Next Article in Journal
Construction of Artificial Ovaries with Decellularized Porcine Scaffold and Its Elicited Immune Response after Xenotransplantation in Mice
Next Article in Special Issue
Interstitial Fluid Flows along Perivascular and Adventitial Clearances around Neurovascular Bundles
Previous Article in Journal
Evaluation of the Microbiological Effect of Colloidal Nanosilver Solution for Root Canal Treatment
Previous Article in Special Issue
Finite Element Analysis of the Non-Uniform Degradation of Biodegradable Vascular Stents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Zinc-Based Biodegradable Materials for Orthopaedic Internal Fixation

1
Department of Biomedical Engineering, Faculty of Environment and Life, Beijing University of Technology, Beijing 100124, China
2
School of Metallurgy, Northeastern University, Shenyang 110819, China
*
Author to whom correspondence should be addressed.
J. Funct. Biomater. 2022, 13(4), 164; https://doi.org/10.3390/jfb13040164
Submission received: 2 August 2022 / Revised: 15 September 2022 / Accepted: 21 September 2022 / Published: 26 September 2022

Abstract

:
Traditional inert materials used in internal fixation have caused many complications and generally require removal with secondary surgeries. Biodegradable materials, such as magnesium (Mg)-, iron (Fe)- and zinc (Zn)-based alloys, open up a new pathway to address those issues. During the last decades, Mg-based alloys have attracted much attention by researchers. However, the issues with an over-fast degradation rate and release of hydrogen still need to be overcome. Zn alloys have comparable mechanical properties with traditional metal materials, e.g., titanium (Ti), and have a moderate degradation rate, potentially serving as a good candidate for internal fixation materials, especially at load-bearing sites of the skeleton. Emerging Zn-based alloys and composites have been developed in recent years and in vitro and in vivo studies have been performed to explore their biodegradability, mechanical property, and biocompatibility in order to move towards the ultimate goal of clinical application in fracture fixation. This article seeks to offer a review of related research progress on Zn-based biodegradable materials, which may provide a useful reference for future studies on Zn-based biodegradable materials targeting applications in orthopedic internal fixation.

1. Introduction

Bone fractures are becoming increasingly common with the rapid increases in aging population, traffic accidents, sports injuries and metabolic diseases [1,2,3]. Fractures have a lifetime prevalence of ~40% and an annual incidence of 3.6% [4]. The most common and burdensome fractures are lower leg fractures of the patella, tibia or fibula, or ankle [5]. Fracture healing is the process of reconstructing bone and restoring its biological and biomechanical functions [6,7]. As one of the common surgical treatments, internal fixation using screws, pins, plates, etc., provides mechanical stability for a fractured bone, allowing weight bearing, early use of the limb, and bone healing [8]. Success in fracture healing is closely related to the internal fixation implants used.
Implants used for internal fixation can be divided into several categories: wires, pins and screws, plates, and intramedullary nails or rods [9]. Staples and clamps are also used occasionally for osteotomy or fracture fixation [9]. Traditional fixation materials are generally nondegradable, including inert stainless steel (SS), titanium (Ti) and its alloys, and cobalt-chromium (Co-Cr) alloys [10]. They possess satisfactory biocompatibility, high wear resistance, and adequate mechanical strength (Table 1) [11,12,13]. However, they have notable shortcomings when being applied in fracture fixation. For example, metallic materials have much higher elastic modulus values (190–200 GPa for 316L SS, 210–240 GPa for Co-Cr alloys, and 90–110 GPa for Ti alloys) compared with bone tissues (3–30 GPa). Although a rigid fixation is required at the beginning of the healing process to provide a sufficient mechanical stability, a large discrepancy in stiffness between bone and the implant can lead to stress shielding and therefore can delay healing [14]. Even for a successful bone healing, a secondary surgery is often required to remove the implant [15].
Biodegradable materials are well suited to solve the issues above. Fixation implants made of biodegradable materials can provide a strong mechanical support of the fracture site at earlier stages of the healing process, and later on degrade naturally as the healed bone takes over the mechanical loading and their by-products can be absorbed and metabolized [16]. Degradable polymers are intended for applications in soft tissue graft fixation and meniscus repair due to their low strength [17,18,19]. Compared to polymers, Mg-based biodegradable materials have higher strength and modulus that are close to cortical bone (Table 1). Also, Mg ions released from Mg-based biodegradable implants have beneficial effects on bone regeneration [20]. Due to their appropriate mechanical property, biocompatibility and biodegradability, Mg-based metals have attracted a great deal of attention of in vitro and in vivo research during the last decades. Several Mg-based implants (bone screws, pins, plates) have been available in clinic or undergoing clinical trials [15,21]. However, the issues with an over-fast degradation rate and generation of hydrogen still need to be overcome. Additionally, current Mg alloys (ultimate tensile strength (UTS) 350 MPa) have relatively low mechanical strength and are only limited to non- or low-load-bearing applications, such as fixation of small bones and cancellous fragments, meniscus repair and soft tissue fixation [21]. Clearly, there remains a critical need for development of biodegradable materials for fixation of fractures at heavy load-bearing skeletal sites where fractures occur most frequently.
The mechanical strength of zinc (Zn) alloys falls in a wide range, from the value of pure Mg to the value of commercial pure Ti and 316 stainless steel (Figure 1). For bone repair, it has been reported that the degradation rates of fixation implants should be between 0.2 and 0.5 mm y−1 to match bone healing [1]. Mg-based alloys have degradation rates ranging from 0.8 to 2.7 mm y−1 [1,15,22,23], which are above the desired degradation rates of bone implants. The degradation rates of Zn-based alloys are mainly between 0.1 and 0.3 mm y−1 [1,24,25]. Moderate corrosion rates and excellent mechanical properties make Zn-based biodegradable metals potential candidates for biomaterial for internal fracture fixation, particularly at heavy load-bearing sites [26,27,28]. In terms of biocompatibility, Zn is the second most abundant transition metal in humans, serving as a structural or enzymatic cofactor for approximately 10% of the proteome [29]. Consequently, perturbations in Zn homeostasis may lead to various disorders, including growth deficiencies, immune defects, neurological disorders, and cancers [30]. Studies also found that Zn ions (Zn2+) play an important role in promoting fracture healing [30,31].
Zn-based alloys have shown a great potential of application in orthopaedics, particularly for internal fixation of fractures at heavy load-bearing bone [24,25,32,33]. There has been a growing body of in vitro studies on the development of new Zn-based biodegradable materials and testing of their biodegradability, mechanical property and biocompatibility, with fewer in vivo animal studies and no clinical application as yet [34,35,36,37,38,39,40,41,42,43,44,45,46] (Figure 2). Although there are several review articles that have elaborated on some aspects of those properties [14,22,26,27,47], it remains unclear if current Zn-based biodegradable material are sufficient to meet clinical needs for orthopaedic internal fixation and what research gap needs to be filled next. In the following sections, we first point out the clinical requirements of implant biomaterials for orthopaedic internal fixation primarily at the heavy load-bearing skeletal sites in terms of their biodegradability, mechanical property and biocompatibility, and then summarize various typical Zn-based biodegradable materials (pure Zn, Zn-based alloys and composites) that have been developed so far and examined in vitro and in vivo for each of these properties. Lastly, unaddressed questions or future research directions are discussed with the aim of moving towards clinical applications of Zn-based biodegradable materials for orthopaedic internal fixation.
Table 1. Characteristics of different typical metallic biomaterials.
Table 1. Characteristics of different typical metallic biomaterials.
ClassificationMaterialsBiodegradabilityMechanical PropertiesBiocompatibilityApplications or Potential ApplicationsRef.
Non-biodegradable metallicmaterials316L SSNon-biodegradableHigh elastic modulus, low wear and corrosion resistance, high tensile strengthHigh biocompatibilityAcetabular cup, bone screws, bone plates, pins, etc.[11]
Co–Cr alloysNon-biodegradableHigh elastic modulus, high wear and corrosion resistanceLow biocompatibilityBone screws, bone plates, femoral stems, total hip replacements, etc.[12]
Ti alloysNon-biodegradablePoor fatigue strength, light weightHigh biocompatibilityDental implants, bone screws, bone plates, etc.[11,13]
Biodegradable metallic materialsMg-based alloysBiodegradable, high degradation ratePoor mechanical properties, elastic modulus are close to cortical boneHigh biocompatibility, H2 evolutionBone screws, bone plates (non-load bearing parts), etc.[2,9]
Fe-based alloysBiodegradable, low degradation rateHigh elastic modulus, poor mechanical propertiesLow biocompatibilityBone screws, bone plates, etc.[9]
Zn-based alloysBiodegradable, moderate corrosion rateHigh elastic modulus, high mechanical properties, low creep resistanceCytotoxicity, no gas production, high biocompatibilityBone screws, bone plates (load-bearing parts (potential applications)), etc.[3,9,10]

2. Biodegradability of Zn-Based Biodegradable Materials

It is well known that human bodies are full of fluid solutions and bear mechanical loading, generating corrosive and mechanical environments for biodegradable materials. It is expected that after biodegradable metals are implanted in the human body, they can gradually degrade at a suitable rate that matches the healing rate of bone tissues (Figure 3). However, different types of fractures at different skeletal sites require different fixation implants (as well as the amounts of degradable materials). Therefore, considering designs of suitable Zn-based devices with a proper degradation rate to meet clinical fixation requirements of different fractures, it is necessary to understand the degradation mechanisms, regulation of the degradation rate, and mechanical factors influencing the degradation of biodegradable metals. It is evident that pure Zn has a relatively low degradation rate. Adding alloy elements or reinforcement materials is commonly used to tune the biodegradability of Zn.

2.1. Biodegradability of Pure Zn

Zn is a relatively low active metal. The standard electrode potential of Zn is −0.76 V, which lies between those of Mg (−2.37 V) and Fe (−0.44 V) [14]. It is prone to corrode in various fluid environments within the human body [22]. Studies have been extensively focused on the in vitro corrosion behavior of pure Zn in different corrosion media, such as Hank’s balanced salt solution (HBSS), phosphate buffer saline (PBS), and stimulated body fluid (SBF) [16,28,48]. The medium electrode potential of Zn is associated with a moderate corrosion rate of approximately 0.1 mm y−1 [28]. The corrosion mechanism of Zn is regulated by the following reactions:
Zn→Zn2+ + 2e,
2H2O + O2 + 4e→4OH,
2Zn + 2H2O + O2→2Zn(OH)2,
Zn(OH)2→ZnO + H2O,
Following the anode reaction, the Zn loses two electrons to generate the Zn2+. Cathode reaction is a process where the electrons of hydrogen dissolve oxygen reduction in the electrolyte to produce hydroxide (OH). The simultaneous increases of Zn2+ and OH in the solution facilitate the precipitation of Zn(OH)2, but the Zn(OH)2 is unstable and may subsequently transform into a thermodynamically more stable ZnO [34,49]. It is evident from these series of reactions above that Zn does not release hydrogen gas during biodegradation like Mg, indicating one of the major benefits of Zn.

2.2. Biodegradability of Zn-Based Alloys

It has been reported that the degradation rate of bone implants should be somewhere between 0.2 and 0.5 mm y−1 to match bone healing [1]. Hence, pure Zn clearly does not meet the requirements of biodegradable orthopaedic implants. Adding other alloying elements is one way to alter the corrosion rate of biodegradable metals. To establish binary Zn alloy systems, some studies added alloying elements that are beneficial for bone health (e.g., Mg, Ca, Sr, Li, Mn, Fe, Cu, and Ag) into Zn-based alloys (Table 2) [24,32,40,42,44,50,51,52,53,54]. They used the same melting and extrusion process to prepare a variety of binary Zn-based alloys. Various alloying elements affect the corrosion rate of Zn-based alloys to different degrees [24]. Alloying can lead to an accelerated degradation, and Fe, Ag and Cu had the most significant roles in accelerating corrosion, followed by Li, Sr, Ca and Mg (Table 2) [24]. Studies have shown that corrosion rates of current Zn-based alloys are mainly between 0.1 and 0.3 mm y−1 [1]. The choice of Zn-based alloys generally depends on clinical demands, considering the fracture site, and the shape and size of fixators. In general, the degradation rate of the currently developed Zn alloy system is relatively slow. More ternary and quaternary Zn alloys may need to be developed to match the rate of fracture healing.

2.3. Biodegradability of the Zn-Based Composites

Other than alloying, adding reinforcement to form Zn matrix composites is another way to regulate the degradation rate of Zn metals. Composites including pure Zn as a matrix and hydroxyapatite (HA) as reinforcements were prepared by spark plasma sintering (SPS) [55]. A wide range of degradation rates (0.3–0.85 mm y−1) can be achieved by changing the concentration of HA. In addition, the immersion experiment of another beta-tricalcium phosphate (β-TCP)/Zn-Mg composite showed that the corrosion resistance of the composite is slightly decreased with the increase in β-TCP content [56].
Table 2. In vitro experiments of binary Zn-based alloys.
Table 2. In vitro experiments of binary Zn-based alloys.
Composition (wt%)Mechanical PropertiesCorrosion TestCytocompatibilityRef.
σYS (MPa)σUTS (MPa)ε (%)Corrosion MediumCorrosion
Rate (mm y−1)
Cell TypeKey Findings
Zn-0.8Mg20330113MEM0.071U-2OS,
L-929
Zn is less biocompatible than magnesium and the maximum safe concentrations of Zn2+ for the U-2OS and L929 cells are 120 μM and 80 μM.[50]
Zn-1.0Ca20625212.7HBSS0.09MG63Adding the alloying elements Ca into Zn can significantly increase the viability of MG63 and can promote the MG63 cell proliferation compared with the pure Zn and negative control groups.[51]
Zn-1.1Sr22025022SBF0.4HOBs, hMSCsThe proliferation ability of the two kinds of cells did not decrease in the zinc alloy leaching solution. When the concentration of the leaching solution was low, the growth of the two kinds of cells was promoted.[32]
Zn-0.4Li3875205.0SBF0.019MC3T3-E1Zn-0.4Li alloy extract can significantly promote the proliferation of MC3T3-E1 cells.[24]
Zn-5.0Ge17523722HBSS0.051MC3T3-E1The diluted extracts at a concentration 12.5% of both the as-cast Zn-5Ge alloy and pure Zn showed grade 0 cytotoxicity; the diluted extracts at the concentrations of 50% and 25% of Zn-5Ge alloy showed a significantly higher cell viability than those of pure Zn.[52]
Zn-6.0Ag-290-SBF0.114--[44]
Zn-0.8Fe12716328.1SBF0.022MC3T3-E1MC3T3-E1 cells had unhealthy morphology and low cell viability.[24]
Zn-4Cu32739344.6HBSS0.13L-929, TAG,
SAOS-2
Zn-4Cu alloy had no obvious cytotoxic effect on L929, TAG and Saos-2 cells.[53]
Zn-0.8Mn98.4104.71.0--L-929Zn-0.8Mn alloy showed 29% to 44% cell viability in 100% extract, indicating moderate cytotoxicity.[40]
Zn-2Al14219212SBF0.13MG63Cell viability decreased to 67.5 ± 5.3% in 100% extract cultured for one day, indicating that high concentrations of ions have a negative effect on cell growth. With the extension of culture time, the number of cells increased significantly.[42]
Zn-0.0.5Zr10415722----[54]
YS: yield Strength; UTS: ultimate tensile strength; SBF: stimulated body fluid; MEM: minimum essential medium; HBSS: Hank’s balanced salt solution; L-929: mouse fibroblasts; MG63: human osteosarcoma cells; HOBs: human osteoblasts; MSCs: human bone marrow mesenchymal stem cells; MC3T3-E1: mouse preosteoblasts; TAG: human immortalized periosteal cells; SAOS-2: human osteosarcoma cells; U-2OS: human osteosarcoma cells.

2.4. Biodegradability of Zn-Based Biomaterials under Mechanical Loading

The response of biodegradable materials to the combined effect of physiological loading and corrosion environment is an important issue in vivo since stress-induced degradation and cracking are common [57]. Particularly, for load-bearing fracture fixation where biodegradable implant undertakes loading, it is critical to understand how mechanical stress affects the biodegradation behavior of the implant.
The combination of mechanical loading and a specific corrosive medium environment can lead to sudden cracking and failure of degradable metals. This phenomenon is called stress corrosion cracking [57]. In vivo animal experiments and clinical studies have indicated the role of mechanical stress in the early failure of biodegradable implants [19,57,58,59]. Li et al. conducted slow-strain rate testing (SSRT) and constant-load immersion tests on a promising Zn-0.8 wt%Li alloy [60]. They investigated its stress corrosion cracking susceptibility and examined its feasibility as biodegradable metals with pure Zn serving as a control group. They observed that the Zn-0.8 wt%Li alloy exhibited a low stress corrosion cracking susceptibility. This was attributed to variations in microstructure and deformation mechanism after alloying with Li. In addition, compared to the “no stress” condition (0.124 mm y−1), the corrosion rate of the Zn-Li alloy only increased slightly under tensile stress of 11.1 MPa (0.129 mm y−1) and compressive stress of 17.7 MPa (0.125 mm y−1). Both pure Zn and Zn-0.8 wt%Li alloy did not fracture over a period of 28 days during the constant-load immersion test. The magnitude of the applied stress was close to the physiological loading condition and thus the authors proved the feasibility of both materials as biodegradable metals. So far, there are only a few experimental studies on the stress corrosion of Zn-based biodegradable materials. Since previous experimental studies have shown in degradable polymers or Mg-based alloys that the corrosion rate is affected by the loading mode (tension or compression) and magnitude [18,61,62,63,64], it is assumed that these effects also exist in Zn-based biodegradable materials. Identification of the quantitative relationships between various forms of applied loading and degradation behaviors of Zn-based materials is important for the design of load-bearing fixation implants. However, the corrosion behaviors of Zn-based biomaterials under different loading conditions need to be further explored.
Cyclic loading-induced fatigue fractures are very common in engineered metals, where the fatigue strength is further reduced in a corrosive environment [65]. It was reported that under the combined effects of stress and corrosive media, fatigue cracks propagate faster [57]. Corrosion fatigue is of primary concern for metallic internal fixation which commonly bear cyclic dynamic loads in vivo. The corrosion pit propagation rate is influenced by the magnitude of stress, frequency, and cycle number [66,67]. Previous studies have compared the compression-induced fatigue behavior of additively manufactured porous Zn in air and in revised simulated body fluid (r-SBF) [68]. The fatigue strength of the additively-manufactured porous Zn was high in air (i.e., 70% of its yield strength) and even higher in r-SBF (i.e., 80% of its yield strength). The high value of the relative fatigue strength in air could be attributed to the high ductility of pure Zn itself. The formation of corrosion products around the strut junctions might explain the higher fatigue strength of additive manufacturing Zn in r-SBF. The favorable fatigue behavior of additive manufacturing porous Zn further highlights its potential as a promising bone-substituting biomaterial. Another study found in their fatigue testing of Zn-0.5Mg-WC nanocomposites that the material survived after 10 million cycles of tensile loading when the maximum stress was 80% of the yield stress [69]. These results suggest that the Zn-0.5Mg-WC nanocomposite is a promising candidate for biodegradable materials. So far, there has been no report on the fatigue corrosion behavior of Zn-based alloys in vivo. Since the resistance of a material to fatigue and corrosion is an important consideration for designing implants, future relevant studies on Zn-based biomaterials may be required.

3. Mechanical Properties of Zn-Based Biodegradable Materials

In addition to the biodegradable properties, the mechanical properties of the biodegradable metals are also important considerations for designing orthopaedic implants for internal fixation. Yield strength (YS), ultimate tensile strength (UTS), elongation (ε) and elastic modulus (E) are common parameters which are used to indicate the mechanical properties of biomedical materials [11,37,70,71,72,73]. Extensive studies have determined those mechanical parameters of Zn-based biodegradable materials. The reported mechanical criteria for degradable metals (e.g., Mg-based) are UTS 300 MPa and ε 20% [22]. On the other hand, the current gold standard for medical metal materials, such as Ti and its alloys, has a tensile strength of over 600 MPa [13]. To certain extent, these criteria could provide guides of mechanical properties for development of Zn-based degradable materials. However, the requirement may vary with different load-bearing sites.

3.1. Mechanical Properties of Pure Zn

Pure Zn has extremely low yield strength (29.3 MPa) and elongation (1.2%) in its as-cast condition [74]. The Young’s modulus of pure Zn is around 94 GPa [16]. Obviously, it is difficult to meet the mechanical criteria as biodegradable metals [22]. On the other hand, owing to the low melting point of Zn, several additional uncertainties exist with regard to the mechanical properties of biodegradable Zn and Zn-based alloys. Low creep resistance, high susceptibility due to natural aging, and static recrystallization may lead to the failure of Zn-based biodegradable materials during storage at a room temperature and usage at a body temperature [26]. Studies showed that Zn-based alloys underwent appreciable creep deformation under human body temperature (37°) [75]. In addition, recrystallization of Zn-based alloys under stress can reduce their resistance to creep [42]. Thus, creep deformation is an important factor that should be considered in the studies of pure Zn.

3.2. Mechanical Properties of Zn-Based Alloys

Alloying is a common approach to change the mechanical properties of metals, where alloy ratio is essential for studies of Zn-based alloys. Attempts have been made to optimizing the Zn-based alloys by changing the alloy ratio, in order to obtain better mechanical performance in vitro and then move to in vivo conditions [24,32,40,42,44,50,51,52,53,54]. Zn-based alloys have Young’s modulus values ranging from 100 to 110 GPa depending on alloying conditions [16]. As summarized in Table 2, the Zn-based alloys with improved mechanical properties to various degrees are generated by adding elements of Mg [50], Ca [51], Sr [32], Li [24], Ge [52], Ag [44], Fe [24], Cu [53], Mn [40], Al [42], Zr [54]. The improvement of adding Li elements is particularly obvious, but the elongation of Zn-Li is only 5%. Following addition of the Cu element, the elongation of the Zn-based alloys reaches 44.6%. Binary Zn-based alloys have poor mechanical properties and may not be applicable in load-bearing sites of the skeleton. Table 3 summarizes the mechanical properties of ternary Zn-based alloys on the basis of binary Zn-based alloys.
Different mechanical processing methods have great influences on the mechanical properties of the same Zn-based alloys. Among the three common mechanical processing operations (hot extrusion, hot rolling, and casting), the hot extrusion can produce the greatest improvement in mechanical properties of Zn-based alloys. Compared with binary Zn-based alloys, ternary Zn-based alloys have largely improved mechanical properties. For example, the tensile strength of Zn-0.8Li-0.4Mg is 646 MPa, which is greater than those of pure titanium or 316L SS (Figure 1) [24]. In addition, reasonable mechanical integrity of Zn-0.8Li-0.4Mg was maintained in vitro, and is expected to be used for bone repair at load-bearing sites.

3.3. Mechanical Properties of Zn-Based Composites

Apart from the addition of alloying elements, adding reinforcement matrix as composite could also regulate the mechanical properties of Zn metals. The biocompatibility and the mechanical properties were improved by controlling the type and content of the second phase to form a composite material. In a previous study, Zn-HA composites were prepared with pure Zn as matrix and hydroxyapatite (HA) as reinforcement by spark plasma sintering [55]. In vivo tests showed that the addition of HA resulted in a better performance in osteogenesis with prolonged fixation time. In another study, Zn-Mg-β-TCP composites were prepared with Zn-Mg as matrix and β-TCP as enhancer by the mechanical stirring combined with ultrasonic assisted casting and hot extrusion technology [56]. This material had an ultimate tensile strength of 330.5 MPa and showed better biocompatibility than Zn-Mg alloys in cellular experiments. A barrier layer of ZrO2 nanofilm was constructed on the surface of Zn-0.1 wt%Li alloy via atomic layer deposition (ALD) [76]. Their results indicated that the addition of ZrO2 could effectively improve cell adhesion and vitality, and promote osseointegration, but the non-degradation of ZrO2 brought new challenges. Composites often have advantages over alloys due to the addition of second-phase enhancers. Compared with pure Zn, the addition of a second-phase material largely enhances its mechanical strength and biocompatibility. However, it was reported that the ductility of Zn-based composite materials is only 10% or even lower, with a greater brittleness [22], bringing difficulties to the processing of orthopaedic devices (such as bone screws and bone plates). In addition, the complex manufacturing process, high cost of composite materials, and a lack of sufficient basic theoretical supports in the field of preparation and processing still limit their developments.
Table 3. In vitro experiments of ternary Zn-based alloys.
Table 3. In vitro experiments of ternary Zn-based alloys.
Composition (wt%) and Manufacturing ProcessMechanical PropertiesCorrosion TestCytocompatibilityRef.
σYS (MPa)σUTS (MPa)ε (%)Corrosion MediumCorrosion
Rate (mm/y)
Cell TypeKey Findings
Zn-1.5Mg-0.5Zr
HE
35042512--L-929Overall, the L-929 cells exhibit polygonal or spindle shape, and well spread and proliferated in the extracts of pure Zn and Zn alloys.[39]
Zn-1.0Ca-1Sr
Cast
861401.2SBF-MG63Adding the alloying elements Mg, Ca and Sr into Zn can significantly increase the viability of MG63 and can promote the MG63 cell proliferation compared with the pure Zn and negative control groups.[77]
Zn-1.0Ca-1Sr
HE
2122606.7SBF0.11
Zn-1.0Ca-1Sr
HR
1442038.8SBF-
Zn-0.8Li-0.4Mg
HE
4386463.68----[24]
Zn-3Ge-0.5Mg
Cast
66.988.31.4HBSS0.062MG63The extract with a concentration of 100% had obvious cytotoxicity to MG63 cells. When the concentration of the extract was diluted to 12.5% or lower, the survival rate of MG-63 cells was all above 90%.[78]
Zn-3Ge-0.5Mg
HR
2532089.2HBSS0.075
Zn-4Ag-0.6Mn
HE
-30235HBSS0.012--[79]
Zn-1Fe-1Mg
Cast
1461572.3SBF0.027--[80]
Zn-0.8Mn-0.4
Cast
1121200.3----[68]
Zn-0.8Mn-0.4
HE
2533438--
Zn-0.8Mn-0.4
HR
24532312--
HE: Hot extrusion; HR: Hot rolling.

4. Biocompatibility of Zn-Based Biodegradable Materials

Biocompatibility is the ability of a material to conform to the host response, cell response, and living systems, and it is a vital property of metallic internal fixation for bone repair [1]. The metallic fixation implants directly release ions into the human body, affecting the surrounding cells, tissues, and blood (Figure 4) and leading to either positive or negative results [47,81,82]. Additionally, biomaterial-induced infections are one of the leading causes of implant failure in orthopaedic surgery [25]. Postoperative wound infection may cause an increase in the cost of pain treatment and even sequelae such as limb malformation and dysfunction of the implants [26]. Thus, exploring the biocompatibility of Zn and its alloys is important considering their ultimate implanting in the human body.

4.1. Biocompatibility of Pure Zn

Zn plays a fundamental role in multiple biochemical functions of the human body, including cell division, cell growth, wound healing, and the breakdown of carbohydrates [83]. Dietary Zn2+ deficiency has been linked to impaired skeletal development and bone growth in humans and animals (Figure 5) [83]. Specifically, 85% of Zn in the human body is found in muscle and bone, 11% in the skin and liver, and the rest in other tissues [84]. Zn is located at sites of soft tissue calcification, including osteons and calcified cartilage. Zn levels in bone tissue increase as bone mineralization increases. The skeletal growth was reduced during Zn deficiency. Zn plays a key biological role in the development, differentiation and growth of various tissues in the human body [85], including nucleic acid metabolism, stimulation of new bone formation, signal transduction, protection of bone mass, regulation of apoptosis, and gene expression [14]. Zn not only inhibits related diseases such as bone loss and inflammation, but also plays an important role in cartilage matrix metabolism and cartilage II gene expression [86]. The following symptoms are associated with Zn deficiency, including impaired physical growth and development in infants and young adults, the increased risk of infection, the loss of cognitive function, the problems of memory and behavioral, and learning disability. However, excessive Zn may cause neurotoxicity problems [87]. Based on the RDI (Reference Daily Intake) values reported for mature adults, the biocompatibility of Zn (RDI: 8–20 mg/day) is not as good as that of Mg (RDI: 240–420 mg/day), but very similar to that of Fe (RDI: 8–18 mg/day) [88]. Excessive Zn can cause symptoms such as nausea, vomiting, abdominal pain, diarrhea, fatigue, and can weaken immune function and delay bone development [87]. Therefore, when Zn-based biodegradable materials are implanted into the body as bone implant materials, the toxicity of their degradation products should be considered.

4.2. Biocompatibility of Zn-Based Alloys

The results of cytotoxicity tests can reflect the biological safety of the material to some extent. Table 2 and Table 3 summarize the results of cytocompatibility testing of alloying elements. Specifically, according to the cytocompatibility testing, additions of Mg, Ca and Li do not produce cytotoxicity, but can promote cell proliferation. However, Cu, Al, and Fe show varying degrees of toxic effects on bone cells [34,37,40,42,43,73,74,89].
Regarding the effect of metal ions on antibacterial activity, Sukhodub et al. [90] systematically examined the antibacterial abilities of metal ions and reported that the sterilization rate of the metal ions from high to low was as follows: Ag+, Cu2+, Zn2+, and Mg2+. Among these metal ions, Zn ions have a good antibacterial ability when they reach a certain concentration and can kill various bacteria and fungi. Zinc is an essential element with intrinsic antibacterial and osteoinductive capacity [91]. Zn-based antimicrobial materials generally consist of zinc complexes and ZnO nano-particles. Complexes such as zinc pyrithione and its derivatives are well known antifungal compounds and have been broadly applied in medicines [92]. Lima et al. [93] prepared Zn-doped mesoporous hydroxyapatites (HAps) with various Zn contents by co-precipitation using a phosphoprotein as the porous template. They found that the antibacterial activity of the HAps samples depended strongly on their Zn2+ contents. Tong et al. [94] examined the bacterial distributions of the Zn-Cu foams pre- and post-heat treatment after co-culturing with staphylococcus aureus for 24 h, and observed good antibacterial properties of the Zn-Cu foams. Lin et al. [74] observed better antibacterial properties of Zn-1Cu-0.1Ti than pure Zn. Ren et al. [95] systematically investigated a variety of Cu-containing medical metals including stainless steels, Ti alloys, and Co-based alloys, and demonstrated good antibacterial abilities of those materials stemming from the durable and broad-spectrum antibacterial characteristics of Cu ions. Therefore, Cu-containing Zn alloys may be expected to be promising implant materials with intrinsic antibacterial ability.

4.3. Biocompatibility of Zn-Based Composites

HA is a well-known bioceramic with bioactivity that supports cell proliferation, bone ingrowth and osseointegration. HA has similar chemical and crystallographic structures to bone, which can form a chemical bond with osseous tissue, and act like nucleation for new bone [17]. Yang et al. [55] fabricated Zn-(1, 5, 10 wt%) HA composites using the SPS technique and investigated their in vitro degradation behaviors. Zn-HA composites showed significantly improved cell viability of osteoblastic MC3T3-E1 cells compared with pure Zn. An effective antibacterial property was observed as well. As a bioactive ceramic, β-TCP has good biocompatibility, osteoconductivity and biodegradability [96]. In a study by Pan et al. [56], the biocompatibility of Zn-1Mg-xβ-TCP (x = 0, 1, 3, 5 vol%) composites were investigated. When L-929 and MC3T3 cells were cultured in different concentrations for one day, the relative proliferation rate of the cells is above 80%, and the cytotoxicity is 0–1. Moreover, the addition of β-TCP makes the compatibility of the composite material to MC3T3 cells significantly higher than that of the Zn-Mg alloy.

5. In Vivo Evaluation of Zn-Based Biodegradable Materials with Animal Models

In addition to in vitro testing, in vivo animal experiments are a necessary step in assessing the performances of Zn-based biodegradable materials prior to translation into clinical applications. Different from in vitro experiments where the biodegradability, mechanical property, and biocompatibility of a material are often tested separately, animal models can be used to examine all these properties together in an in vivo condition. Although the in vivo animal experiments may not be able to fully mimic the mechanical, biological and chemical environments in the human body, they are currently the best way to evaluate the interactions between Zn-based biodegradable materials and host [15,24,25,36,97]. There are far fewer in vivo studies on Zn-based biomaterials than in vitro studies. Several representative in vivo studies on Zn-based biodegradable materials were summarized in Table 4.
Yang et al. [24] implanted the pure Zn into the rat femur condyle. A serious fibrous tissue encapsulation was found for pure Zn, resulting in the lack of direct bonding between bone and implant (Figure 6a). The delayed osseointegration of pure Zn is claimed to be attributed to the local high Zn ion concentration. Consistent with the observations in vitro, the in vivo results confirmed that alloying with appropriate elements such as Mg, Ca and Sr can effectively improve the biocompatibility. Yang et al. [55] implanted the pure Zn into the rat femur condyle. A serious fibrous tissue encapsulation was found for pure Zn, resulting in the lack of direct bonding between bone and implant (Figure 6b). Meanwhile, Jia et al. [70] implanted the Zn-0.8 wt.%Mn alloy into the rat femoral condyle for repairing bone defects with pure Ti as control. Their results showed that the new bone tissue at the bone defect site in both groups gradually increased with time, but a large amount of new bone tissue was observed around the Zn-0.8Mn alloy scaffold (Figure 6c). More importantly, in a heavy load-bearing rabbit shaft fracture model, the Zn-0.4Li-based bone plates and screws showed comparable performance in bone fracture fixation compared to the Ti-6Al-4 V counterpart whereas the cortical bone in the Zn-0.4Li alloy group was much thicker (Figure 6d). The results suggest the great potential of Zn-Li based alloys for degradable biomaterials in heavy load-bearing applications [25].
It can be seen from those in vivo studies above (Table 4) that the Zn-based biodegradable materials play an important role in promoting osteogenesis. The corrosion rate of Zn-based biodegradable materials is relatively slow in vivo and can provide a long-term mechanical support in the period of fracture healing [24,25,36]. No incomplete fracture healing and structural collapse of the implant were reported during the animal experiments on load-bearing parts such as the femur [25]. However, the long-term results of the Zn-based implants remain unknown since those animal studies generally lasted for 8–24 weeks [24,25,46]. Additionally, the in vivo studies testing performances of Zn-based implants in fracture fixation are limited [24].
Currently, only small animal models, such as mice, rats and rabbits have been used to examine primarily the biodegradability and biocompatibility of Zn2+ metals on bone defect sites (Table 4) [24,25,33,36,46,51,55,56,70,97]. Although mammals have many similarities, differences across small animals, large animals and humans should be recognized [98]. For example, the difference in skeletal size across various species affects the amount of Zn materials that needs to be degraded or absorbed as well as the mechanical environment, which may lead to varied results between preclinical studies and clinical applications. Therefore, with clinical translations in mind, future studies may be warranted with large animal models. In addition, as Zn-based implants are expected to be used at heavy load-bearing sites for internal fracture fixation, proper site-specific in vivo animal models should be used to test their biodegradability, mechanical properties and biocompatibility (Figure 7).
Table 4. Relevant animal studies of Zn-based biodegradable materials as potential orthopaedic implants.
Table 4. Relevant animal studies of Zn-based biodegradable materials as potential orthopaedic implants.
Zn-Based MetalsDesigned ImplantsControlSurgeriesAnimal SpeciesMajor FindingsRef.
Zn-MnScaffoldPure TiInsertion into
femoral condyle
RatsThe new bone tissues at the bone defect sites gradually increased with time in both groups, and numerous new bone tissues were observed around the Zn-0.8Mn alloy scaffold[70]
Zn-1Mg, Zn-1Ca, Zn-1SrIntramedullary
nails
NAInsertion into
femoral marrow
medullary cavity
MiceThere was no inflammation observed around the implantation site and no mouse died after operation. The new bone thickness of Zn-1Mg, Zn-1Ca and Zn-1Sr pin groups are significantly larger than the sham control group.[51]
Zn-HAPinPure ZnInsertion into
femoral condyle
RatsThere was new bone formation around the Zn-HA composite, and the bone mass increased over time. With prolonged implantation time, the Zn-HA composite was more effective than pure Zn in promoting new bone formation.[55]
Zn-0.05MgPinPure ZnInsertion into
femoral condyle
RabbitsNo inflammatory cells were found at the fracture site, and new bone tissue formation was confirmed at the bone/implant interface, proving that the Zn-0.05Mg alloy promoted the formation of new bone tissue.[46]
Zn-(0.001% Mg 2.5%,
0.01% Fe 2.5%)
Screw and platePLLA, Ti-based alloysMandible fractureBeaglesThe new bone formation in the Zn alloy group and the titanium alloy group was significantly higher than that in the PLLA group. In addition, the new bone formation in the Zn-based alloys group was slightly higher than that in the Ti-based alloys group. The degradation of Zn implants in vivo would not increase the concentration of Zn2+.[97]
Zn-X (Fe, Cu, Ag, Mg, Ca, Sr, Mn, Li)Intramedullary
nails
Pure ZnInsertion into
femoral marrow
medullary cavity
RatsPure Zn, Zn-0.4Fe, Zn-0.4Cu and Zn-2.0Ag alloy implants showed localized degradation patterns with local accumulation of products. In contrast, the degradation of Zn-0.8Mg, Zn-0.8Ca, Zn-0.1Sr, Zn-0.4Li and Zn-0.1Mn was more uniform on the macroscopic scale.[24]
Zn-0.8SrScaffoldPure TiInsertion into
femoral condyle
RatsZn-based alloys promote bone regeneration by promoting the proliferation and differentiation of MC3T3-E1 cells, upregulating the expression of osteogenesis-related genes and proteins, and stimulating angiogenesis.[36]
Zn-0.8Li-0.1CaScaffoldPure TiInsertion into radial defectRabbitsThe Zn-0.8Li-0.1Ca alloy has a similar level of biocompatibility to pure titanium, but it promotes regeneration significantly faster than pure Ti.[33]
Zn-0.4LiScrew and plateTi-6Al-4VFemoral shaft fractureRabbitsPlates and screws made of Zn-0.4Li alloy showed comparable performance to Ti-6Al-4V in fracture fixation, and the fractured bone healed completely six months after surgery.[25]
Zn-1Mg-nvol%β-TCP (n = 0, 1)Columnar samplesZn-1MgSpecimens in lateral thighs.RatsZn-1Mg alloy and Zn-1Mg-β-TCP composites had no significant tissue inflammation and showed good biocompatibility.[56]

6. Summary and Future Directions

A growing number of new Zn-based biodegradable materials have been developed and their biodegradability, mechanical properties, and biocompatibility were tested mostly in vitro and partially in vivo. An ideal biodegradable material for orthopaedic internal fixation should have a suitable combination of biocompatibility, biodegradability, and mechanical properties (YS, UTS, and ε). Although the mechanical properties of pure Zn are difficult to meet the requirements of orthopaedic fixation, Zn-based alloys can achieve the mechanical properties of traditional implants used in internal fixation at load-bearing sites. Zn-based materials have a moderate corrosion rate and good biocompatibility. Their degradation by-product Zn2+ can promote bone growth and mineralization. These properties support Zn-based biodegradable materials as an alternative for internal fixation implants at heavy load-bearing skeletal sites. However, many questions still need to be addressed before Zn-based biodegradable materials can be used for fracture fixation in clinics (Figure 8).
In terms of biodegradability, (1) the current degradation rate of Zn-based biodegradable materials remains relatively slow, and it needs to be further tuned according to the target skeletal site to match its healing rate; (2) due to existence of static and dynamic loads on the skeleton, the stress corrosion and fatigue corrosion of the materials need to be better understood. In terms of the mechanical properties, (1) the elastic modulus (94–110 GPa) of current Zn-based biodegradable materials is higher than that of bone and should be reduced to avoid stress shielding when being used in internal fixation; (2) the creep effect of Zn-based alloys on the failure of the internal fixation implant at a physiological temperature of the human body should be explored further. In terms of biocompatibility, (1) since high content of Zn2+ has toxic effects on cells, attempts should be made to regulate the degradation rate of internal fixation to ensure that the concentration of degradation products does not exceed the safe concentration range of the implant site; (2) antibacterial properties may be explored further. Additionally, in vivo experiments should move from small animal models to large animal models for heavy load-bearing fracture fixation.

Author Contributions

Conceptualization, Y.L. and H.Y.; validation, T.D., A.Q., Y.M. and H.Y.; formal analysis, Y.M.; investigation, Y.L.; writing—original draft preparation, Y.L.; writing—review and editing, H.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China, grant numbers 11702008 and 12272017; Beijing Natural Science Foundation, grant number 7202003; Beijing Municipal Education Commission Research Program, grant number KM202010005035; China Postdoctoral Science Foundation funded project, grant number 2021TQ0020; Cultivation fund of Faculty of Enviroment and Life.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Shuai, C.; Li, S.; Peng, S.; Feng, P.; Lai, Y.; Gao, C. Biodegradable metallic bone implants. Mater. Chem. Front. 2019, 3, 544–562. [Google Scholar] [CrossRef]
  2. Fu, R.; Feng, Y.; Liu, Y.; Willie, B.M.; Yang, H. The combined effects of dynamization time and degree on bone healing. J. Orthop. Res. 2022, 40, 634–643. [Google Scholar] [CrossRef] [PubMed]
  3. Fu, R.; Feng, Y.; Bertrand, D.; Du, T.; Liu, Y.; Willie, B.M.; Yang, H. Enhancing the efficiency of distraction osteogenesis through rate-varying distraction: A computational study. Int. J. Mol. Sci. 2021, 22, 11734. [Google Scholar] [CrossRef] [PubMed]
  4. Tucker, A. Management of common upper limb fractures in adults. Surgery 2022, 40, 184–191. [Google Scholar] [CrossRef]
  5. Wu, A.M.; Bisignano, C.; James, S.L.; Abady, G.G.; Abedi, A.; Abu-Gharbieh, E.; Alhassan, R.K.; Alipour, V.; Arabloo, J.; Asaad, M.; et al. Global, regional, and national burden of bone fractures in 204 countries and territories, 1990–2019: A systematic analysis from the Global Burden of Disease Study 2019. Lancet Healthy Longev. 2021, 2, e580–e592. [Google Scholar] [CrossRef]
  6. Fu, R.; Feng, Y.; Liu, Y.; Yang, H. Mechanical regulation of bone regeneration during distraction osteogenesis. Med. Nov. Technol. Devices 2021, 11, 100077. [Google Scholar] [CrossRef]
  7. Fu, R.; Bertrand, D.; Wang, J.; Kavaseri, K.; Feng, Y.; Du, T.; Liu, Y.; Willie, B.M.; Yang, H. In vivo and in silico monitoring bone regeneration during distraction osteogenesis of the mouse femur. Comput. Methods Programs Biomed. 2022, 216, 106679. [Google Scholar] [CrossRef]
  8. Stiffler, K.S. Internal fracture fixation. Clin. Tech. Small Anim. Pract. 2004, 19, 105–113. [Google Scholar] [CrossRef]
  9. Taljanovic, M.S.; Jones, M.D.; Ruth, J.T.; Benjamin, J.B.; Sheppard, J.E.; Hunter, T.B. Fracture Fixation. Radiographics 2003, 23, 1569–1590. [Google Scholar] [CrossRef]
  10. Manam, N.S.; Harun, W.S.W.; Shri, D.N.A.; Ghani, S.A.C.; Kurniawan, T.; Ismail, M.H.; Ibrahim, M.H.I. Study of corrosion in biocompatible metals for implants: A review. J. Alloys Compd. 2017, 701, 698–715. [Google Scholar] [CrossRef] [Green Version]
  11. Hasiak, M.; Sobieszczanska, B.; Laszcz, A.; Bialy, M.; Checmanowski, J.; Zatonski, T.; Bozemska, E.; Wawrzynska, M. Production, mechanical properties and biomedical characterization of ZrTi-based bulk metallic glasses in comparison with 316L stainless steel and Ti6Al4V alloy. Materials 2021, 15, 252. [Google Scholar] [CrossRef] [PubMed]
  12. Iatecola, A.; Longhitano, G.A.; Antunes, L.H.M.; Jardini, A.L.; Miguel, E.C.; Beres, M.; Lambert, C.S.; Andrade, T.N.; Buchaim, R.L.; Buchaim, D.V.; et al. Osseointegration improvement of Co-Cr-Mo alloy produced by additive manufacturing. Pharmaceutics 2021, 13, 724. [Google Scholar] [CrossRef] [PubMed]
  13. Okazaki, Y.; Katsuda, S.I. Biological safety evaluation and surface modification of biocompatible Ti-15Zr-4Nb alloy. Materials 2021, 14, 731. [Google Scholar] [CrossRef] [PubMed]
  14. Kabir, H.; Munir, K.; Wen, C.; Li, Y. Recent research and progress of biodegradable zinc alloys and composites for biomedical applications: Biomechanical and biocorrosion perspectives. Bioact. Mater. 2021, 6, 836–879. [Google Scholar] [CrossRef]
  15. Wang, J.L.; Xu, J.K.; Hopkins, C.; Chow, D.H.; Qin, L. Biodegradable magnesium-based implants in orthopedics-A general review and perspectives. Adv. Sci. 2020, 7, 1902443. [Google Scholar] [CrossRef]
  16. Witte, F.; Hort, N.; Vogt, C.; Cohen, S.; Kainer, K.U.; Willumeit, R.; Feyerabend, F. Degradable biomaterials based on magnesium corrosion. Curr. Opin. Solid State Mater. Sci. 2008, 12, 63–72. [Google Scholar] [CrossRef]
  17. Jang, H.Y.; Shin, J.Y.; Oh, S.H.; Byun, J.H.; Lee, J.H. PCL/HA hybrid microspheres for effective osteogenic differentiation and bone regeneration. ACS Biomater. Sci. Eng. 2020, 6, 5172–5180. [Google Scholar] [CrossRef]
  18. Yang, Y.; Zhao, Y.; Tang, G.; Li, H.; Yuan, X.; Fan, Y. In vitro degradation of porous poly(l-lactide-co-glycolide)/β-tricalcium phosphate (PLGA/β-TCP) scaffolds under dynamic and static conditions. Polym. Degrad. Stab. 2008, 93, 1838–1845. [Google Scholar] [CrossRef]
  19. Zhou, H.; Lawrence, J.G.; Bhaduri, S.B. Fabrication aspects of PLA-CaP/PLGA-CaP composites for orthopedic applications: A review. Acta Biomater. 2012, 8, 1999–2016. [Google Scholar] [CrossRef]
  20. Zhang, Y.; Xu, J.; Ruan, Y.C.; Yu, M.K.; O’Laughlin, M.; Wise, H.; Chen, D.; Tian, L.; Shi, D.; Wang, J.; et al. Implant-derived magnesium induces local neuronal production of CGRP to improve bone-fracture healing in rats. Nat. Med. 2016, 22, 1160–1169. [Google Scholar] [CrossRef] [Green Version]
  21. Lee, J.W.; Han, H.S.; Han, K.J.; Park, J.; Jeon, H.; Ok, M.R.; Seok, H.K.; Ahn, J.P.; Lee, K.E.; Lee, D.H.; et al. Long-term clinical study and multiscale analysis of in vivo biodegradation mechanism of Mg alloy. Proc. Natl. Acad. Sci. USA 2016, 113, 716–721. [Google Scholar] [CrossRef] [PubMed]
  22. Zheng, Y.F.; Gu, X.N.; Witte, F. Biodegradable metals. Mater. Sci. Eng. R Rep. 2014, 77, 1–34. [Google Scholar] [CrossRef]
  23. Kumar, K.; Gill, R.S.; Batra, U. Challenges and opportunities for biodegradable magnesium alloy implants. Mater. Technol. 2017, 33, 153–172. [Google Scholar] [CrossRef]
  24. 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, 401. [Google Scholar] [CrossRef] [PubMed]
  25. Yang, H.; Qu, X.; Wang, M.; Cheng, H.; Jia, B.; Nie, J.; Dai, K.; Zheng, Y. Zn-0.4Li alloy shows great potential for the fixation and healing of bone fractures at load-bearing sites. Chem. Eng. J. 2021, 417, 129317. [Google Scholar] [CrossRef]
  26. 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] [PubMed]
  27. Li, H.F.; Shi, Z.Z.; Wang, L.N. Opportunities and challenges of biodegradable Zn-based alloys. J. Mater. Sci. Technol. 2020, 46, 136–138. [Google Scholar] [CrossRef]
  28. Torne, K.; Larsson, M.; Norlin, A.; Weissenrieder, J. Degradation of Zinc in saline solutions, plasma, and whole blood. J. Biomed. Mater. Res. B Appl. Biomater. 2016, 104, 1141–1151. [Google Scholar] [CrossRef]
  29. Andreini, C.; Banci, L.; Bertini, I.; Rosato, A. Counting the Zinc-proteins encoded in the human genome. J. Proteome Res. 2006, 5, 196–201. [Google Scholar] [CrossRef]
  30. Weiss, A.; Murdoch, C.C.; Edmonds, K.A.; Jordan, M.R.; Monteith, A.J.; Perera, Y.R.; Rodriguez Nassif, A.M.; Petoletti, A.M.; Beavers, W.N.; Munneke, M.J.; et al. Zn-regulated GTPase metalloprotein activator 1 modulates vertebrate zinc homeostasis. Cell 2022, 185, 2148–2163. [Google Scholar] [CrossRef]
  31. Qiao, W.; Pan, D.; Zheng, Y.; Wu, S.; Liu, X.; Chen, Z.; Wan, M.; Feng, S.; Cheung, K.M.C.; Yeung, K.W.K.; et al. Divalent metal cations stimulate skeleton interoception for new bone formation in mouse injury models. Nat. Commun. 2022, 13, 535. [Google Scholar] [CrossRef]
  32. 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]
  33. Zhang, Z.; Jia, B.; Yang, H.; Han, Y.; Wu, Q.; Dai, K.; Zheng, Y. Biodegradable ZnLiCa ternary alloys for critical-sized bone defect regeneration at load-bearing sites: In vitro and in vivo studies. Bioact. Mater. 2021, 6, 3999–4013. [Google Scholar] [CrossRef]
  34. Bowen, P.K.; Seitz, J.M.; Guillory, R.J.; 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]
  35. Dambatta, M.S.; Izman, S.; Kurniawan, D.; Hermawan, H. Processing of Zn-3Mg alloy by equal channel angular pressing for biodegradable metal implants. J. King Saud. Univ. Sci. 2017, 29, 455–461. [Google Scholar] [CrossRef]
  36. Jia, B.; Yang, H.; Zhang, Z.; Qu, X.; Jia, X.; Wu, Q.; Han, Y.; Zheng, Y.; Dai, K. Biodegradable Zn-Sr alloy for bone regeneration in rat femoral condyle defect model: In vitro and in vivo studies. Bioact. Mater. 2021, 6, 1588–1604. [Google Scholar] [CrossRef]
  37. Kafri, A.; Ovadia, S.; Goldman, J.; Drelich, J.; Aghion, E. The suitability of Zn–1.3%Fe alloy as a biodegradable implant material. Metals 2018, 8, 153. [Google Scholar] [CrossRef]
  38. Krezel, A.; Maret, W. The biological inorganic chemistry of zinc ions. Arch. Biochem. Biophys. 2016, 611, 3–19. [Google Scholar] [CrossRef] [PubMed]
  39. Ren, T.; Gao, X.; Xu, C.; Yang, L.; Guo, P.; Liu, H.; Chen, Y.; Sun, W.; Song, Z. Evaluation of as-extruded ternary Zn–Mg–Zr alloys for biomedical implantation material: In vitro and in vivo behavior. Mater. Corros. 2019, 70, 1056–1070. [Google Scholar] [CrossRef]
  40. Shi, Z.Z.; Yu, J.; Liu, X.F.; Zhang, H.J.; Zhang, D.W.; Yin, Y.X.; Wang, L.N. Effects of Ag, Cu or Ca addition on microstructure and comprehensive properties of biodegradable Zn-0.8Mn alloy. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 99, 969–978. [Google Scholar] [CrossRef]
  41. Shi, Z.; Yu, J.; Liu, X. Microalloyed Zn-Mn alloys: From extremely brittle to extraordinarily ductile at room temperature. Mater. Des. 2018, 144, 343–352. [Google Scholar] [CrossRef]
  42. Shuai, C.; Cheng, Y.; Yang, Y.; Peng, S.; Yang, W.; Qi, F. Laser additive manufacturing of Zn-2Al part for bone repair: Formability; microstructure and properties. J. Alloys Compd. 2019, 798, 606–615. [Google Scholar] [CrossRef]
  43. Shuai, C.; Xue, L.; Gao, C.; Peng, S.; Zhao, Z. Rod-like eutectic structure in biodegradable Zn-Al-Sn alloy exhibiting enhanced mechanical strength. ACS Biomater. Sci. Eng. 2020, 6, 3821–3831. [Google Scholar] [CrossRef]
  44. Shuai, C.; Xue, L.; Gao, C.; Yang, Y.; Peng, S.; Zhang, Y. Selective laser melting of Zn–Ag alloys for bone repair: Microstructure; mechanical properties and degradation behavior. Virtual Phys. Prototyp. 2018, 13, 146–154. [Google Scholar] [CrossRef]
  45. Shi, Z.; Yu, J.; Liu, X.; Wang, L. Fabrication and characterization of novel biodegradable Zn-Mn-Cu alloys. J. Mater. Sci. Technol. 2018, 34, 1008–1015. [Google Scholar] [CrossRef]
  46. Xiao, C.; Wang, L.; Ren, Y.; Sun, S.; Zhang, E.; Yan, C.; Liu, Q.; Sun, X.; Shou, F.; Duan, J.; et al. Indirectly extruded biodegradable Zn-0.05wt%Mg alloy with improved strength and ductility: In vitro and in vivo studies. J. Mater. Sci. Technol. 2018, 34, 1618–1627. [Google Scholar] [CrossRef]
  47. Yuan, W.; Xia, D.; Wu, S.; Zheng, Y.; Guan, Z.; Rau, J.V. A review on current research status of the surface modification of Zn-based biodegradable metals. Bioact. Mater. 2022, 7, 192–216. [Google Scholar] [CrossRef] [PubMed]
  48. 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]
  49. Yang, J.; Yim, C.D.; You, B.S. Effects of solute Zn on corrosion film of Mg–Sn–Zn alloy formed in NaCl solution. J. Electrochem. Soc. 2016, 163, C839–C844. [Google Scholar] [CrossRef]
  50. 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]
  51. 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]
  52. Tong, X.; Zhang, D.; Zhang, X.; Su, Y.; Shi, Z.; Wang, K.; Lin, J.; Li, Y.; Lin, J.; Wen, C. Microstructure; mechanical properties; biocompatibility; and in vitro corrosion and degradation behavior of a new Zn-5Ge alloy for biodegradable implant materials. Acta Biomater. 2018, 82, 197–204. [Google Scholar] [CrossRef] [PubMed]
  53. Li, P.; Zhang, W.; Dai, J.; Xepapadeas, A.B.; Schweizer, E.; Alexander, D.; Scheideler, L.; Zhou, C.; Zhang, H.; Wan, G.; et al. Investigation of zinccopper alloys as potential materials for craniomaxillofacial osteosynthesis implants. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 103, 109826. [Google Scholar] [CrossRef]
  54. Wątroba, M.; Bednarczyk, W.; Kawałko, J.; Bała, P. Effect of zirconium microaddition on the microstructure and mechanical properties of Zn-Zr alloys. Mater. Charact. 2018, 142, 187–194. [Google Scholar] [CrossRef]
  55. Yang, H.; Qu, X.; Lin, W.; Wang, C.; Zhu, D.; Dai, K.; Zheng, Y. In vitro and in vivo studies on zinc-hydroxyapatite composites as novel biodegradable metal matrix composite for orthopedic applications. Acta Biomater. 2018, 71, 200–214. [Google Scholar] [CrossRef]
  56. Pan, C.; Sun, X.; Xu, G.; Su, Y.; Liu, D. The effects of beta-TCP on mechanical properties; corrosion behavior and biocompatibility of beta-TCP/Zn-Mg composites. Mater. Sci. Eng. C Mater. Biol. Appl. 2020, 108, 110397. [Google Scholar] [CrossRef]
  57. Li, X.; Chu, C.; Chu, P.K. Effects of external stress on biodegradable orthopedic materials: A review. Bioact. Mater. 2016, 1, 77–84. [Google Scholar] [CrossRef]
  58. Li, N.; Zheng, Y. Novel magnesium alloys developed for biomedical application: A review. J. Mater. Sci. Technol. 2013, 29, 489–502. [Google Scholar] [CrossRef]
  59. Kirkland, N.T.; Birbilis, N.; Staiger, M.P. Assessing the corrosion of biodegradable magnesium implants: A critical review of current methodologies and their limitations. Acta Biomater. 2012, 8, 925–936. [Google Scholar] [CrossRef]
  60. Li, G.N.; Zhu, S.M.; Nie, J.F.; Zheng, Y.; Sun, Z. Investigating the stress corrosion cracking of a biodegradable Zn-0.8 wt%Li alloy in simulated body fluid. Bioact. Mater. 2021, 6, 1468–1478. [Google Scholar] [CrossRef]
  61. Li, P.; Feng, X.; Jia, X.; Fan, Y. Influences of tensile load on in vitro degradation of an electrospun poly(L-lactide-co-glycolide) scaffold. Acta Biomater. 2010, 6, 2991–2996. [Google Scholar] [CrossRef] [PubMed]
  62. Guo, M.; Chu, Z.; Yao, J.; Feng, W.; Wang, Y.; Wang, L.; Fan, Y. The effects of tensile stress on degradation of biodegradable PLGA membranes: A quantitative study. Polym. Degrad. Stab. 2016, 124, 95–100. [Google Scholar] [CrossRef]
  63. 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. 2019, 83, 477–486. [Google Scholar] [CrossRef] [PubMed]
  64. Gao, Y.; Wang, L.; Gu, X.; Chu, Z.; Guo, M.; Fan, Y. A quantitative study on magnesium alloy stent biodegradation. J. Biomech. 2018, 74, 98–105. [Google Scholar] [CrossRef]
  65. Vasudevan, A.K.; Sadananda, K. Classification of environmentally assisted fatigue crack growth behavior. Int. J. Fatigue 2009, 31, 1696–1708. [Google Scholar] [CrossRef]
  66. Jafari, S.; Singh Raman, R.K.; Davies, C.H.J. Corrosion fatigue of a magnesium alloy in modified simulated body fluid. Eng. Fract. Mech. 2015, 137, 2–11. [Google Scholar] [CrossRef]
  67. Zhao, J.; Gao, L.L.; Gao, H.; Yuan, X.; Chen, X. Biodegradable behaviour and fatigue life of ZEK100 magnesium alloy in simulated physiological environment. Fatigue Fract. Eng. Mater. Struct. 2015, 38, 904–913. [Google Scholar] [CrossRef]
  68. Shi, Z.; Li, H.; Xu, J.; Gao, X.; Liu, X. Microstructure evolution of a high-strength low-alloy Zn–Mn–Ca alloy through casting; hot extrusion and warm caliber rolling. Mater. Sci. Eng. A 2020, 771, 138626. [Google Scholar] [CrossRef]
  69. Guan, Z.; Linsley, C.S.; Pan, S.; Yao, G.; Wu, B.M.; Levi, D.S.; Li, X. Zn-Mg-WC nanocomposites for bioresorbable cardiovascular stents: Microstructure, mechanical properties, fatigue, shelf life, and corrosion. ACS Biomater. Sci. Eng. 2022, 8, 328–339. [Google Scholar] [CrossRef]
  70. Jia, B.; Yang, H.; Han, Y.; Zhang, Z.; Qu, X.; Zhuang, Y.; Wu, Q.; Zheng, Y.; Dai, K. In vitro and in vivo studies of Zn-Mn biodegradable metals designed for orthopedic applications. Acta Biomater. 2020, 108, 358–372. [Google Scholar] [CrossRef]
  71. Katarivas Levy, G.; Leon, A.; Kafri, A.; Ventura, Y.; Drelich, J.W.; Goldman, J.; Vago, R.; Aghion, E. Evaluation of biodegradable Zn-1%Mg and Zn-1%Mg-0.5%Ca alloys for biomedical applications. J. Mater. Sci. Mater. Med. 2017, 28, 174. [Google Scholar] [CrossRef] [PubMed]
  72. Liu, C.; Li, Y.; Ge, Q.; Liu, Z.; Qiao, A.; Mu, Y. Mechanical characteristics and in vitro degradation of biodegradable Zn-Al alloy. Mater. Lett. 2021, 300, 130181. [Google Scholar] [CrossRef]
  73. Tang, Z.; Niu, J.; Huang, H.; Zhang, H.; Pei, J.; Ou, J.; Yuan, G. Potential biodegradable Zn-Cu binary alloys developed for cardiovascular implant applications. J. Mech. Behav. Biomed. Mater. 2017, 72, 182–191. [Google Scholar] [CrossRef]
  74. Lin, J.; Tong, X.; Shi, Z.; Zhang, D.; Zhang, L.; Wang, K.; Wei, A.; Jin, L.; Lin, J.; Li, Y.; et al. A biodegradable Zn-1Cu-0.1Ti alloy with antibacterial properties for orthopedic applications. Acta Biomater. 2020, 106, 410–427. [Google Scholar] [PubMed]
  75. Zhu, S.; Wu, C.; Li, G.; Zheng, Y.; Nie, J. Creep properties of biodegradable Zn-0.1Li alloy at human body temperature: Implications for its durability as stents. Mater. Res. Lett. 2019, 7, 347–353. [Google Scholar] [CrossRef]
  76. Yuan, W.; Xia, D.; Zheng, Y.; Liu, X.; Wu, S.; Li, B.; Han, Y.; Jia, Z.; Zhu, D.; Ruan, L.; et al. Controllable biodegradation and enhanced osseointegration of ZrO2-nanofilm coated Zn-Li alloy: In vitro and in vivo studies. Acta Biomater. 2020, 105, 290–303. [Google Scholar] [CrossRef]
  77. Li, H.; Yang, H.; Zheng, Y.; Zhou, F.; Qiu, K.; Wang, X. Design and characterizations of novel biodegradable ternary Zn-based alloys with IIA nutrient alloying elements Mg, Ca and Sr. Mater. Des. 2015, 83, 95–102. [Google Scholar] [CrossRef]
  78. Lin, J.; Tong, X.; Sun, Q.; Luan, Y.; Zhang, D.; Shi, Z.; Wang, K.; Lin, J.; Li, Y.; Dargusch, M.; et al. Biodegradable ternary Zn-3Ge-0.5X (X = Cu; Mg; and Fe) alloys for orthopedic applications. Acta Biomater. 2020, 115, 432–446. [Google Scholar] [CrossRef]
  79. Mostaed, E.; Sikora-Jasinska, M.; Ardakani, M.S.; Mostaed, A.; Reaney, I.M.; Goldman, J.; Drelich, J.W. Towards revealing key factors in mechanical instability of bioabsorbable Zn-based alloys for intended vascular stenting. Acta Biomater. 2020, 105, 319–335. [Google Scholar] [CrossRef]
  80. Xue, P.; Ma, M.; Li, Y.; Li, X.; Yuan, J.; Shi, G.; Wang, K.; Zhang, K. Microstructure, mechanical properties, and in vitro corrosion behavior of biodegradable Zn-1Fe-xMg alloy. Materials 2020, 13, 4835. [Google Scholar] [CrossRef]
  81. Bowen, P.K.; Shearier, E.R.; Zhao, S.; Guillory II, R.J.; Zhao, F.; Goldman, J.; Drelich, J.W. Biodegradable metals for cardiovascular stents: From clinical concerns to recent Zn-alloys. Adv. Healthc. Mater. 2016, 5, 1121–1140. [Google Scholar] [CrossRef]
  82. Plum, L.M.; Rink, L.; Haase, H. The essential toxin: Impact of zinc on human health. Int. J. Environ. Res. Publ. Health. 2010, 7, 1342–1365. [Google Scholar] [CrossRef] [PubMed]
  83. O’Connor, J.P.; Kanjilal, D.; Teitelbaum, M.; Lin, S.S.; Cottrell, J.A. Zinc as a therapeutic agent in bone regeneration. Materials 2020, 13, 2211. [Google Scholar] [CrossRef] [PubMed]
  84. Tapiero, H.; Tew, K.D. Trace elements in human physiology and pathology: Zinc and metallothioneins. Biomed. Pharmacother. 2003, 57, 399–411. [Google Scholar] [CrossRef]
  85. Glutsch, V.; Hamm, H.; Goebeler, M. Zinc and skin: An update. J. Dtsch. Dermatol. Ges. 2019, 17, 589–596. [Google Scholar] [CrossRef]
  86. Jimenez, M.; Abradelo, C.; San Roman, J.; Rojo, L. Bibliographic review on the state of the art of strontium and zinc based regenerative therapies. Recent developments and clinical applications. J. Mater. Chem. B 2019, 7, 1974–1985. [Google Scholar] [CrossRef] [PubMed]
  87. Hernandez-Escobar, D.; Champagne, S.; Yilmazer, H.; Dikici, B.; Boehlert, C.J.; Hermawan, H. Current status and perspectives of zinc-based absorbable alloys for biomedical applications. Acta Biomater. 2019, 97, 1–22. [Google Scholar] [CrossRef]
  88. Venezuela, J.; Dargusch, M.S. 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] [PubMed]
  89. Bakhsheshi-Rad, H.R.; Hamzah, E.; Low, H.T.; Kasiri-Asgarani, M.; Farahany, S.; Akbari, E.; Cho, M.H. Fabrication of biodegradable Zn-Al-Mg alloy: Mechanical properties, corrosion behavior, cytotoxicity and antibacterial activities. Mater. Sci. Eng. C Mater. Biol. Appl. 2017, 73, 215–219. [Google Scholar] [CrossRef] [PubMed]
  90. Sukhodub, L.B. Antimicrobial activity of Ag+, Cu2+, Zn2+, Mg2+ ions doped chitosan nanoparticles. Ann. Mechnikov’s Inst. 2015, 1, 39–43. [Google Scholar]
  91. Serrano-Aroca, A.; Cano-Vicent, A.; Sabater, I.S.R.; El-Tanani, M.; Aljabali, A.; Tambuwala, M.M.; Mishra, Y.K. Scaffolds in the microbial resistant era: Fabrication, materials, properties and tissue engineering applications. Mater Today Bio 2022, 16, 100412. [Google Scholar] [CrossRef]
  92. Riduan, S.N.; Zhang, Y. Recent Advances of Zinc-based Antimicrobial Materials. Chem. Asian J. 2021, 16, 2588–2595. [Google Scholar] [CrossRef]
  93. de Lima, C.O.; de Oliveira, A.L.M.; Chantelle, L.; Silva Filho, E.C.; Jaber, M.; Fonseca, M.G. Zn-doped mesoporous hydroxyapatites and their antimicrobial properties. Colloids Surf. 2021, 198, 111471. [Google Scholar] [CrossRef] [PubMed]
  94. Tong, X.; Shi, Z.; Xu, L.; Lin, J.; Zhang, D.; Wang, K.; Li, Y.; Wen, C. Degradation behavior, cytotoxicity, hemolysis, and antibacterial properties of electro-deposited Zn-Cu metal foams as potential biodegradable bone implants. Acta Biomater. 2020, 102, 481–492. [Google Scholar] [CrossRef] [PubMed]
  95. Ren, L.; Yang, K. Antibacterial design for metal implants. In Metallic Foam Bone; Woodhead Publishing: Sawston, UK, 2017; pp. 203–216. [Google Scholar]
  96. Cui, Z.; Zhang, Y.; Cheng, Y.; Gong, D.; Wang, W. Microstructure; mechanical, corrosion properties and cytotoxicity of betacalcium polyphosphate reinforced ZK61 magnesium alloy composite by spark plasma sintering. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 99, 1035–1047. [Google Scholar] [CrossRef] [PubMed]
  97. Wang, X.; Shao, X.; Dai, T.; Xu, F.; Zhou, J.G.; Qu, G.; Tian, L.; Liu, B.; Liu, Y. In vivo study of the efficacy, biosafety, and degradation of a zinc alloy osteosynthesis system. Acta Biomater. 2019, 92, 351–361. [Google Scholar] [CrossRef]
  98. Taguchi, T.; Lopez, M.J. An overview of de novo bone generation in animal models. J. Orthop. Res. 2021, 39, 7–21. [Google Scholar] [CrossRef]
Figure 1. Mechanical properties of biodegradable and non-biodegradable materials for orthopaedic devices and their clinical applications [24].
Figure 1. Mechanical properties of biodegradable and non-biodegradable materials for orthopaedic devices and their clinical applications [24].
Jfb 13 00164 g001
Figure 2. A schematic indicating the state of the art of research on Zn-based biodegradable materials for orthopaedic internal fixation.
Figure 2. A schematic indicating the state of the art of research on Zn-based biodegradable materials for orthopaedic internal fixation.
Jfb 13 00164 g002
Figure 3. Schematics illustrating the processes of bone healing and implant degradation under a perfect matching scene.
Figure 3. Schematics illustrating the processes of bone healing and implant degradation under a perfect matching scene.
Jfb 13 00164 g003
Figure 4. (a) Biological roles of Zn. Reprinted with permission from Ref. [81], Copyright 2016 Wiley. (b) Comparison of the influence of zinc excess versus deficiency [82].
Figure 4. (a) Biological roles of Zn. Reprinted with permission from Ref. [81], Copyright 2016 Wiley. (b) Comparison of the influence of zinc excess versus deficiency [82].
Jfb 13 00164 g004
Figure 5. Zn2+ deficiency has been linked to impaired skeletal development and bone growth in humans and animals [83].
Figure 5. Zn2+ deficiency has been linked to impaired skeletal development and bone growth in humans and animals [83].
Jfb 13 00164 g005
Figure 6. (a) Hard tissue sections of pure Zn, Zn-0.4Li, Zn-0.1Mn, Zn-0.8 Mg, Zn-0.8Ca, Zn-0.1Sr, Zn-0.4Fe, Zn-0.4Cu and Zn-2Ag in metaphysis. The magnified region is marked by red rectangle. NB, new bone; DP, degradation products; FT, fibrous tissue. Scale bar, 0.5 mm in low magnification, 500 μm in high magnification [24]. (b) Histological characterization of hard tissue sections at implant sites. Van Gieson staining of pure Zn. Reprinted with permission from Ref. [55], Copyright 2018, Elsevier. (c) The Van Gieson staining results of specimens 4 weeks, 8 weeks, and 12 weeks postoperatively. Within each row, full-view images of bone defect areas (20×), medium magnification images (50×), and higher magnification images (100×) arranged from left to right. Reprinted with permission from Ref. [70], Copyright 2020, Elsevier. (d) Van Gieson staining of representative histological images of femoral fracture healing at 6 months. The fracture healing and fixation screws are magnified and marked by red rectangles. Reprinted with permission from Ref. [25], Copyright 2021, Elsevier.
Figure 6. (a) Hard tissue sections of pure Zn, Zn-0.4Li, Zn-0.1Mn, Zn-0.8 Mg, Zn-0.8Ca, Zn-0.1Sr, Zn-0.4Fe, Zn-0.4Cu and Zn-2Ag in metaphysis. The magnified region is marked by red rectangle. NB, new bone; DP, degradation products; FT, fibrous tissue. Scale bar, 0.5 mm in low magnification, 500 μm in high magnification [24]. (b) Histological characterization of hard tissue sections at implant sites. Van Gieson staining of pure Zn. Reprinted with permission from Ref. [55], Copyright 2018, Elsevier. (c) The Van Gieson staining results of specimens 4 weeks, 8 weeks, and 12 weeks postoperatively. Within each row, full-view images of bone defect areas (20×), medium magnification images (50×), and higher magnification images (100×) arranged from left to right. Reprinted with permission from Ref. [70], Copyright 2020, Elsevier. (d) Van Gieson staining of representative histological images of femoral fracture healing at 6 months. The fracture healing and fixation screws are magnified and marked by red rectangles. Reprinted with permission from Ref. [25], Copyright 2021, Elsevier.
Jfb 13 00164 g006
Figure 7. Schematic representation of common animal bone defect models [98].
Figure 7. Schematic representation of common animal bone defect models [98].
Jfb 13 00164 g007
Figure 8. Future directions of Zn-based biodegradable materials.
Figure 8. Future directions of Zn-based biodegradable materials.
Jfb 13 00164 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, Y.; Du, T.; Qiao, A.; Mu, Y.; Yang, H. Zinc-Based Biodegradable Materials for Orthopaedic Internal Fixation. J. Funct. Biomater. 2022, 13, 164. https://doi.org/10.3390/jfb13040164

AMA Style

Liu Y, Du T, Qiao A, Mu Y, Yang H. Zinc-Based Biodegradable Materials for Orthopaedic Internal Fixation. Journal of Functional Biomaterials. 2022; 13(4):164. https://doi.org/10.3390/jfb13040164

Chicago/Turabian Style

Liu, Yang, Tianming Du, Aike Qiao, Yongliang Mu, and Haisheng Yang. 2022. "Zinc-Based Biodegradable Materials for Orthopaedic Internal Fixation" Journal of Functional Biomaterials 13, no. 4: 164. https://doi.org/10.3390/jfb13040164

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