18, 2285. [Google Scholar] [CrossRef]
  284. Vallee, B.L.; Falchuk, K.H.; Hashimoto, A.; Nakagawa, M.; Tsujimura, N.; Miyazaki, S.; Kizu, K.; Goto, T.; Komatsu, Y.; Matsunaga, A.; et al. The biochemical basis of zinc physiology. Physiol. Rev. 1993, 73, 79–118. [Google Scholar] [CrossRef] [PubMed]
  285. Beyersmann, D.; Haase, H. Functions of zinc in signaling, proliferation and differentiation of mammalian cells. BioMetals 2001, 14, 331–341. [Google Scholar] [CrossRef]
  286. Haase, H.; Maret, W. Intracellular zinc fluctuations modulate protein tyrosine phosphatase activity in insulin/insulin-like growth factor-1 signaling. Exp. Cell Res. 2003, 291, 289–298. [Google Scholar] [CrossRef]
  287. Andreini, C.; Bertini, I.; Rosato, A. Metalloproteomes: A Bioinformatic Approach. Acc. Chem. Res. 2009, 42, 1471–1479. [Google Scholar] [CrossRef]
  288. Brylinski, M.; Skolnick, J. FINDSITE-metal: Integrating evolutionary information and machine learning for structure-based metal-binding site prediction at the proteome level. Proteins: Struct. Funct. Bioinform. 2010, 79, 735–751. [Google Scholar] [CrossRef]
  289. Maret, W. Zinc Coordination Environments in Proteins as Redox Sensors and Signal Transducers. Antioxid. Redox Signal. 2006, 8, 1419–1441. [Google Scholar] [CrossRef]
  290. Cousins, R.J.; Liuzzi, J.P.; Lichten, L.A. Mammalian zinc transport, trafficking, and signals. J. Biol. Chem. 2006, 281, 24085–24089. [Google Scholar] [CrossRef]
  291. Eide, D.J. Zinc transporters and the cellular trafficking of zinc. Biochim. Biophys. Acta (BBA) Mol. Cell Res. 2006, 1763, 711–722. [Google Scholar] [CrossRef]
  292. Bentley, P.J. Influx of Zinc by Channel Catfish (Ictalurus-Punctatus)—Uptake from External Environmental Solutions. Comp. Biochem. Physiol. C-Pharmacol. Toxicol. Endocrinol. 1992, 101, 215–217. [Google Scholar] [CrossRef]
  293. Hogstrand, C.; Verbost, P.M.; Bonga, S.E.; Wood, C.M. Mechanisms of zinc uptake in gills of freshwater rainbow trout: Interplay with calcium transport. Am. J. Physiol. Integr. Comp. Physiol. 1996, 270, R1141–R1147. [Google Scholar] [CrossRef]
  294. Haase, H.; Ober-Blöbaum, J.L.; Engelhardt, G.; Hebel, S.; Heit, A.; Heine, H.; Rink, L. Zinc Signals Are Essential for Lipopolysaccharide-Induced Signal Transduction in Monocytes. Perspect. Surg. 2008, 181, 6491–6502. [Google Scholar] [CrossRef]
  295. Haase, H.; Rink, L. Functional Significance of Zinc-Related Signaling Pathways in Immune Cells. Annu. Rev. Nutr. 2009, 29, 133–152. [Google Scholar] [CrossRef]
  296. Haase, H.; Hebel, S.; Engelhardt, G.; Rink, L. Flow cytometric measurement of labile zinc in peripheral blood mononuclear cells. Anal. Biochem. 2006, 352, 222–230. [Google Scholar] [CrossRef]
  297. Frassinetti, S.; Bronzetti, G.; Caltavuturo, L.; Cini, M.; Della Croce, C. The role of zinc in life: A review. J. Environ. Pathol. Tox. 2006, 25, 597–610. [Google Scholar] [CrossRef] [PubMed]
  298. Lansdown, A.B.G.; Mirastschijski, U.; Stubbs, N.; Scanlon, E.; Agren, M.S. Zinc in wound healing: Theoretical, experimental, and clinical aspects. Wound Repair Regener. 2007, 15, 2–16. [Google Scholar] [CrossRef] [PubMed]
  299. Ågren, M.S. Percutaneous Absorption of Zinc from Zinc Oxide Applied Topically to Intact Skin in Man. Dermatology 1990, 180, 36–39. [Google Scholar] [CrossRef]
  300. Henzel, J.H.; DeWeese, M.S.; Lichti, E.L. Zinc concentrations within healing wounds. Significance of postoperative zincuria on availability and requirements during tissue repair. Arch. Surg. 1970, 100, 349–357. [Google Scholar] [CrossRef] [PubMed]
  301. Michaelsson, G.; Ljunghall, K.; Danielson, B.G. Zinc in epidermis and dermis in healthy subjects. Acta Derm. Venereol. 1980, 60, 295–299. [Google Scholar] [CrossRef] [PubMed]
  302. Portnoy, B.; Dyer, A.; Molokhia, A. Neutron activation analysis of trace elements in skin. IX. Rubidium in normal skin. Br. J. Dermatol. 1981, 105, 445–450. [Google Scholar] [CrossRef] [PubMed]
  303. Hanada, K.; Sawamura, D.; Hashimoto, I.; Kida, K.; Naganuma, A. Epidermal Proliferation of the Skin in Metallothionein-Null Mice. J. Investig. Dermatol. 1998, 110, 259–262. [Google Scholar] [CrossRef] [PubMed]
  304. Iwata, M.; Takebayashi, T.; Ohta, H.; Alcalde, R.E.; Itano, Y.; Matsumura, T. Zinc accumulation and metallothionein gene expression in the proliferating epidermis during wound healing in mouse skin. Histochem. Cell Biol. 1999, 112, 283–290. [Google Scholar] [CrossRef]
  305. Lansdown, A.B.G. Metallothioneins: Potential therapeutic aids for wound healing in the skin. Wound Repair Regen. 2002, 10, 130–132. [Google Scholar] [CrossRef]
  306. Lansdown, A.B.G. Calcium: A potential central regulator in wound healing in the skin. Wound Repair Regen. 2002, 10, 271–285. [Google Scholar] [CrossRef]
  307. Sakamoto, M.; Tzeng, S.; Fukuyama, K.; Epstein, W.L. Light-scattering studies of cation-stimulated filament assembly of newborn rat epidermal keratin. Biochim. Et Biophys. Acta (BBA) Protein Struct. 1980, 624, 205–210. [Google Scholar] [CrossRef]
  308. Lansdown, A. Morphological variations in keratinising epithelia in the beagle. Veter. Rec. 1985, 116, 127–130. [Google Scholar] [CrossRef]
  309. Henkin, R.I.; Schecter, P.J.; Friedewald, W.T.; Demets, D.L.; Raff, M. A double blind study of the effects of zinc sulfate on taste and smell dysfunction. Am. J. Med. Sci. 1976, 272, 285–299. [Google Scholar] [CrossRef]
  310. Ravanti, L.; Kähäri, V.M. Matrix metalloproteinases in wound repair (review). Int. J. Mol. Med. 2000, 6, 391–798. [Google Scholar] [CrossRef]
  311. Plum, L.M.; Rink, L.; Haase, H. The Essential Toxin: Impact of Zinc on Human Health. Int. J. Environ. Res. Public Health 2010, 7, 1342–1365. [Google Scholar] [CrossRef] [PubMed]
  312. Wessels, I.; Maywald, M.; Rink, L. Zinc as a Gatekeeper of Immune Function. Nutrients 2017, 9, 1286. [Google Scholar] [CrossRef]
  313. Ranasinghe, P.; Wathurapatha, W.; Ishara, M.; Jayawardana, R.; Galappatthy, P.; Katulanda, P.; Constantine, G. Effects of Zinc supplementation on serum lipids: A systematic review and meta-analysis. Nutr. Metab. 2015, 12, 1–16. [Google Scholar] [CrossRef]
  314. Sum, E.Y.; O’Reilly, L.A.; Jonas, N.; Lindeman, G.J.; Visvader, J.E. The LIM Domain Protein Lmo4 Is Highly Expressed in Proliferating Mouse Epithelial Tissues. J. Histochem. Cytochem. 2005, 53, 475–486. [Google Scholar] [CrossRef] [PubMed]
  315. Zhu, C.-H.; Ying, D.-J.; Mi, J.-H.; Zhang, W.; Dong, S.-W.; Sun, J.-S.; Zhang, J.-P. The zinc finger protein A20 protects endothelial cells from burns serum injury. Burns 2004, 30, 127–133. [Google Scholar] [CrossRef] [PubMed]
  316. Cousins, R.J.; Leinart, A.S. Tissue-specific regulation of zinc metabolism and metallothionein genes by interleukin 1. FASEB J. 1988, 2, 2884–2890. [Google Scholar] [CrossRef]
  317. Lansdown, A.B.G.; Sampson, B.; Rowe, A. Sequential changes in trace metal, metallothionein and calmodulin concentrations in healing skin wounds. J. Anat. 1999, 195 Pt 3, 375–386. [Google Scholar] [CrossRef]
  318. Savlov, E.D.; Strain, W.H.; Huegin, F. Radiozinc studies in experimental wound healing. J. Surg. Res. 1962, 2, 209–212. [Google Scholar] [CrossRef]
  319. Bernstein, I.; Chakrabarti, S.; Kumaroo, K.; Sibrack, L. Synthesis of protein in the mammalian epidermis. J. Investig. Dermatol. 1970, 55, 291–302. [Google Scholar] [CrossRef]
  320. Tenaud, I.; Leroy, S.; Chebassier, N.; Dreno, B. Zinc, copper and manganese enhanced keratinocyte migration through a functional modulation of keratinocyte integrins. Exp. Dermatol. 2000, 9, 407–416. [Google Scholar] [CrossRef]
  321. Zhu, D.; Cockerill, I.; Su, Y.; Zhang, Z.; Fu, J.; Lee, K.-W.; Ma, J.; Okpokwasili, C.; Tang, L.; Zheng, Y.; et al. Mechanical Strength, Biodegradation, and in Vitro and in Vivo Biocompatibility of Zn Biomaterials. ACS Appl. Mater. Interfaces 2019, 11, 6809–6819. [Google Scholar] [CrossRef] [PubMed]
  322. Mostaed, E.; Sikora-Jasinska, M.; Drelich, J.W.; Vedani, M. Zinc-based alloys for degradable vascular stent applications. Acta Biomater. 2018, 71, 1–23. [Google Scholar] [CrossRef] [PubMed]
  323. Vojtěch, D.; Kubásek, J.; Šerák, J.; Novák, P. Mechanical and corrosion properties of newly developed biodegradable Zn-based alloys for bone fixation. Acta Biomater. 2011, 7, 3515–3522. [Google Scholar] [CrossRef]
  324. Venezuela, J.J.D.; Johnston, S.; Dargusch, M.S. The Prospects for Biodegradable Zinc in Wound Closure Applications. Adv. Healthc. Mater. 2019, 8, e1900408. [Google Scholar] [CrossRef] [PubMed]
  325. Bowen, P.K.; Drelich, J.; Goldman, J. Zinc Exhibits Ideal Physiological Corrosion Behavior for Bioabsorbable Stents. Adv. Mater. 2013, 25, 2577–2582. [Google Scholar] [CrossRef]
  326. Wu, J.; Wang, L.; He, J.; Zhu, C. In vitro cytotoxicity of Cu(2)(+), Zn(2)(+), Ag(+) and their mixtures on primary human endometrial epithelial cells. Contraception 2012, 85, 509–518. [Google Scholar] [CrossRef]
  327. Yoo, M.H.; Lee, J.Y.; Lee, S.E.; Koh, J.Y.; Yoon, Y.H. Protection by pyruvate of rat retinal cells against zinc toxicity in vitro, and pressure-induced ischemia in vivo. Investig. Ophthalmol. Vis. Sci. 2004, 45, 1523–1530. [Google Scholar] [CrossRef]
  328. Rodilla, V.; Miles, A.T.; Jenner, W.; Hawksworth, G.M. Exposure of cultured human proximal tubular cells to cadmium, mercury, zinc and bismuth: Toxicity and metallothionein induction. Chem. Interact. 1998, 115, 71–83. [Google Scholar] [CrossRef]
  329. Cheng, J.; Liu, B.; Wu, Y.; Zheng, Y. Comparative in vitro Study on Pure Metals (Fe, Mn, Mg, Zn and W) as Biodegradable Metals. J. Mater. Sci. Technol. 2013, 29, 619–627. [Google Scholar] [CrossRef]
  330. Bordbar-Khiabani, A.; Ebrahimi, S.; Yarmand, B. In-vitro corrosion and bioactivity behavior of tailored calcium phosphate-containing zinc oxide coating prepared by plasma electrolytic oxidation. Corros. Sci. 2020, 173, 108781. [Google Scholar] [CrossRef]
  331. Yang, H.; Wang, C.; Liu, C.; Chen, H.; Wu, Y.; Han, J.; Jia, Z.; Lin, W.; Zhang, D.; Li, W.; et al. Evolution of the degradation mechanism of pure zinc stent in the one-year study of rabbit abdominal aorta model. Biomaterials 2017, 145, 92–105. [Google Scholar] [CrossRef] [PubMed]
  332. Li, H.F.; Xie, X.H.; Zheng, Y.F.; Cong, Y.; Zhou, F.Y.; Qiu, K.J.; Wang, X.; Chen, S.H.; Huang, L.; Tian, L.; et al. Development of biodegradable Zn-1X binary alloys with nutrient alloying elements Mg, Ca and Sr. Sci. Rep. 2015, 5, 10719. [Google Scholar] [CrossRef] [PubMed]
  333. Su, Y.; Cockerill, I.; Wang, Y.; Qin, Y.-X.; Chang, L.; Zheng, Y.; Zhu, D. Zinc-Based Biomaterials for Regeneration and Therapy. Trends Biotechnol. 2019, 37, 428–441. [Google Scholar] [CrossRef] [PubMed]
  334. Xia, P.; Lian, S.; Wu, Y.; Yan, L.; Quan, G.; Zhu, G. Zinc is an important inter-kingdom signal between the host and microbe. Veter. Res. 2021, 52, 1–14. [Google Scholar] [CrossRef]
  335. Wei, Y.; Wang, J.; Wu, S.; Zhou, R.; Zhang, K.; Zhang, Z.; Liu, J.; Qin, S.; Shi, J. Nanomaterial-Based Zinc Ion Interference Therapy to Combat Bacterial Infections. Front. Immunol. 2022, 13, 899992. [Google Scholar] [CrossRef]
  336. Rutherford, D.; Jíra, J.; Kolářová, K.; Matolínová, I.; Mičová, J.; Remeš, Z.; Rezek, B. Growth Inhibition of Gram-Positive and Gram-Negative Bacteria by Zinc Oxide Hedgehog Particles. Int. J. Nanomed. 2021, 16, 3541–3554. [Google Scholar] [CrossRef]
  337. Su, Y.; Wang, K.; Gao, J.; Yang, Y.; Qin, Y.-X.; Zheng, Y.; Zhu, D. Enhanced cytocompatibility and antibacterial property of zinc phosphate coating on biodegradable zinc materials. Acta Biomater. 2019, 98, 174–185. [Google Scholar] [CrossRef]
  338. Li, G.; Yang, H.; Zheng, Y.; Chen, X.-H.; Yang, J.-A.; Zhu, D.; Ruan, L.; Takashima, K. Challenges in the use of zinc and its alloys as biodegradable metals: Perspective from biomechanical compatibility. Acta Biomater. 2019, 97, 23–45. [Google Scholar] [CrossRef]
  339. Zheng, Y.F.; Gu, X.N.; Witte, F. Biodegradable metals. Mat. Sci. Eng. R 2014, 77, 1–34. [Google Scholar] [CrossRef]
  340. Venezuela, J.; Dargusch, M. The influence of alloying and fabrication techniques on the mechanical properties, biodegradability and biocompatibility of zinc: A comprehensive review. Acta Biomater. 2019, 87, 1–40. [Google Scholar] [CrossRef]
  341. Chen, K.; Lu, Y.; Tang, H.; Gao, Y.; Zhao, F.; Gu, X.; Fan, Y. Effect of strain on degradation behaviors of WE43, Fe and Zn wires. Acta Biomater. 2020, 113, 627–645. [Google Scholar] [CrossRef] [PubMed]
  342. Bowen, P.K.; Seitz, J.M.; Guillory, R.J., 2nd; Braykovich, J.P.; Zhao, S.; Goldman, J.; Drelich, J.W. Evaluation of wrought Zn-Al alloys (1, 3, and 5 wt % Al) through mechanical and in vivo testing for stent applications. J. Biomed. Mater. Res. B Appl. Biomater. 2018, 106, 245–258. [Google Scholar] [CrossRef] [PubMed]
  343. Mostaed, E.; Sikora-Jasinska, M.; Mostaed, A.; Loffredo, S.; Demir, A.; Previtali, B.; Mantovani, D.; Beanland, R.; Vedani, M. Novel Zn-based alloys for biodegradable stent applications: Design, development and in vitro degradation. J. Mech. Behav. Biomed. Mater. 2016, 60, 581–602. [Google Scholar] [CrossRef] [PubMed]
  344. Zhang, W.; Li, P.; Shen, G.; Mo, X.; Zhou, C.; Alexander, D.; Rupp, F.; Geis-Gerstorfer, J.; Zhang, H.; Wan, G. Appropriately adapted properties of hot-extruded Zn-0.5Cu-xFe alloys aimed for biodegradable guided bone regeneration membrane application. Bioact. Mater. 2021, 6, 975–989. [Google Scholar] [CrossRef]
  345. García-Mintegui, C.; Córdoba, L.C.; Buxadera-Palomero, J.; Marquina, A.; Jiménez-Piqué, E.; Ginebra, M.-P.; Cortina, J.L.; Pegueroles, M. Zn-Mg and Zn-Cu alloys for stenting applications: From nanoscale mechanical characterization to in vitro degradation and biocompatibility. Bioact. Mater. 2021, 6, 4430–4446. [Google Scholar] [CrossRef]
  346. Hu, Y.; Guo, X.; Qiao, Y.; Wang, X.; Lin, Q. Preparation of medical Mg–Zn alloys and the effect of different zinc contents on the alloy. J. Mater. Sci. Mater. Med. 2022, 33, 1–13. [Google Scholar] [CrossRef]
  347. Kubasek, J.; Dvorsky, D.; Sedy, J.; Msallamova, S.; Levorova, J.; Foltan, R.; Vojtech, D. The Fundamental Comparison of Zn-2Mg and Mg-4Y-3RE Alloys as a Perspective Biodegradable Materials. Materials 2019, 12, 3745. [Google Scholar] [CrossRef]
  348. Yang, H.; Jia, B.; Zhang, Z.; Qu, X.; Li, G.; Lin, W.; Zhu, D.; Dai, K.; Zheng, Y. Alloying design of biodegradable zinc as promising bone implants for load-bearing applications. Nat. Commun. 2020, 11, 1–16. [Google Scholar] [CrossRef]
  349. Sezer, N.; Evis, Z.; Kayhan, S.M.; Tahmasebifar, A.; Koç, M. Review of magnesium-based biomaterials and their applications. J. Magnes. Alloy 2018, 6, 23–43. [Google Scholar] [CrossRef]
  350. Liu, J.; Sonshine, D.A.; Shervani, S.; Hurt, R.H. Controlled Release of Biologically Active Silver from Nanosilver Surfaces. ACS Nano 2010, 4, 6903–6913. [Google Scholar] [CrossRef]
  351. Jiang, J.; Pi, J.; Cai, J. The Advancing of Zinc Oxide Nanoparticles for Biomedical Applications. Bioinorg. Chem. Appl. 2018, 2018, 1062562. [Google Scholar] [CrossRef] [PubMed]
  352. He, M.; Chen, L.; Yin, M.; Xu, S.; Liang, Z. Review on magnesium and magnesium-based alloys as biomaterials for bone immobilization. J. Mater. Res. Technol. 2023, 23, 4396–4419. [Google Scholar] [CrossRef]
  353. Mutlu, N.; Liverani, L.; Kurtuldu, F.; Galusek, D.; Boccaccini, A.R. Zinc improves antibacterial, anti-inflammatory and cell motility activity of chitosan for wound healing applications. Int. J. Biol. Macromol. 2022, 213, 845–857. [Google Scholar] [CrossRef] [PubMed]
  354. Gu, X.; Zheng, Y.; Cheng, Y.; Zhong, S.; Xi, T. In vitro corrosion and biocompatibility of binary magnesium alloys. Biomaterials 2009, 30, 484–498. [Google Scholar] [CrossRef] [PubMed]
  355. Levy, G.K.; Goldman, J.; Aghion, E. The Prospects of Zinc as a Structural Material for Biodegradable Implants—A Review Paper. Metals 2017, 7, 402. [Google Scholar] [CrossRef]
  356. Gong, H.; Wang, K.; Strich, R.; Zhou, J.G. In vitro biodegradation behavior, mechanical properties, and cytotoxicity of biodegradable Zn-Mg alloy. J. Biomed. Mater. Res. B Appl. Biomater. 2015, 103, 1632–1640. [Google Scholar] [CrossRef]
  357. Chen, Y.; Xu, Z.; Smith, C.; Sankar, J. Recent advances on the development of magnesium alloys for biodegradable implants. Acta Biomater. 2014, 10, 4561–4573. [Google Scholar] [CrossRef]
  358. Agarwal, S.; Curtin, J.; Duffy, B.; Jaiswal, S. Biodegradable magnesium alloys for orthopaedic applications: A review on corrosion, biocompatibility and surface modifications. Mater. Sci. Eng. C Mater. Biol. Appl. 2016, 68, 948–963. [Google Scholar] [CrossRef]
  359. Kubasek, J.; Vojtech, D.; Jablonska, E.; Pospisilova, I.; Lipov, J.; Ruml, T. Structure, mechanical characteristics and in vitro degradation, cytotoxicity, genotoxicity and mutagenicity of novel biodegradable Zn-Mg alloys. Mater. Sci. Eng. C Mater. Biol. Appl. 2016, 58, 24–35. [Google Scholar] [CrossRef]
  360. Kubasek, J.; Pospisilova, I.; Vojtech, D.; Jablonska, E.; Ruml, T. Structural, Mechanical and Cytotoxicity Characterization of As-Cast Biodegradable Zn-xMg (x = 0.8–8.3%) alloys. Mater. Tehnol. 2014, 48, 623–629. [Google Scholar]
  361. Murni, N.; Dambatta, M.; Yeap, S.; Froemming, G.; Hermawan, H. Cytotoxicity evaluation of biodegradable Zn–3Mg alloy toward normal human osteoblast cells. Mater. Sci. Eng. C Mater. Biol. Appl. 2015, 49, 560–566. [Google Scholar] [CrossRef] [PubMed]
  362. Sheu, T.-J.; Schwarz, E.M.; Martinez, D.A.; O’Keefe, R.J.; Rosier, R.N.; Zuscik, M.J.; Puzas, J.E. A Phage Display Technique Identifies a Novel Regulator of Cell Differentiation. J. Biol. Chem. 2003, 278, 438–443. [Google Scholar] [CrossRef] [PubMed]
  363. Wang, L.; Luo, Q.; Zhang, X.; Qiu, J.; Qian, S.; Liu, X. Co-implantation of magnesium and zinc ions into titanium regulates the behaviors of human gingival fibroblasts. Bioact. Mater. 2020, 6, 64–74. [Google Scholar] [CrossRef] [PubMed]
  364. Sazegar, G.; Reza, A.H.S.; Behravan, E. The effects of supplemental zinc and honey on wound healing in rats. Iran. J. Basic Med. Sci 2011, 14, 391–398. [Google Scholar]
  365. Lansdown, A. Influence of zinc oxide in the closure of open skin wounds. Int. J. Cosmet. Sci. 1993, 15, 83–85. [Google Scholar] [CrossRef]
  366. Apelqvist, J.; Larsson, J.; Lstenstrom, A. Topical treatment of necrotic foot ulcers in diabetic patients: A comparative trial of DuoDerm and MeZinc. Br. J. Dermatol. 1990, 123, 787–792. [Google Scholar] [CrossRef]
  367. Agren, M. Zinc oxide increases degradation of collagen in necrotic wound tissue. Br. J. Dermatol. 1993, 129, 221–222. [Google Scholar] [CrossRef]
  368. Nagase, H.; Visse, R.; Murphy, G. Structure and function of matrix metalloproteinases and TIMPs. Cardiovasc. Res. 2006, 69, 562–573. [Google Scholar] [CrossRef]
  369. Moreno-Eutimio, M.A.; Espinosa-Monroy, L.; Orozco-Amaro, T.; Torres-Ramos, Y.; Montoya-Estrada, A.; Hicks, J.J.; Rodríguez-Ayala, E.; Del Moral, P.; Moreno, J.; Cueto-García, J. Enhanced healing and anti-inflammatory effects of a carbohydrate polymer with zinc oxide in patients with chronic venous leg ulcers: Preliminary results. Arch. Med. Sci. 2016, 14, 336–344. [Google Scholar] [CrossRef]
  370. Mirastschijski, U.; Haaksma, C.J.; Tomasek, J.J.; Ågren, M.S. Matrix metalloproteinase inhibitor GM 6001 attenuates keratinocyte migration, contraction and myofibroblast formation in skin wounds. Exp. Cell Res. 2004, 299, 465–475. [Google Scholar] [CrossRef]
  371. A Cabral-Pacheco, G.; Garza-Veloz, I.; La Rosa, C.C.-D.; Ramirez-Acuña, J.M.; A Perez-Romero, B.; Guerrero-Rodriguez, J.F.; Martinez-Avila, N.; Martinez-Fierro, M.L. The Roles of Matrix Metalloproteinases and Their Inhibitors in Human Diseases. Int. J. Mol. Sci. 2020, 21, 9739. [Google Scholar] [CrossRef] [PubMed]
  372. Ågren, M.S.; Andersen, L.; Heegaard, A.M.; Jorgensen, L.N. Effect of parenteral zinc sulfate on colon anastomosis repair in the rat. Int. J. Color. Dis. 2008, 23, 857–861. [Google Scholar] [CrossRef] [PubMed]
Figure 2. Schematic sequence of tissue expander inflation. Initially, the skin is in a state of rest, with no tension present (top). A tissue expander which is inserted underneath the skin, starts to inflate between the epidermis and dermis layers and the hypodermis. Upon inflation of the expander, the skin becomes taut and stretched (middle). The mechanical stretching results in cellular proliferation, leading to either skin growth and proliferation or apoptosis or tumor formation. Eventually, the skin grows enough to return to its original state of rest, with no tension (bottom) [34]. Parts of the figure were drawn using elements from Servier Medical Art. (https://creativecommons.org/licenses/by/3.0/) (accessed on 16 May 2022).
Figure 2. Schematic sequence of tissue expander inflation. Initially, the skin is in a state of rest, with no tension present (top). A tissue expander which is inserted underneath the skin, starts to inflate between the epidermis and dermis layers and the hypodermis. Upon inflation of the expander, the skin becomes taut and stretched (middle). The mechanical stretching results in cellular proliferation, leading to either skin growth and proliferation or apoptosis or tumor formation. Eventually, the skin grows enough to return to its original state of rest, with no tension (bottom) [34]. Parts of the figure were drawn using elements from Servier Medical Art. (https://creativecommons.org/licenses/by/3.0/) (accessed on 16 May 2022).
Polymers 15 03854 g002
Figure 3. Essential characteristics of an optimal metallic implant [207].
Figure 3. Essential characteristics of an optimal metallic implant [207].
Polymers 15 03854 g003
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

Hassan, N.; Krieg, T.; Zinser, M.; Schröder, K.; Kröger, N. An Overview of Scaffolds and Biomaterials for Skin Expansion and Soft Tissue Regeneration: Insights on Zinc and Magnesium as New Potential Key Elements. Polymers 2023, 15, 3854. https://doi.org/10.3390/polym15193854

AMA Style

Hassan N, Krieg T, Zinser M, Schröder K, Kröger N. An Overview of Scaffolds and Biomaterials for Skin Expansion and Soft Tissue Regeneration: Insights on Zinc and Magnesium as New Potential Key Elements. Polymers. 2023; 15(19):3854. https://doi.org/10.3390/polym15193854

Chicago/Turabian Style

Hassan, Nourhan, Thomas Krieg, Max Zinser, Kai Schröder, and Nadja Kröger. 2023. "An Overview of Scaffolds and Biomaterials for Skin Expansion and Soft Tissue Regeneration: Insights on Zinc and Magnesium as New Potential Key Elements" Polymers 15, no. 19: 3854. https://doi.org/10.3390/polym15193854

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