Skip to main content

The impact of Zn-doped synthetic polymer materials on bone regeneration: a systematic review

Abstract

Introduction

To repair bone defects, a variety of bone substitution materials have been used, such as ceramics, metals, natural and synthetic polymers, and combinations thereof. In recent decades, a wide range of synthetic polymers have been used for bone regeneration. These polymers have the advantages of biocompatibility, biodegradability, good mechanical properties, low toxicity, and ease of processing. However, when used alone, they are unable to achieve ideal bone formation. Incorporating zinc (Zn) into synthetic polymers has been considered, as previous studies have shown that Zn2+ promotes stem cell osteogenesis and mineral deposition. The purpose of this systematic review was to provide an overview of the application and effectiveness of Zn in synthetic polymers for bone regeneration, whether used alone or in combination with other biomaterials. This study was performed according to the PRISMA guidelines.

Materials and methods

A search of the PubMed, Embase, and the Cochrane Library databases for articles published up to June 2020 revealed 153 relevant studies. After screening the titles, abstracts, and full texts, 13 articles were included in the review; 9 of these were in vitro, 3 were in vivo, and 1 included both in vitro and in vivo experiments.

Results

At low concentrations, Zn2+ promoted cell proliferation and osteogenic differentiation, while high-dose Zn2+ resulted in cytotoxicity and inhibition of osteogenic differentiation. Additionally, one study showed that Zn2+ reduced apatite formation in simulated body fluid. In all of the in vivo experiments, Zn-containing materials enhanced bone formation.

Conclusions

At appropriate concentrations, Zn-doped synthetic polymer materials are better able to promote bone regeneration than materials without Zn.

Background

Bone defects caused by various conditions, such as congenital malformation, trauma, tumor resection, and infection, not only cause great pain but also place tremendous pressure on healthcare systems [1]. Bone tissue engineering (BTE) emerging in recent decades now provides an option for the therapy of bone defects. The method involves combining biomaterials with donor cells to promote bone regeneration, and a variety of biomaterials have been used in BTE to repair and replace traumatized/damaged bone tissue, such as ceramics, metals, natural and synthetic polymers, and combinations thereof [2, 3].

As a commonly used material both in conventional processing methods, like freeze-drying and solvent-casting, and emerging three-dimensional (3D) printing technology, a wide range of synthetic polymers have been applied for bone regeneration, including polylactic acid (PLA), polycaprolactone (PCL), poly (glycolic acid) (PGA), and their copolymers, such as poly (lactic-co-glycolic) acid (PLGA). Approved by the Food and Drug Administration, these polymers are considered biocompatible with natural tissues [4] and have several other advantages. For example, as synthetic materials, their composition, architecture, and physical properties are adjustable; also, these materials are highly reproducible. Moreover, these polymers are degradable, so they can provide mechanical support initially, before gradually degrading to make room for newly formed tissue [5, 6]. However, although synthetic polymers can be osteoconductive, they usually do not achieve the desired effect of complete bone regeneration when used alone [4].

Zinc (Zn), an essential trace element, is usually found in skeletal tissue, and approximately 30% of Zn in the body is contained in bone tissue. Zn plays a significant role in the formation, development, mineralization, and maintenance of healthy bones [7]. Also, as a mediator of bone development and growth, Zn deficiency in humans can lead to diseases such as dwarfism, osteoporosis, and stunted bone development. Previous in vitro studies have demonstrated that Zn2+ improves stem cell osteogenesis and enhances mineral deposition [8].

Thus, Zn incorporation is expected to improve the osteogenic ability of synthetic polymer materials, as the release of Zn2+ may enhance the osteogenic differentiation of cells for accelerated bone regeneration [9].

However, there is no consensus on the precise role of Zn in these effects, and different types of cells seem to respond differently depending on the Zn2+ concentration. In a study by Bertels et al., 0.04–0.08 mM of zinc sulfate (ZnSO4) enhanced mineral nodule formation in swine adipose-derived stem cells (swine ASC) in osteogenic media; moreover, the response was biphasic: concentrations of ZnSO4 above 0.08 mM were detrimental to cell growth [10]. And Tiffany et al. found that Zn2+ in amounts below 0.04 mM still enhanced the cell number and metabolic activity of porcine adipose-derived stem cells (pASCs) [11]. In another study, Xiong et al. showed that 10.91–27.15 μM of Zn2+ in cell culture medium significantly enhanced the proliferation and alkaline phosphatase (ALP) activity of mouse bone marrow-derived mesenchymal stem cells (mBMSCs). Similarly, high concentrations (128.58 μM) of Zn have been shown to significantly inhibit ALP activity [12].

In this systematic review, we describe the incorporation of Zn into synthetic polymer materials (whether used alone or in combination with other biomaterials), evaluate the effectiveness of these Zn-doped materials for promoting osteogenesis, and discuss the viable concentration of Zn2+ for osteogenesis in different types of cells.

Methods

Objectives

The objective of this research was to review the literature on Zn-doped synthetic polymer materials, to provide an overview of the application and effectiveness of incorporating Zn into synthetic polymers to improve their osteogenic ability.

Guidelines

The PRISMA guidelines were followed in this research [13].

PICO

P: Synthetic polymer-based materials

I: Incorporation of Zn into synthetic polymers

C: Comparison between synthetic polymers used alone and in combination with Zn

O: Overview of the bone regeneration effect

Inclusion and exclusion criteria

The following are the inclusion criteria:

  • In vitro studies

  • In vivo studies with animal bone defect models

  • Zn-doped synthetic polymer materials used alone or in combination with other biomaterials

  • Materials used in the form of membranes, scaffolds, disks, etc.

  • Published in English

The following are the exclusion criteria:

  • Synthetic polymer materials not used

  • No control group

Search strategy

The PubMed, Embase, and Cochrane Library databases were searched for studies about the application and effectiveness of Zn-doped synthetic polymer materials for promoting osteogenic ability, published at any time up to June 2020. The following keywords were used: “Zinc,” “Polymers,” and “Osteogenesis.” Details of the search strategies are provided in Supplementary Materials Tables S1, S2, and S3.

Study selection and data collection

Studies were selected by two reviewers (Li R, Zhao X), who independently screened the titles, abstracts, and full texts of all retrieved articles. Disagreements were resolved by discussion or through consultation with a third reviewer (Wang S). Potentially suitable articles were then assessed according to the inclusion and exclusion criteria, and data were then extracted. A flow chart of the study selection process is shown in Fig. 1.

Fig. 1
figure 1

Flow chart of the study selection process

Quality assessment

Two other researchers (Zhu Y, Gu R) used the Methodological Index for Non-Randomized Studies (MINORS) scale to evaluate the risk of bias [14, 15]. The revised and validated version of MINORS contains 12 items. Each item is scored from 0 to 2 (0, not reported; 1, inadequately reported; 2, sufficiently reported). The ideal total score is 16 for non-comparative studies and 24 for comparative studies. Disagreements regarding scores were resolved by discussion.

Results

Study selection and characteristics

Of the 153 articles screened, 13 were deemed eligible for the review. In total, 28 articles were excluded due to being duplicates, and a further 55, 38, and 19 studies were excluded after screening the title, abstract, and full text, respectively. The contents of the 13 included articles are systematically summarized in Tables 1 and 2, and the results of the quality assessments are listed in Table 3. All included studies had scores above 16 on the MINORS scale, indicating a low risk of bias. In these studies, synthetic polymers were typically used in conjunction with other biomaterials.

Table 1 Results of in vitro studies
Table 2 Results of in vivo studies
Table 3 MINORS bias scale

Nine of the studies were conducted in vitro, three were in vivo, and one included both in vitro and in vivo experiments. Three of the studies were performed on New Zealand rabbits (tibia or skull) [26,27,28], and one was performed on Sprague-Dawley (SD) rats (femur) [24]. The in vitro studies were concerned with cell proliferation and viability, and osteogenic differentiation, assessed using MTS, MTT, and CCK-8 assays, and ALP activity, RT-PCR, immunofluorescence staining, Alizarin Red S (ARS) staining, and calcium content assay, respectively.

Results of individual studies

To assess the effect of Zn-doped synthetic polymer materials on bone regeneration, proliferation/viability and osteogenic differentiation data were obtained, and in vivo osteogenesis evaluations were performed, as discussed above.

Cell proliferation/viability assessments

Seven of the thirteen studies assessed cellular proliferation and viability using the MTS assay (n = 1) [16], MTT assay (n = 3) [17, 21, 25], or CCK-8 assay (n = 3) [18, 19, 22]. The results mostly showed that the Zn-containing materials did not exhibit cytotoxicity. However, Li et al. demonstrated that the response of bone marrow-derived mesenchymal stem cells (BMSCs) to a Zn-containing coating was highly dose-dependent [22]. Specifically, an appropriate dose (polyacrylic acid [PAA] powder dispersed in 0.025 or 0.05 mol/L zinc chloride solution; PAA-0.025 M and PAA-0.05 M, respectively) promoted cell proliferation, whereas a high dose (PAA powder dispersed in 0.1 mol/L zinc chloride solution; PAA-0.1 M) exhibited cytotoxicity [22].

Cell osteogenic differentiation assessments

Of the 13 articles, 10 assessed osteogenic differentiation based on ALP activity, RT-PCR, immunofluorescence staining, ARS staining, and calcium assay. One study used all of the abovementioned evaluation methods [17], and the others used at least one of the techniques. Overall, Zn-containing materials significantly promoted osteogenic differentiation compared to the control groups without Zn. Amiri et al. fabricated a Zn2SiO4-PES-PEG scaffold and showed the osteogenic differentiation rate of human mesenchymal stem cells (hMSCs) on this scaffold was increased compared to that of PES-PEG fibrous scaffolds [17]. Li et al. demonstrated that, consistent with cell proliferation results, the PAA-0.025M and PAA-0.05M groups exhibited higher ALP activity and more calcium nodules compared to the PAA and PAA-0.1M groups [22]. Interestingly, Rajzer et al. found that PCL-A2Zn5 (PCL combined with 49CaO-5ZnO-6P2O5-40SiO2 bioactive glass [BG]) membranes promoted ALP expression, whereas in simulated body fluid (SBF), apatite formation on the surface of the PCL-A2 (PCL combined with 40SiO2-54CaO-6P2O5) membranes was markedly increased compared to pure PCL and PCL-A2Zn5 membranes [20].

It is worth mentioning that some of the studies added other metal ions to the polymer materials besides Zn2+ and thus achieved better results with respect to promoting bone regeneration. Deng et al. reported that a silver (Ag)/Zn-codecorated SPEEK surface showed enhanced cell proliferation and osteogenic differentiation compared to a Zn-SPEEK surface [19]. Liang et al. compared the efficacy of PLGA/CPC-Zn scaffolds and PLGA/CPC-Si/Zn scaffolds in terms of osteogenesis; their results showed that the expression of bone morphogenetic protein 2 (BMP-2) was higher in the PLGA/CPC-Si/Zn scaffolds [24].

In vivo experiments

In the in vivo experiments, bone defect sites were created in the tibia, femur, and skull, typically in New Zealand rabbits or SD rats (Fig. 2). All of the studies showed that Zn-containing materials can play a role in promoting bone formation, consistent with in vitro studies. It should be noted that Liang et al. demonstrated that PLGA/CPC-Si/Zn scaffolds yielded a substantial increase in the amount of regenerated bone volume over the PLGA/CPC-Zn scaffolds [24].

Fig. 2
figure 2

In vivo experiments with animal bone defect models. a Skull defects in rabbits [27, 28]. b Tibia defects in rabbits [26]. c Femur defects in SD rats [24]

Discussion

The unique advantages of synthetic polymers include biocompatibility, biodegradability, good mechanical properties, and low toxicity; additionally, their shape, porosity, and mechanical properties can be well defined and controlled, and they can be stably mass-produced [29]. However, hydrophilia deficiencies and a lack of bioactivity compromise the ability of synthetic polymers to facilitate biomaterial-host interactions. Furthermore, they typically show poor antibacterial properties, and there is some concern that their degradation products generate an acidic environment conducive to inflammation [5]. Therefore, synthetic polymers are commonly combined with other biomaterials to obtain two-dimensional or 3D scaffolds that promote the desired cell behavior and tissue regeneration.

Some of the 13 studies included in our review support the conclusion that the addition of BGs can address the abovementioned issues. In the study of Rajzer et al., an SBF bioactivity test demonstrated that the presence of BG on PCL membranes induced mineralization, thus indicating its bioactivity [20]. As a solution to the problem of acidic products, Oh et al. showed that the ions dissolved from BGs neutralize the acidic environment caused by polymers, and reported significantly higher expression of ALP and osteocalcin in samples containing BG and, especially, ZnBG [16]. To improve antibacterial performance, Bejarano et al. incorporated BG doped with copper (Cu) and Zn into poly (d, l-lactide) (PDLLA). The results showed that the Cu- and Zn-doped BG improved antibacterial activity [18].

When used alone, synthetic polymers usually cannot achieve the desired effect of complete bone regeneration. In recent years, biodegradable metals (BMs) have gradually become a hotspot in the field of biomedical materials used in BTE because of their good biocompatibility, degradability, and appropriate mechanical properties. Among BMs, Zn has received extensive attention due to its satisfying biological properties in promoting bone regeneration [30].

Thus, incorporating Zn into the synthetic polymers combines the advantages of both and continues to receive much attention. Previous studies have indicated that Zn can regulate cells, which stimulates osteoblastogenesis and attenuates osteoclastogenesis. In the process of bone formation, Zn regulates the secretion and expression of osteogenic markers, such as ALP, and the deposition of minerals [31, 32]. However, the role of Zn in the process of bone regeneration depends on its content. In the selected articles, Zn is usually incorporated into the polymers as a composite or coating (Fig. 3). The content of Zn in the composite materials ranged from 0.11 to 2.9 wt%; the results indicated that, in this range, Zn exerts a satisfactory bone-promoting function. While in the Zn-containing coatings, the concentration of Zn did not show consistency, generally within the range of 20 wt%, as the surface area, thickness, and density of the coatings may also count greatly in the Zn-release process. Nevertheless, it can be concluded that Zn-containing coatings have viable osteogenic effects. Moreover, several studies have indicated that Zn-coating bone implants enhance osseointegration [33, 34].

Fig. 3
figure 3

Zn incorporated into polymers as composites or coatings

Previous studies have discussed the possible toxic effect of Zn on cells. Some researchers pointed out that in the case of BGs, ~ 5 mol% Zn doping is the upper limit before cytotoxicity occurs. Additionally, only the highest concentration (5 mg/L) elution extracts inhibited the growth of mouse embryonic fibroblasts (MEFs) [35,36,37]. However, Neščáková et al. [7] showed that Zn-doped BGs (~ 8 mol% of ZnO) did not exhibit cytotoxicity towards MG-63 or MEF cells; this can be explained by the relatively low Zn2+ concentration released into the medium (1.2 mg/L). Thus, it can be inferred that, in terms of the likely effects of Zn on the biological behavior of cells, the release concentration of Zn2+ in aqueous solutions is a more accurate indicator than the proportion of Zn in materials.

In aqueous solutions, different cells respond differently to the Zn2+ concentration (Table 4).

It has been shown that Zn2+ improves the adhesion, spreading, proliferation, and migration of vascular smooth muscle cells, up to concentrations of 60–80 μM (3.9–5.2 mg/L), while higher Zn2+ concentrations (80–120 μM, 5.2–7.8 mg/L) cause opposite responses [38]. Aina et al. [34, 35] reported that a Zn2+ concentration of 1.1 mg/L increased the proliferation rate of endothelial cells, whereas cytotoxicity was observed at 2.7 mg/L. Yamaguchi et al. [39] demonstrated that, in the presence of Zn at concentrations of 10−6–10−4 M (65 μg/L–6.5 mg/L), MC3T3-E1 osteoblasts showed an increase in osteoprotegerin, Runx-2, and regucalcin mRNA expression. The results presented by the same author indicated that a Zn2+ concentration between 10 and 250 μM (0.65 and 16.25 mg/L) suppressed osteoclastogenesis of RAW264.7 [40]. The behavior of human osteoblast-like cell line SaOS-2 was also Zn-concentration-dependent; ALP expression and mineral deposition were stimulated under Zn2+ concentrations of 1–10 μM (65 μg/L–0.65 mg/L) but were inhibited at concentrations exceeding 25 μM (1.6 mg/L) [41]. Holloway et al. found that Zn concentrations lower than 10−4 mol/L (6.5 mg/L) did not affect osteoclast activity, whereas higher concentrations promoted the proliferation of tartrate-resistant acid phosphatase (TRAP)-positive cells [42]. Under Zn2+ concentrations lower than 10−9 M (65 ng/L), osteoblasts and primary murine bone marrow stromal cells showed normal proliferation, while osteogenic and adipogenic differentiation were repressed [43]. Wang et al. reported that, without affecting cell proliferation under conditions of long-term stimulation, Zn2+-passivated carbon dots (Zn-CDs) with a Zn2+ release concentration of 10−5 mol/L (0.65 mg/L) achieved the best osteogenic effect in rat bone marrow-derived mesenchymal stem cells (rBMSCs) [44].

Table 4 Summary of the effects of Zn2+ content on different cells

Rajzer et al. found out that in SBF, BG containing Zn showed less apatite formation on the surface compared to BG without Zn [20]. This could be explained that in SBF, Zn may influence the kinetics of hydroxyapatite (HA) formation and retard HA nucleation [45]. By binding to the active growth sites of HA, Zn2+ prevents its nucleation [46] (Fig. 4). Moreover, as the Zn2+ content increases, the deposition rate of HA decreases. Previous studies have pointed out that, at low concentrations, the released Zn2+ may stimulate osteoblast differentiation, osteogenic differentiation of mesenchymal stem cells (MSCs), and extracellular matrix (ECM) mineralization in vitro [47,48,49], whereas at high concentrations, Zn2+ reduces ECM mineralization and may cause cytotoxicity [3031]. It can be concluded that, with appropriate concentrations and release behavior, the addition of Zn to a bone scaffold is likely to promote bone regeneration.

Fig. 4
figure 4

Zn2+ prevents HA nucleation by binding to its active growth sites

The mechanism by which Zn promotes bone formation has been partially explored but remains unclear. Zhu et al. used pure Zn disks and human mesenchymal stem cells (hMSCs) to show that intracellular Zn2+ triggers Ca2+ responses by activating the cyclic adenosine monophosphate protein-kinase A (cAMP-PKA) pathway, followed by the mitogen-activated protein kinase (MAPK) pathway. The Gαq-PLC-AKT pathway is also activated by Zn2+; the activity in all of these pathways enhances cell growth and differentiation, ECM mineralization, and differential regulation of genes [50]. Similarly, Park et al. suggested that ZnSO4 exerts osteogenic effects on human bone marrow-derived mesenchymal stem cells (hBMSCs) by activating RUNX2 via the cAMP-PKA-CREB pathway [51]. Fathi and Farahzadi demonstrated that, in electromagnetic fields, ZnSO4 activated the extracellular signal-regulated protein kinase (ERK) 1/2, PKA, and Wnt/β-catenin signaling pathways to promote the osteogenesis of adipose tissue-derived MSCs [52]. Yamaguchi et al. showed that ZnSO4 attenuated the nuclear factor-kappa B (NF-κB) pathway activation induced by tumor necrosis factor-alpha (TNF-α), to promote osteoblast mineralization and suppress osteoclast differentiation, and could also mitigate the transforming growth factor-beta/bone morphogenic protein-2 (TGF-β/BMP-2)-induced inhibitory effect of TNF-α on the activation of Smad [3940]. In addition, there have been several studies of Zn-containing composite materials. Vimalraj et al. revealed that Zn-silibinin complexes regulated the miR-590/Smad7 signaling pathway to enhance osteoblast differentiation [53]. Fernandes et al. incorporated Zn and citrate into HA nanoparticles and demonstrated osteogenic differentiation of BMSCs in the absence of osteoinductive factors, with the functioning regulated, at least in part, by activation of the canonical Wnt pathway [54]. Using Zn-modified calcium silicate coatings, Yu et al. showed that the TGF-β/Smad signaling pathway plays a significant role in regulating the osteoblastic differentiation of rat bone marrow-derived pericytes [55]. Despite these findings, the mechanism by which Zn-doped synthetic polymer materials promote osteogenesis has not been fully elucidated.

Some of the 13 studies in this review used multiple ions. Deng et al. demonstrated that Ag nanoparticle-decorated and Ag/ZnO-codecorated SPEEK effectively inhibit the reproduction of Gram-positive and Gram-negative bacteria. Moreover, Ag/ZnO-codecorated SPEEK substrates were better able to enhance the biocompatibility and osteodifferentiation of MG-63 cells, likely due to the combined effect of micro-/nanoscale topological cues and Zn induction [19]. In the study of Telgerd et al., Cu not only acted as an angiogenic agent, but also enhanced angiogenesis in vitro and increased the proliferation rate of endothelial cells [21]. Liang et al. showed that silicon (Si) promotes vascularization; they integrated Si-Zn and PLGA microspheres into CPC scaffolds, in which Si4+ was released at a faster rate than Zn2+ to initially promote angiogenesis and, later, osteogenesis. This was attributed to the biodegradable PLGA microspheres, which allowed for the successional release of Si and Zn. As the sequential vascularization and osteogenesis corresponded to the natural process of bone defect restoration, the PLGA/CPC-Si/Zn group showed better bone regeneration than the PLGA/CPC-Zn group [24]. Therefore, the synergistic effect of different metal ions on the biological properties of synthetic polymers is a topic worthy of further investigation and could be explored using Zn-doped synthetic polymer materials.

Based on the MINORS, the studies in this review were of satisfactory quality. However, no clinical studies were obtained through our database searches, indicating that synthetic polymers have not yet been applied clinically in humans. Meanwhile, the lack of standardization among the studies prevented us from conducting a meta-analysis.

We believe that the main significance of this systematic review is to prove the addition of Zn to synthetic polymers promotes bone regeneration; moreover, the viable ranges of Zn2+ concentrations for different cell types were demonstrated. However, additional research is necessary to uncover the mechanism by which Zn-doped synthetic polymer materials promote bone regeneration and other potential clinical applications. Finally, the effect of Zn addition on other properties (e.g., mechanical, degradation, and biocompatibility properties) should be examined further.

Conclusion

Synthetic polymers are commonly used in bone regeneration; however, their ability to promote bone formation is limited when used alone. The introduction of Zn may address this issue. In this systematic review, an overview of the application and effectiveness of Zn incorporation into synthetic polymers is presented. The results showed that, with appropriate concentrations and release behavior, Zn-containing synthetic polymers have the potential to promote bone regeneration. However, the mechanism of action and feasibility for clinical application require further study.

Availability of data and materials

All supporting data are included in the article and its supplementary files.

Abbreviations

Zn:

Zinc

3D:

Three-dimensional

PLA:

Polylactic acid

PCL:

Polycaprolactone

PGA:

Poly (glycolic acid)

PLGA:

Poly (lactic-co-glycolic) acid

ZnSO4 :

Zinc sulfate

swine ASC:

Swine adipose-derived stem cells

pASCs:

Porcine adipose-derived stem cells

ALP:

Alkaline phosphatase

mBMSCs:

Mouse bone marrow-derived mesenchymal stem cells

PLDLA:

Poly-l-d,l-lactic acid

MTS:

3-(4,5-Dimethylthiazol-2-yl)-5(3-carboxymethonyphenol)-2-(4-sulfophenyl)-2H-tetrazolium

rBMSCs:

Rat bone marrow mesenchymal stem cells

PES:

Polyethersulphone

PEG:

Polyethyleneglycol

MTT:

3-(4,5-Dimethylthiazolyl-2)-2,5-diphenyltetrazolium bromide

hMSCs:

Human mesenchymal stem cells

PDLLA:

Poly (d, l-lactide)

CCK-8:

Cell Counting Kit-8

ST-2:

Bone marrow stromal cells

PEEK:

Polyetheretherketone

OD:

Optical density

PLLA:

Poly-l-lactic acid

PAA:

Poly (amino acids)

PDLA:

Poly (dl-lactide)

PEA:

Poly (ester amide)

PEU:

Poly (ester urea)

PLGA:

Poly (lactic-co-glycolic acid)

CPC:

Calcium phosphate cement

CMC:

Carboxymethyl cellulose

PVP:

Polyvinyl pyrrolidone

HAP:

Hydroxyapatite

Runx2:

runt-related transcription factor 2

Col I:

Collagen type 1

OCN:

Osteocalcin

BMP-2:

Bone morphogenetic protein-2

NPs:

Nanoparticles

Dox:

Doxycycline

SD:

Sprague-Dawley

ARS:

Alizarin Red S

BMSCs:

Bone marrow-derived mesenchymal stem cells

PAA:

Polyacrylic acid

BG:

Bioactive glass

SBF:

Simulated body fluid

BMP-2:

Bone morphogenetic protein 2

Cu:

Copper

MEFs:

Mouse embryonic fibroblasts

TRAP:

Tartrate-resistant acid phosphatase

HA:

Hydroxyapatite

MSCs:

Mesenchymal stem cells

ECM:

Extracellular matrix

cAMP-PKA:

Cyclic adenosine monophosphate protein-kinase A

MAPK:

Mitogen-activated protein kinase

ERK:

Extracellular signal-regulated protein kinase

NF-κB:

Nuclear factor-kappa B

TNF-α:

Tumor necrosis factor-alpha

TGF-β:

Transforming growth factor-beta

Si:

Silicon

References

  1. Wang S, Li R, Xu Y, Xia D, Zhu Y, Yoon J, et al. Fabrication and application of a 3D-printed poly-ε-caprolactone cage scaffold for bone tissue engineering. Biomed Res Int. 2020;2020:2087475.

    PubMed  PubMed Central  Google Scholar 

  2. Kashte S, Jaiswal AK, Kadam S. Artificial bone via bone tissue engineering: current scenario and challenges. Tissue Eng Regen Med. 2017;14(1):1–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Bose S, Roy M, Bandyopadhyay A. Recent advances in bone tissue engineering scaffolds. Trends Biotechnol. 2012;30(10):546–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Babilotte J, Guduric V, Le Nihouannen D, Naveau A, Fricain JC, Catros S. 3D printed polymer-mineral composite biomaterials for bone tissue engineering: fabrication and characterization. J Biomed Mater Res B Appl Biomater. 2019;107(8):2579–95.

    Article  CAS  PubMed  Google Scholar 

  5. Hasan A. Tissue engineering for artificial organs: regenerative medicine, smart diagnostics and personalized medicine. Weinheim: WILEY-VCH Verlag GmbH & Co. KGaA; 2017. p. 243–95.

    Book  Google Scholar 

  6. Kenry LB. Recent advances in biodegradable conducting polymers and their biomedical applications. Biomacromolecules. 2018;19(6):1783–803.

    Article  CAS  PubMed  Google Scholar 

  7. Neščáková Z, Zheng K, Liverani L, Nawaz Q, Galusková D, Kaňková H, et al. Multifunctional zinc ion doped sol - gel derived mesoporous bioactive glass nanoparticles for biomedical applications. Bioact Mater. 2019;4:312–21.

    Article  PubMed  PubMed Central  Google Scholar 

  8. Yusa K, Yamamoto O, Iino M, Takano H, Fukuda M, Qiao Z, et al. Eluted zinc ions stimulate osteoblast differentiation and mineralization in human dental pulp stem cells for bone tissue engineering. Arch Oral Biol. 2016;71:162–9.

    Article  CAS  PubMed  Google Scholar 

  9. Luo X, Barbieri D, Davison N, Yan Y, de Bruijn JD, Yuan H. Zinc in calcium phosphate mediates bone induction: in vitro and in vivo model. Acta Biomater. 2014;10(1):477–85.

    Article  CAS  PubMed  Google Scholar 

  10. Bertels JC, Rubessa M, Schreiber SR, Wheeler MB. The effect of zinc on the differentiation of adipose-derived stem cells into osteoblasts. Reprod Fertil Dev. 2016;29:207.

    Article  Google Scholar 

  11. Tiffany AS, Gray DL, Woods TJ, Subedi K, Harley BAC. The inclusion of zinc into mineralized collagen scaffolds for craniofacial bone repair applications. Acta Biomater. 2019;93:86–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Xiong K, Zhang J, Zhu Y, Chen L, Ye J. Zinc doping induced differences in the surface composition, surface morphology and osteogenesis performance of the calcium phosphate cement hydration products. Mater Sci Eng C Mater Biol Appl. 2019;105:110065.

    Article  CAS  PubMed  Google Scholar 

  13. Moher D, Liberati A, Tetzlaff J, Altman DG. Preferred reporting items for systematic reviews and meta-analyses: the PRISMA statement. Plos Med. 2009;6(7):e1000097.

    Article  PubMed  PubMed Central  Google Scholar 

  14. Slim K, Nini E, Forestier D, Kwiatkowski F, Panis Y, Chipponi J. Methodological index for non-randomized studies (minors): development and validation of a new instrument. ANZ J Surg. 2003;73(9):712–6.

    Article  PubMed  Google Scholar 

  15. Leão RS, Moraes SLD, Gomes JML, Lemos CAA, Casado B, Vasconcelos B, et al. Influence of addition of zirconia on PMMA: a systematic review. Mater Sci Eng C Mater Biol Appl. 2020;106:110292.

    Article  PubMed  CAS  Google Scholar 

  16. Oh SA, Won JE, Kim HW. Composite membranes of poly (lactic acid) with zinc-added bioactive glass as a guiding matrix for osteogenic differentiation of bone marrow mesenchymal stem cells. J Biomater Appl. 2012;27(4):413–22.

    Article  CAS  PubMed  Google Scholar 

  17. Amiri B, Ghollasi M, Shahrousvand M, Kamali M, Salimi A. Osteoblast differentiation of mesenchymal stem cells on modified PES-PEG electrospun fibrous composites loaded with Zn2SiO4 bioceramic nanoparticles. Differentiation. 2016;92(4):148–58.

    Article  CAS  PubMed  Google Scholar 

  18. Bejarano J, Detsch R, Boccaccini AR, Palza H. PDLLA scaffolds with Cu- and Zn-doped bioactive glasses having multifunctional properties for bone regeneration. J Biomed Mater Res A. 2017;105(3):746–56.

    Article  CAS  PubMed  Google Scholar 

  19. Deng Y, Yang L, Huang X, Chen J, Shi X, Yang W, et al. Dual Ag/ZnO-decorated micro-/nanoporous sulfonated polyetheretherketone with superior antibacterial capability and biocompatibility via layer-by-layer self-assembly strategy. Macromol Biosci. 2018;18(7):e1800028.

    Article  PubMed  CAS  Google Scholar 

  20. Rajzer I, Dziadek M, Kurowska A, Cholewa-Kowalska K, Ziąbka M, Menaszek E, et al. Electrospun polycaprolactone membranes with Zn-doped bioglass for nasal tissues treatment. J Mater Sci Mater Med. 2019;30(7):80.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  21. Telgerd MD, Sadeghinia M, Birhanu G, Daryasari MP, Zandi-Karimi A, Sadeghinia A, et al. Enhanced osteogenic differentiation of mesenchymal stem cells on metal-organic framework based on copper, zinc, and imidazole coated poly-l-lactic acid nanofiber scaffolds. J Biomed Mater Res A. 2019;107(8):1841–8.

    CAS  PubMed  Google Scholar 

  22. Li S, Chen X, Wang X, Xiong Y, Yan Y, Tan Z, et al. Simonkolleite coating on poly (amino acids) to improve osteogenesis and suppress osteoclast formation in vitro. Polymers (Basel). 2019;11(9):1505–20.

  23. Neto AS, Fonseca AC, Abrantes JCC, Coelho JFJ, Ferreira JMF. Surface functionalization of cuttlefish bone-derived biphasic calcium phosphate scaffolds with polymeric coatings. Mater Sci Eng C Mater Biol Appl. 2019;105:110014.

    Article  CAS  PubMed  Google Scholar 

  24. Liang W, Gao M, Lou J, Bai Y, Zhang J, Lu T, et al. Integrating silicon/zinc dual elements with PLGA microspheres in calcium phosphate cement scaffolds synergistically enhances bone regeneration. J Mater Chem B. 2020;8(15):3038–49.

    Article  CAS  PubMed  Google Scholar 

  25. Kandasamy S, Narayanan V, Sumathi S. Zinc and manganese substituted hydroxyapatite/CMC/PVP electrospun composite for bone repair applications. Int J Biol Macromol. 2020;145:1018–30.

    Article  CAS  PubMed  Google Scholar 

  26. Ahmadzadeh E, Talebnia F, Tabatabaei M, Ahmadzadeh H, Mostaghaci B. Osteoconductive composite graft based on bacterial synthesized hydroxyapatite nanoparticles doped with different ions: from synthesis to in vivo studies. Nanomedicine. 2016;12(5):1387–95.

    Article  CAS  PubMed  Google Scholar 

  27. Toledano M, Toledano-Osorio M, Osorio R, Carrasco-Carmona Á, Gutiérrez-Pérez JL, Gutiérrez-Corrales A, et al. Doxycycline and zinc loaded silica-nanofibrous polymers as biomaterials for bone regeneration. Polymers (Basel). 2020;12(5):1201–22.

  28. Toledano M, Gutierrez-Pérez JL, Gutierrez-Corrales A, Serrera-Figallo MA, Toledano-Osorio M, Rosales-Leal JI, et al. Novel non-resorbable polymeric-nanostructured scaffolds for guided bone regeneration. Clin Oral Investig. 2020;24(6):2037–49.

    Article  PubMed  Google Scholar 

  29. Grad S, Kupcsik L, Gorna K, Gogolewski S, Alini M. The use of biodegradable polyurethane scaffolds for cartilage tissue engineering: potential and limitations. Biomaterials. 2003;24(28):5163–71.

    Article  CAS  PubMed  Google Scholar 

  30. Zheng Y, Gu X, Witte F. Biodegradable metals. Mater Sci Eng R Rep. 2014;77:1–34.

    Article  Google Scholar 

  31. Wang W, Yeung KWK. Bone grafts and biomaterials substitutes for bone defect repair: a review. Bioact Mater. 2017;2(4):224–47.

    Article  PubMed  PubMed Central  Google Scholar 

  32. Aina V, Perardi A, Bergandi L, Malavasi G, Menabue L, Morterra C, et al. Cytotoxicity of zinc-containing bioactive glasses in contact with human osteoblasts. Chem Biol Interact. 2007;167(3):207–18.

    Article  CAS  PubMed  Google Scholar 

  33. Qiao Y, Zhang W, Tian P, Meng F, Zhu H, Jiang X, et al. Stimulation of bone growth following zinc incorporation into biomaterials. Biomaterials. 2014;35(25):6882–97.

    Article  CAS  PubMed  Google Scholar 

  34. Li X, Li Y, Peng S, Ye B, Lin W, Hu J. Effect of zinc ions on improving implant fixation in osteoporotic bone. Connect Tissue Res. 2013;54(4–5):290–6.

    Article  CAS  PubMed  Google Scholar 

  35. Aina V, Malavasi G, Fiorio Pla A, Munaron L, Morterra C. Zinc-containing bioactive glasses: surface reactivity and behaviour towards endothelial cells. Acta Biomater. 2009;5(4):1211–22.

    Article  CAS  PubMed  Google Scholar 

  36. Haimi S, Gorianc G, Moimas L, Lindroos B, Huhtala H, Räty S, et al. Characterization of zinc-releasing three-dimensional bioactive glass scaffolds and their effect on human adipose stem cell proliferation and osteogenic differentiation. Acta Biomater. 2009;5(8):3122–31.

    Article  CAS  PubMed  Google Scholar 

  37. Salih V, Patel A, Knowles JC. Zinc-containing phosphate-based glasses for tissue engineering. Biomed Mater. 2007;2(1):11–20.

    Article  CAS  PubMed  Google Scholar 

  38. Ma J, Zhao N, Zhu D. Bioabsorbable zinc ion induced biphasic cellular responses in vascular smooth muscle cells. Sci Rep. 2016;6:26661.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Yamaguchi M, Goto M, Uchiyama S, Nakagawa T. Effect of zinc on gene expression in osteoblastic MC3T3-E1 cells: enhancement of Runx2, OPG, and regucalcin mRNA expressions. Mol Cell Biochem. 2008;312(1–2):157–66.

    Article  CAS  PubMed  Google Scholar 

  40. Yamaguchi M, Weitzmann MN. Zinc stimulates osteoblastogenesis and suppresses osteoclastogenesis by antagonizing NF-κB activation. Mol Cell Biochem. 2011;355(1–2):179–86.

    Article  CAS  PubMed  Google Scholar 

  41. Cerovic A, Miletic I, Sobajic S, Blagojevic D, Radusinovic M, El-Sohemy A. Effects of zinc on the mineralization of bone nodules from human osteoblast-like cells. Biol Trace Elem Res. 2007;116(1):61–71.

    Article  CAS  PubMed  Google Scholar 

  42. Holloway WR, Collier FM, Herbst RE, Hodge JM, Nicholson GC. Osteoblast-mediated effects of zinc on isolated rat osteoclasts: inhibition of bone resorption and enhancement of osteoclast number. Bone. 1996;19(2):137–42.

    Article  CAS  PubMed  Google Scholar 

  43. Wang T, Zhang JC, Chen Y, Xiao PG, Yang MS. Effect of zinc ion on the osteogenic and adipogenic differentiation of mouse primary bone marrow stromal cells and the adipocytic trans-differentiation of mouse primary osteoblasts. J Trace Elem Med Biol. 2007;21(2):84–91.

    Article  PubMed  CAS  Google Scholar 

  44. Wang B, Yang M, Liu L, Yan G, Yan H, Feng J, et al. Osteogenic potential of Zn2+-passivated carbon dots for bone regeneration in vivo. Biomater Sci. 2019;7(12):5414–23.

    Article  CAS  PubMed  Google Scholar 

  45. Miola M, Verné E, Ciraldo FE, Cordero-Arias L, Boccaccini AR. Electrophoretic deposition of chitosan/45S5 bioactive glass composite coatings doped with Zn and Sr. Front Bioeng Biotechnol. 2015;3:159.

    Article  PubMed  PubMed Central  Google Scholar 

  46. Du RL, Chang J, Ni SY, Zhai WY, Wang JY. Characterization and in vitro bioactivity of zinc-containing bioactive glass and glass-ceramics. J Biomater Appl. 2006;20(4):341–60.

    Article  CAS  PubMed  Google Scholar 

  47. Su Y, Cockerill I, Wang Y, Qin YX, Chang L, Zheng Y, et al. Zinc-based biomaterials for regeneration and therapy. Trends Biotechnol. 2019;37(4):428–41.

    Article  CAS  PubMed  Google Scholar 

  48. Oh SA, Kim SH, Won JE, Kim JJ, Shin US, Kim HW. Effects on growth and osteogenic differentiation of mesenchymal stem cells by the zinc-added sol-gel bioactive glass granules. J Tissue Eng. 2011;2010:475260.

    PubMed  PubMed Central  Google Scholar 

  49. Wang X, Li X, Ito A, Sogo Y. Synthesis and characterization of hierarchically macroporous and mesoporous CaO-MO-SiO (2)-P (2) O (5) (M=Mg, Zn, Sr) bioactive glass scaffolds. Acta Biomater. 2011;7(10):3638–44.

    Article  CAS  PubMed  Google Scholar 

  50. Zhu D, Su Y, Young ML, Ma J, Zheng Y, Tang L. Biological responses and mechanisms of human bone marrow mesenchymal stem cells to Zn and Mg biomaterials. ACS Appl Mater Interfaces. 2017;9(33):27453–61.

    Article  CAS  PubMed  Google Scholar 

  51. Park KH, Choi Y, Yoon DS, Lee KM, Kim D, Lee JW. Zinc promotes osteoblast differentiation in human mesenchymal stem cells via activation of the cAMP-PKA-CREB signaling pathway. Stem Cells Dev. 2018;27(16):1125–35.

    Article  CAS  PubMed  Google Scholar 

  52. Fathi E, Farahzadi R. Enhancement of osteogenic differentiation of rat adipose tissue-derived mesenchymal stem cells by zinc sulphate under electromagnetic field via the PKA, ERK1/2 and Wnt/β-catenin signaling pathways. PLoS One. 2017;12(3):e0173877.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  53. Vimalraj S, Rajalakshmi S, Saravanan S, Raj Preeth D, LAV R, Shairam M, et al. Synthesis and characterization of zinc-silibinin complexes: a potential bioactive compound with angiogenic, and antibacterial activity for bone tissue engineering. Colloids Surf B Biointerfaces. 2018;167:134–43.

    Article  CAS  PubMed  Google Scholar 

  54. Fernandes MH, Alves MM, Cebotarenco M, Ribeiro IAC, Grenho L, Gomes PS, et al. Citrate zinc hydroxyapatite nanorods with enhanced cytocompatibility and osteogenesis for bone regeneration. Mater Sci Eng C Mater Biol Appl. 2020;115:111147.

    Article  CAS  PubMed  Google Scholar 

  55. Yu J, Xu L, Li K, Xie N, Xi Y, Wang Y, et al. Zinc-modified calcium silicate coatings promote osteogenic differentiation through TGF-β/Smad pathway and osseointegration in osteopenic rabbits. Sci Rep. 2017;7(1):3440.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

Not applicable.

Author information

Authors and Affiliations

Authors

Contributions

Wang S contributed to the writing of the first draft of this review. Li R and Zhao X performed the literature search and collected the data. Zhu Y and Gu R made the quality assessment. Xia D and Yoon J contributed to the finalizing of the review. Liu Y contributed to the methodology. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Yunsong Liu.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1

: Table S1. Search strategies in PubMed database and related results.

Additional file 2

: Table S2. Search strategies in Embase database and related results.

Additional file 3

: Table S3. Search strategies in Cochrane Library database and related results.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, S., Li, R., Xia, D. et al. The impact of Zn-doped synthetic polymer materials on bone regeneration: a systematic review. Stem Cell Res Ther 12, 123 (2021). https://doi.org/10.1186/s13287-021-02195-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13287-021-02195-y

Keywords