Skip to main content

A review on the effect of nanocomposite scaffolds reinforced with magnetic nanoparticles in osteogenesis and healing of bone injuries

Abstract

Many problems related to disorders and defects of bone tissue caused by aging, diseases, and injuries have been solved by the multidisciplinary research field of regenerative medicine and tissue engineering. Numerous sciences, especially nanotechnology, along with tissue engineering, have greatly contributed to the repair and regeneration of tissues. Various studies have shown that the presence of magnetic nanoparticles (MNPs) in the structure of composite scaffolds increases their healing effect on bone defects. In addition, the induction of osteogenic differentiation of mesenchymal stem cells (MSCs) in the presence of these nanoparticles has been investigated and confirmed by various studies. Therefore, in the present article, the types of MNPs, their special properties, and their application in the healing of damaged bone tissue have been reviewed. Also, the molecular effects of MNPs on cell behavior, especially in osteogenesis, have been discussed. Finally, the present article includes the potential applications of MNP-containing nanocomposite scaffolds in bone lesions and injuries. In summary, this review article highlights nanocomposite scaffolds containing MNPs as a solution for treating bone defects in tissue engineering and regenerative medicine.

Introduction

Bone tissue as a mineral connective tissue has important functions in the body, which is disturbed by many bone disorders such as bone fractures or defects caused by surgery, trauma, or primary tumor removal. Therefore, there is a growing need for safe and cost-effective treatment approaches for damaged bone tissues. On the other hand, the use of autograft and allograft products to repair damaged bone tissue is associated with the risk of disease transmission, chronic pain, infection, possible immunogenicity, lack of supply, and increased operation time [1]. It is anticipated that in the future, the high potential of bone for regeneration will lead to the introduction and development of bone tissue engineering (BTE) as a promising supporter of bone grafting techniques in regenerative medicine. Compared to conventional micro-sized materials, nanosized biomaterials have been shown to have the ability to induce cell adhesion and proliferation, and further bone regeneration [2]. Therefore, nanostructured biomaterials including biopolymer matrices and bioactive fillers in nanodimensions have been developed in nanocomposites-based scaffolds for use in tissue engineering, especially BTE [2,3,4]. Due to their small size and large surface area, these structures show good interaction with different cells and tissues and can facilitate absorption by cells.

In bone defects, in addition to the biological stimuli that originate from the understanding of bone biology, bone regeneration, and fracture repair, they can be stimulated by exogenous or endogenous physical factors such as fluid shear stresses, tensile and compressive stresses, and heat [5, 6]. Interestingly, magnetic stimulations from electromagnetic fields (EMFs) and static magnetic fields (SMFs) can also significantly progress bone repair and regeneration [7, 8]. Magnetic nanoparticles (MNPs), mainly superparamagnetic iron oxide nanoparticles (SPIONs), of small size (about 1–100 nm), with or without a magnetic field, can influence the function of ion channels in stem cells and regulate osteogenic differentiation. It also affects various biochemical pathways, scaffold activity, and growth factor turnover and can be used as a modulator in BTE [9,10,11,12]. In addition to dynamic magneto-mechanical stimulation, MNPs and magnetic fields provide the necessary growth factors, drugs, and gene transfections to accelerate the regeneration and repair of damaged bone [13,14,15]. Surprisingly, static magnetic field (SMF) expands the proliferation, migration, and differentiation of bone marrow mesenchymal stem cells (MSCs) into osteoblast-like cells and osteogenic cells [16]. In addition, electromagnetic field (EMF) also regulates the expression of type II collagen, thereby stimulating cartilage formation through the differentiation of human bone marrow MSCs into chondrocytes [17]. MNPs also have the potential to be used as magnetic resonance imaging (MRI) contrast agents for tracking implanted cells, bone regeneration, and scaffold degradation. Therefore, making an optimal magnetic scaffold in BTE can be very useful and effective.

MNPs are one of the most important nanomaterials that are assembled in nanocomposite scaffolds, and hence these types of scaffolds are called magnetic nanocomposite scaffolds. Among the scaffolds that support cell activity, nanofibrous scaffolds provide ideal substrates for cell adhesion and nutrient transport due to their highly porous structure, adjustable fiber diameter, and special network shape [18]. In particular, the integration of MNPs with polymer fibers solves the agglomeration problem and improves the stability of magnetic nanocomposite scaffolds produced with diverse morphologies [19, 20]. Compared to microfibers, nanofibers display better surface adsorption to proteins and enhanced cell adhesion and proliferation, which designates the importance of using a bimodal structure to achieve larger pores and an appropriate surface for cell adhesion and biological applications [21]. On the other hand, the use of these scaffolds in combination with MNPs increases cell attachment, and cell survival, and also increases the mechanical strength of the scaffold, all of which are needed for bone regeneration [22].

Therefore, it is important to have an overview of magnetic nanocomposite scaffolds used in bone tissue repair and regeneration. In this review, after the introduction of bone tissue engineering and types of magnetic nanoparticles and their application, the effects of various magnetic nanocomposite scaffolds used in bone tissue engineering on adhesion, penetration, proliferation, and specific differentiation of stem cells have been discussed.

Bone tissue engineering

Bone tissue engineering (BTE) is an emerging field that aims to incorporate three components, including (1) osteogenic cells generating the bone tissue matrix, (2) a biocompatible scaffold mimicking bone extracellular matrix (ECM), and (3) physico-chemical stimuli affecting cell behavior. A successfully engineered bone product has no permanent graft site complications, exhibits adequate vascularization, and does not induce an immunologic reaction at the defect site [23].

Over time, the polymer scaffold that supports the tissue regeneration process is absorbed or replaced by newly produced bone tissue [24]. As three-dimensional structures with high porosity, various scaffolds are widely used in bone tissue engineering and support cell-biomaterial interactions, cell adhesion, growth, and migration [2, 4]. At the same time, they also facilitate the transport, survival, proliferation, and differentiation of progenitor cells [25]. A suitable biomaterial for the construction of scaffolds in bone tissue engineering should have critical behaviors such as osteo-induction and osteo-conduction and also be able to maintain bone integrity so that the components of the biomaterial integrate into the surrounding bone tissue [26, 27]. In addition, easy sterilizability, easy fabrication, non-thrombogenicity, and stability in different chemical and mechanical conditions are other key parameters that should be considered in biomaterials used in BTE [28]. Furthermore, when the scaffolds are implanted in the bone defect, they must have sufficient stability and elasticity to withstand the suture site and, in addition, be able to support bone formation with homogeneous morphology. Finally, the implanted scaffold must be degraded in a controlled manner in vivo without or with a minimal degree of inflammatory or toxic side effects [29, 30].

Magnetic nanoparticles and magnetic nanocomposite

According to the number of nanophases (i.e., < 100 nm) of homogeneous solid materials used to make nanocomposites, they can be one-dimensional (e.g., thin films and surface coatings), two-dimensional (e.g., nanowires or nanotubes), or three-dimensional (e.g., multilayer structures). In other definitions, in the nanocomposite structure, the interphase spacing is repeated in the nanoscale range [31, 32]. Nanoparticles (NPs) are materials that are articulated with all three external dimensions at the nanoscale, exhibit a high total interfacial area and high surface-to-volume ratio, and create extraordinary interactions with molecular and supramolecular structures in biological environments [33]. Magnetism is the magnetic moment per unit volume of a particle, which is typically dependent on the spin or orbital energy that the dipole possesses. Therefore, the magnetic behavior is affected by the sample temperature and the degree of magnetic order [34].

In general, the synthesis of magnetic nanoparticles is done using two processes: top down or bottom up, which include different techniques such as ultrasonication, radiation, electrochemical, vapor deposition, and microwave [35, 36]. In this regard, MNPs (metal oxide NPs and SPIONs) can be fabricated using various methods, including co-precipitation, surfactant-assisted techniques, solgel and hydro-thermal processing, and emulsion techniques. These methods are used to control the structure, surface morphology, and stability of the manufactured nanoparticles [37,38,39]. Among the challenges of creating monodisperse magnetic nanostructures are size control, particle surface effects, and dipole interactions. However, some new chemical synthesis methods have made it easier to make functional MNPs. MNPs with a size of less than 30 nm are the dominant superparamagnetic nanoparticles [40, 41]. For example, superparamagnetic magnetite nanocrystal clusters (SMNC) as developed MNPs can be synthesized via mini-emulsion/solgel and polyol systems and used for multifunctional applications including combined drug targeting and cell imaging [42].

Iron oxide nanoparticles in two forms, Fe3O4 and Fe2O3, are the most common MNPs that are usually synthesized by conventional co-precipitation method [43] and are widely used in MRI applications for imaging cancer cells, and in vivo tracking and monitoring of cells and transplanted tissues [44]. The excellent biocompatibility and low toxicity of these types of MNPs led researchers to use them in the biomedical field, especially for monitoring engineered tissues [45]. Fe3O4 has been accepted by the US Food and Drug Administration (FDA) for clinical use [46]. Potential applications of MNPs include organ regeneration, tissue implants, drug delivery, imaging, and improved diagnostics.

In line with what was mentioned, in the field of tissue repair and tissue engineering, the unique feature of large surface area in biodegradable nanofillers complements the critical parameters mentioned above for classical scaffolds (e.g., biocompatibility, physicochemical stability, support of cell adhesion, and differentiation) which ultimately effectively improves the BTE approach [47, 48]. In bone tissue engineering, various technologies, including the foam replica method [49], solvent casting and particulate-leaching [50], freeze-drying [51], phase separation [52], gas foaming [53], rapid prototyping [54], and electrospinning [55], have been used to develop and introduce efficient nanocomposite scaffolds with controllable size and porosity that show a high surface-to-volume ratio. However, most scaffolds cannot be controlled after implantation in vivo, and since the repair process can only be finalized by scaffolds in vivo, then the repair will not always be good enough. Therefore, one of the suggested approaches to achieve appropriate tissue repair and the possibility of controlling tissue fate using external stimuli is the construction of composite scaffolds containing magnetic nanoparticles.

Since MNPs can respond to EMF, scaffolds containing MNPs can also respond to external magnetic fields. In addition, MNPs and magnetic responsive scaffolds (MRS) can deliver various peptide agents, and also improve implant stabilization [56], improve mechanical properties and biocompatibility [57, 58], and improve the wettability of the scaffolds. Moreover, MNPs and MRS increase alkaline phosphatase (ALP) activity and osteogenic gene expression of bone cells [59, 60]. Therefore, combining the potentials of MNPs and nanocomposite scaffolds leads to the development of implantable and functional magnetic nanocomposites (MNCs) [61, 62]. The most important aspect of these scaffolds is the magnetic response, which leads to great progress in tissue engineering such as magnetic patterning of cells and 3D tissue-like structures [63,64,65].

Recently, Panseri et al. made magnetic scaffolds composed of hydroxyapatite/collagen and MNP and showed that the presence of magnetic particles causes the attraction of growth factors and cells [65]. Also, embedding iron oxide nanoparticles in macro-porous ferrogel scaffolds led to the production of MNC with a porous structure optimized for cell delivery [66]. In addition, doping of magnetic poly (1-caprolactone)/iron into hydroxyapatite leads to the fabrication of MNCs capable of simultaneous use in the repair of damaged tissues and the treatment of further hyperthermia [67]. It is noteworthy that each particle in the structure of MNPs owns a single magnetic domain, and thus, the incorporation of MNPs into scaffolds generates a nanoscale magnetic field that affects the cells and scaffolds interaction in the exposed microenvironments. In addition, the endogenous force exerted by MNPs or exogenous stimulation using a magnetic field has been shown to affect multiple cell surface receptors as well as associated signaling pathways to modulate cell function toward a specific target [68,69,70].

Magnetic field stimulation also accelerates the bone healing process by enhancing the integration of scaffolds and host bone and increases calcium content for bone density and new bone formation [71,72,73]. It has also been shown that functionalized MNPs injected near the scaffold can be absorbed in the damaged site under the influence of an external magnetic field to promote tissue regeneration [74]. Remarkably, as previously reported for wound healing, a moderate external static magnetic field can modulate osteoblast differentiation even without the presence of MNPs [75,76,77]. In general, the intrinsic magnetic properties of MNCs or exposure to far magnetic fields combined with appropriate mechanical support lead to the regulation of signaling pathways and various biological responses, the promotion of osteogenic cell differentiation, and ultimately bone regeneration and injury repair [78, 79].

Despite all the positive features and wide applications, one of the limitations of using magnetic nanoparticles is their instability and low solubility in water environments [80]. To deal with this problem and to increase their stability, using a hydrophilic polymer substrate and covering magnetic nanoparticles is recommended [81]. Also, to prevent aggregation, magnetic nanoparticles should either be stabilized by electrostatic or steric repulsion [82]. Another limitation of MNPs is the toxicity related to size, shape, and chemistry, which should be considered before clinical use. Therefore, while MNPs seem to have significant therapeutic, restorative, and diagnostic potential, long-term evaluation is necessary to reduce human health risks [83]. Hence, they require comprehensive pre-evaluation in terms of bio-distribution and biocompatibility. It should also be noted that MNPs, either alone or in a polymer substrate, must be quickly removed from the body after reaching the therapeutic endpoint. Considering all these, it is concluded that the development of MNPs for therapeutic purposes is in its infancy. It appears that the collaborative multidisciplinary science that will deliver the next generation of MNPs is related to theranostics where combined diagnosis and treatment will eventually become part of standard medical practice [84]. Therefore, it is expected that in the near future interdisciplinary research in the fields of biology, engineering, physics, and chemistry can provide new technologies based on magnetic nanoparticles to improve global health.

Molecular effect of MNPs on osteogenic differentiation and bone repair

The differentiation of mesenchymal stem cells (MSCs) plays a crucial role in bone repair and regeneration. The process of osteogenic differentiation of MSCs is a complex and intricate process that involves the activation of several signaling pathways and transcription factors, such as Wnt/β-catenin, Notch, BMP/TGF-β, PI3K/Akt/mTOR, MAPK, PDGF, IGF, FGF, and Ca2+ pathways [85,86,87]. Magnetic stimulations from MNPs alter cell behavior and have great potential for BTE applications. The use of magnetic field stimulation has been shown to activate multiple sensitive receptors on the surface of cells and trigger related signaling pathways, resulting in enhanced cell activity [69]. Continuous application of magnetic field acts as sustained stimulation to further increase cellular activity in the bone defect [70]. Additionally, magnetic field stimulation can cause better integration of scaffolds and host bone, promote new bone density through increased calcium content, and ultimately accelerate bone healing. These nanoparticles that have a particle size smaller than 30 nm exhibit a superparamagnetic effect and can be seen as individual magnetic domains [40, 41]. Therefore, it is reasonable to assume that MNPs in the microenvironments within the scaffolds can create a nanoscale magnetic field and exert a micromagnetic driving force at the interface between the scaffolds and cells, thereby affecting the fate of cells and subsequently in tissue repair. This happens by activating several sensitive receptors on cell surfaces, increasing cell activity and enhancing bone formation during bone healing.

The incorporation of MNPs into the structure of scaffolds can promote the osteogenic differentiation of MSCs in several ways. MNPs increase the surface roughness of scaffolds, which enhances surface energy and protein absorption, promoting the interaction between cells and scaffolds and activating the integrin signaling pathway. From the perspective of molecular mechanisms, magnetic induction stimulates cellular receptors and ECM components to activate the signaling pathways including the Wnt/β-catenin, integrin, and BMP2 (Fig. 1). The studies indicated that exposure of stem cells to a low-frequency magnetic field enhanced the expression of Wnt3a. This ligand activates the Wnt/β-catenin pathway and improves osteogenic differentiation of human bone marrow-derived mesenchymal stem cells (BMSCs) [88, 89].

Fig. 1
figure 1

Schematic image of the signaling pathways activated by magnetic nanocomposite scaffolds inducing osteogenesis (Designed by CorelDRAW 2019)

ECM proteins (e.g., Integrins and collagens) offer a cell adhesion substrate for cellular functions that allows the development of subsequent mineralization. Integrins are transmembrane receptors that can potentially transmit ECM physicochemical and mechanical conditions to the cells. MNPs have excellent specific hardness and strength as well as large specific surface area. These properties can be utilized to enhance the mechanical properties of scaffolds through their ability to reinforce and strengthen the structure [90]. The mechanical properties of scaffolds, including their stiffness, can influence the fate of combined stem cells. High matrix stiffness can induce MSCs to differentiate into osteoblasts, so that the upper and lower stiffness of the matrix causes the osteoblast and nerve differentiation of MSCs, respectively [91]. Therefore, Integrins can mediate the matrix stiffness affecting cells and regulate osteogenic differentiation of stem cells. From a molecular point of view (Fig. 1), in the presence of high matrix stiffness, integrins can activate osteogenic differentiation-related signaling pathways such as the PI3K/Akt pathway and FAK or ERK1/2 phosphoric acid toward enabling stem cells to differentiate into osteoblasts [92, 93]. Indeed, matrix stiffness regulates the ERK pathway, which subsequently up-regulates the expression of osteogenic genes [94]. At the same time, integrin α5 activation upregulates the expression of collagen type I and Runx2 (runt-related transcription factor 2), leading to increased deposition of calcium [95]. Notably, signaling pathways such as BMP and Wnt are critical pathways promoting osteogenic differentiation that activated by integrin β3 and phosphorylation of GSK3 (glycogen synthase kinase-3) via integrin-linked kinase and β-catenin nuclear translocation [96]. The BMP signaling pathways mediated by Smad1/5/8 phosphorylation [97].

In a study using decellularized cancellous bone scaffolds coated with various proportions of collagen/HA mixtures, MSCs implanted on the high-stiffness scaffolds demonstrated superior osteogenic ability [98]. Moreover, matrix stiffness can affect cell morphology, which is related to the cytoskeleton and intercellular interactions [99]. Cells on stiffer matrices are more spread out, with enhanced expression of vinculin, promoting the formation of focal adhesion and activation of β-catenin signaling, and inducing bone formation and remodeling [100]. For example, the treatment of rat BM-MSCs via graphene oxide (GO) combination with Fe3O4 (MGO), leads to the up-regulation of BMP2 and Wnt/β-catenin signal pathways. It also provides cell protective activity where ferrous iron from Fe3O4 reacts with hydrogen peroxide (H2O2) to produce hydroxide and hydroxyl radicals [101].

Regarding osteogenic differentiation, bioinformatics analyses of genes microarray test revealed that MNPs induced MSCs osteogenic differentiation through modulation of expression of genes (e.g., ALP, COL1, RUNX2, and OCN) and induction of the signal pathway such as MAPK (classic mitogen-activated protein kinase) [11]. ZEB2 is a regulatory factor that inhibits the BMP/Smad-related osteogenic differentiation. SPIONs support sustained ossification by upregulating the INZEB2 (long non-coding RNA) in MSCs, that overexpress INZEB2 and downregulate ZEB2 expression [102]. In this regard, it was observed that gelatin sponges containing SPION increased bone regeneration with about a 1.5 times increase in BMD and BV/TV (bone volume per tissue volume) in comparison with gelatin sponges SPION-free [12].

Similar to the magnetic field, a pulsed electromagnetic field (PEMF) also induces and activates ERK1/2 and PKA (protein kinase A) signaling pathways to promote bone repair [103]. PEMF also increases the Ca2+ concentration in the cytoplasm by opening ion channels (voltage-gated) and inducing the expression of osteogenesis-related genes [104]. Moreover, static magnetic field (SMF) has a similar effect and induces the expression of Wnt and Smad4 and other signaling pathways related to osteogenic differentiation, such as nuclear factor-κB, p38, and JNK (c-Jun N-terminal kinase)/MAPK pathways [101, 105]. For example, SMFs (15 mT) and PCL/MNPs composites synergistically upregulated the osteogenesis-related genes such as Runx2 and Osterix and induced alkaline phosphatase activity in mice osteoblasts to enhance osteoblastic differentiation [78]. Overall, these pathways detect changes in microenvironments, cytoskeletons, cell membranes, and nucleoproteins induced by magnetic forces [106, 107]. In summary, the incorporation of MNPs into scaffolds by creating a magnetic field can influence their surface roughness, wettability, and mechanical properties, increase their surface energy, protein absorption, cell-scaffold interaction and stiffness, and subsequently regulate bone differentiation.

On the other hand, angiogenesis supports the delivery of nutrients and signaling factors in damaged tissues to promote the formation of new tissues [108, 109]. In this case, magnetic systems affect the secretory function of osteoblasts and indirectly on other cells. The significant point in angiogenic responses is that magnetism upregulates VEGF (vascular endothelial growth factor) and angiogenin-1 genes in endothelial cells and promotes the formation of capillary tubes [110, 111].

Magnetic nanocomposite scaffolds in bone tissue engineering

Along with other medical applications (i.e., drug delivery, biosensors, imaging, etc.), nanomedicine has a special place in regenerative medicine. Today, nanocomposite biomaterials have emerged as a new class of biocompatible materials that are used as bioactive, absorbable, and nanosized fillers for developing matrix structures [48]. Various studies have been conducted to prepare efficient magnetic nanocomposite scaffolds for use in BTE (Table 1).

Table 1 Applied MNPs in bone tissue engineering

In 2014, Wang et al. used the electrospinning technique to create a composite material containing of MNPs based on Fe3O4. They used poly (l-lactic acid) (PLLA) polymer concentrations of 2% and 5% to load nanoparticles and finally showed that the spreading and attachment of fibroblast cells to the fabricated nanocomposites was well done [67]. In addition to unique magnetic properties and high surface area, Fe3O4 MNPs have high specific hardness and specific strength, which improves the mechanical attributes of magnetic scaffolds [121]. Due to the advantages of Fe3O4 MNPs, several studies have used biomaterials containing Fe3O4 MNPs for BTE. Pan et al., who used the extrusion process to integrate Fe3O4 into Poly-l-lactide polymer to make Fe3O4/Poly-l-lactide composites, finally showed the ability of osteogenic differentiation without a cytotoxic effect on fibroblast cells in vitro [122].

Similarly, D’Amora et al. prepared the poly(e-caprolactone)/iron-doped hydroxyapatite scaffolds that enhanced cell proliferation in vitro while causing minimal toxicity [123]. Also, in a study regarding the induction of bone formation by MNPs, the incorporation of SPIONs (superparamagnetic iron oxide nanoparticles) (Fe2O3) into calcium phosphate cement (CPC) led to the construction of nanocomposite scaffolds. It was reported that compared to CPC scaffolds without SPIONs, the shape of the scaffold surface caused better adhesion and osteogenic differentiation of human dental pulp stem cells (hDPSCs). These scaffolds also released SPION into hDPSCs, thereby regulating osteogenic gene expression and ALP activity, and bone matrix mineral synthesis in cells [59, 124].

Another study using ultrasonic irradiation o synthesize nanocomposite scaffolds containing bacterial cellulose (BC), Fe3O4, and hydroxyapatite (HA) NPs indicated that HA and Fe3O4 NPs were uniformly distributed on the surface and cross-section of the BC matrix. They have also shown high porosity (81.1%) and good mechanical attributes (9.87 MPa and 1.85 GPa). BC/Fe3O4/HA developed MNC showed significant attachment to living bone cells without toxicity and also induced cell differentiation and proliferation [125]. With all these advantages, preparing magnetic nanofibers with a high content of Fe3O4 NPs is difficult. To solve this problem, the combination of different manufacturing methods, such as the combination of cooperative assembly methods with electrospinning technology, is used. This strategy has effectively overcome the mentioned challenge and displayed excellent magnetic performance, BMSCs biocompatibility, and viability, and more importantly, provided a dynamic cell culture microenvironment for a model of actual 3D growth under alternating magnetic field without direct contact with cells [126].

The freeze-casting method is another strategy to gain highly interconnected porous magnetic scaffolds appropriate for BTE. For example, a study investigated the effect of different percentages of MNPs integrated into the chitosan/silk fibroin structures. As shown in the SEM micrographs in Fig. 2, the structure of the scaffolds was layered. It was also shown that the incorporated nanoparticles did not meaningfully affect the microstructural features of the scaffolds. However, interestingly, the samples containing 1% MNPs showed a lower degradation rate. In this regard, MNPs-free scaffolds showed the highest degradation, which was attributed to the hydrophobic nature of MNPs in scaffold structures. In addition, in the mentioned study, it was shown that a higher percentage of MNPs decreases cell viability, which indicates the need to optimize the percentage of MNPs used in BTE [127]. MNPs can be considered a single nanoscale magnetic domain that may influence cell membrane ion channels to control cellular responses [128, 129]. Even though the magnetic field intensity of the nanoscale particle is very low, the overall effect is likely to be enhanced by increasing the MNP amount, thus having a more substantial impact on cellular responses. These somehow create a dynamic environment that influences the surrounding cellular reactions.

Fig. 2
figure 2

a Schematic illustration of chitosan/silk/MNPs scaffolds preparation by freeze-casting method. b SEM images of chitosan/silk and c chitosan/silk/MNPs. Reprinted from Ref. [127]. Aliramaji, S., A. Zamanian, and M. Mozafari, Super-paramagnetic responsive silk fibroin/chitosan/magnetite scaffolds with tunable pore structures for bone tissue engineering applications. Materials Science and Engineering: C, 2017. 70: p. 736–744, Copyright (2023), with permission from Elsevier

Cobalt-based NPs are another type of MNPs that has a high magnetic saturation. On the other hand, due to the high biocompatibility of Fe3O4, their combined properties with cobalt in the form of CoFe2O4 MNPs are considered as one of the most useful spinel ferrites, which have suitable characteristics such as large magneto-crystalline anisotropy, medium saturation magnetization, mechanical stiffness, high inductance, high magnetic resistance, and chemical stability [112, 130]. Cobalt can also be used in combination with zinc, as in a study where cobalt-zinc ferrite (CZF) NPs were fabricated and incorporated up to 3 wt% into poly(ε-caprolactone) nanofiber composites scaffolds. In the resulting nanofiber, the diameter was reported to decrease to 466 nm. Also, hydrophilicity, mechanical stability, biodegradability, and biocompatibility were demonstrated in the presence of EMF [131].

Similar to this study, in other studies, the integration of MNPs into other biological materials, such as ceramic or calcium phosphate scaffolds, has been used to induce cell proliferation, differentiation, and bone formation in vivo. For example, in various studies, adding HA nanoparticles to composite scaffolds or calcium phosphate scaffolds significantly increased the attachment, proliferation, and differentiation of osteoblast cells under EMF and showed tissue biocompatibility [132].

The cell-ECM and cell–cell interactions regulation, predicting drug sensitivities, providing nutrients or stimuli, and removing waste products are essential conditions for the native 3D-microenvironment of bone that classical 2D plastic scaffolds in single-cell cultures cannot provide. As a result of the absence of these conditions, cell death or loss of function and impaired repair occurs [133, 134]. These limitations are overcome by a novel 3D scaffold that provides a high degree of porosity, allowing cell interaction, waste removal, and oxygen/nutrient diffusion. In addition, 3D scaffolds show a more natural morphology of cells compared to 2D structures and provide better differentiation of them into physiologically relevant tissue. It should be mentioned that the size of the pattern, and the topography of the scaffold and the cell type can determine morphology of cell, migration, proliferation, and differentiation [135]. These observations indicate the importance of the fabrication of proper microstructures and designing 3D biomimetic scaffolds that mimic the native tissue. Of course, the design of such microenvironments needs the use of computational modeling and computer-aided instruments for tissue engineering [136].

To develop magnetic nanocomposite scaffolds containing Fe3O4, a 3D printing technique can be used and scaffolds with uniform pore size, unique morphology, and architecture can be prepared. For example, in a study, this technology was used to make magnetic nanocomposite scaffolds containing Fe3O4, bioactive glass, and polycaprolactone (Fe3O4/MBG/PCL). The results showed that these scaffolds increased proliferation rate and ALP activity and induced the human bone marrow mesenchymal stem cells (h-BMSCs) differentiation toward bone repair and regeneration [117]. Another study used PCL/Fe/HA nanoparticles to prepare 3D fully biodegradable magnetic nanocomposite scaffolds. The in vitro results showed a 2.2-fold increase in its potential in the growth of BMSCs. In addition, in the mentioned study, after four weeks, the magnetized scaffolds were filled with new bone in vivo, which confirmed the excellent tissue compatibility of the magnetic scaffolds [113]. In another study, electrospinning followed by the layer-by-layer assembly (LbL) was used to assemble a film of SPION on the surface of a magnetic scaffold containing PCL/PLGA. The results showed that hydrophilicity, elasticity, and surface roughness were greatly increased, which subsequently significantly increased cell attachment and osteogenic differentiation of ADSCs [115].

Notably, implantable scaffolds with larger pores provide a bone-like micro-environments that increase cell proliferation and migration [137, 138]. Moreover, structures with high porosity allow the fast diffusion of medium, oxygen, and metabolites, creating a favorable micro-environment for cells [139]. Addressing these notes, the solvent casting technique alongside the overlap of nylon template structures is another way to develop bioinspired 3D porous nanocomposites scaffolds comprised of piezoelectric polymers, such as PVDF (polyvinylidene fluoride) and CoFe2O4. Studies on these scaffolds show that their structures are alike to the spongy bone (pore sizes 5–20 µm). Also, due to the natural crystallization process with magneto-mechanical and magneto-electric stimulation, larger pores with more interconnectivity are created after removing the nylon template [140].

On the other hand, the use of selective laser sintering for the fabrication of PLLA/PGA (polyglycolic acid) scaffolds containing MNPs Fe3O4 leads to the construction of magnetic scaffolds that provide cell adhesion and viability (Fig. 3). According to these studies, the stiffness of MNPs increased the modulus and compressive strength of the scaffold by 71.6% and 81.9%, and at the same time, it improved cell proliferation and alkaline phosphatase activity [116].

Fig. 3
figure 3

a Schematic illustration of PLLA/PGA/MNPs scaffolds preparation by selective laser sintering. b Powders of PLLA, PGA and composite. c PLLA/PGA scaffolds fabrication with different content of MNPs. Reprinted from Ref. [116]. Shuai, C., et al., A magnetic micro-environment in scaffolds for stimulating bone regeneration. Materials & Design, 2020. 185: p. 108275, open access article distributed under the CC BY-NC-ND license

However, their validity and translation to the clinic require their functional evaluation in preclinical animal models, including mice, rats, and rabbits. Hence, in vivo imaging and ex vivo histological analysis is performed to monitor tissue development and measure them at the molecular level [141, 142]. After in vitro approval, the fabricated scaffolds were implanted in orthotopic and ectopic models, and collagen deposition, matrix mineralization, and bone tissue remodeling were investigated [124].

As shown in Fig. 4, implantation of PEG-hydrogels-MNPs loaded with SVF (stromal vascular fraction) cells in nude mice indicated the formation of tissue with dense vascularization and high mineralization [143]. These results confirm that MNPs provide both matrix calcification and proliferation of endothelial cell in vitro and in vivo to form compact bone tissue in an implanted model. Furthermore, mesh of PCL polymer modified with Fe3O4 (15%) can stimulate vascularization and bone regeneration after implantation in segmental bone defects in mouse models [113]. The μCT imaging exposed that magnetic nanofibrous scaffolds and external SMF increased the formation of fresh homogeneous tissue in a rabbit within three months. The magnetic properties accelerate the remodeling of bone via the thorough absorption of the scaffold materials in damaged area [65].

Fig. 4
figure 4

Osteogenic differentiation and bone formation. a After 3 weeks of culture, gel sections' mineralization (calcium deposition) was assessed using Alizarin red staining. b The samples were subcutaneously implanted in vivo, then observed via micro-computed tomography imaging after 1 or 8 weeks after extraction. c In vivo μCT image of gels in 3D volumetric view, and sections with gradient density of the tissue generated after 1 and 8 weeks. d By setting the following thresholds on the image histogram, the pixel intensity was defined based on the intervals of tissue density values: 0 to17000 (low density, in gray), 17,000 to 27,000 (moderate density, in green), 27,000 to 39,000 (intermediate density, in blue), and 39,000 to 65,500 (high density, in red). μCT analysis was used to calculate the tissue volume of gels. Reprinted from Ref. [143]. Filippi, M., et al., Magnetic nanocomposite hydrogels and static magnetic field stimulate the osteoblastic and vasculogenic profile of adipose-derived cells. Biomaterials, 2019. 223: p. 119468, Copyright (2023), with permission from Elsevier

In another study, four weeks of monitoring with MRI imaging revealed that implantation of SPION-enriched gelatin sponges in the Sprague–Dawley rats (incisor sockets) induced new bone formation and preserved alveolar [144]. In addition, CT imaging showed that the biodegradable Fe3O4 NPs/PLLA scaffolds accelerated the bone healing process after 8 weeks of implanting in tibia of rabbits [145]. Indeed, Fe3O4/PLLA nanofibers promoted bone regeneration in a dose-dependent behavior of nanoparticles, such that bone formation improved with increasing Fe3O4 NPs dosage [146] (Fig. 5).

Fig. 5
figure 5

a Rabbit’s tibia was scraped and two implantation holes were prepared with a diameter of 4 mm. Next, neat PLLA was grafted into one defected site and PLLA/MNP was grafted into the other. b, c Micro-Computed Tomography (μCT) images of defect sites after 4 weeks and d, e 8 weeks. Histological images of defected sites grafted with: f blank; g PLLA; h 2% MNP/PLLA; and i 5% MNP/PLLA scaffolds after 8 weeks. Reprinted from Ref. [146]. Lai, W.-Y., et al., In vivo investigation into effectiveness of Fe3O4/PLLA nanofibers for bone tissue engineering applications. Polymers, 2018. 10(7): p. 804, open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license

In another study, Shuai et al. prepared magnetic scaffolds and implanted in rabbit model within radius defect (Fig. 6). The results demonstrated that these structures remarkably increased angiogenesis, fibrous formation, and the creation of new bone tissue after two months [116]. Overall, the magnetic microenvironment may serve as an appropriate and efficient substrate in BTE in the future.

Fig. 6
figure 6

ac The procedure involves the creation of a bone defect model. d Implantation of 7.5 MNPs scaffolds in rabbit radius defect site, e MNP-free scaffolds, and f defect site with no implantation. g Micro-CT images of bone defects treated with 7.5 MNPs scaffolds, 0 MNP scaffolds, and blank groups. h The bone volume fraction and i the bone mineral density in the 7.5 MNPs and 0 MNPs scaffold groups in 1 and 2 months after implantation (**P < 0.01 and ***P < 0.005). Reprinted from Ref. [116]. Shuai, C., et al., A magnetic micro-environment in scaffolds for stimulating bone regeneration. Materials & Design, 2020. 185: p. 108275, open access article distributed under the CC BY-NC-ND license

In general, MNPs and their nanocomposite scaffolds have several advantages over more conventional implants in bone tissue engineering (Table 1). One advantage is that they can be functionalized with ligands that can target specific cells or tissues [147]. This allows for precise control over their localization and function, which is important for promoting tissue regeneration and minimizing unwanted side effects. Another advantage is the ability to manipulate magnetic nanoparticles through external magnetic fields, providing a non-invasive method for stimulating cells and promoting tissue regeneration, which is particularly important for bone tissue engineering [148]. MNPs can also modulate signaling pathways involved in bone formation and remodeling, such as the Wnt/β-catenin pathway, by functionalizing them with Wnt ligands or inhibitors, it is possible to promote or inhibit osteogenic differentiation, respectively, and regulate bone formation and remodeling [149]. Despite these advantages, there are several challenges that must be addressed before MNPs can be translated into clinical use. One important challenge is ensuring the long-term safety and biocompatibility of these materials. Magnetic nanoparticles have the potential to accumulate in tissues and organs over time, which could lead to toxicity and other adverse effects [150]. Researchers must carefully evaluate the biocompatibility of magnetic nanoparticles and their degradation products, as well as their potential to induce inflammation or other immune responses. Another challenge is understanding the mechanisms by which magnetic nanoparticles affect cell behavior and tissue regeneration. While there is growing evidence that magnetic nanoparticles can modulate signaling pathways involved in bone formation and remodeling, the underlying mechanisms are not yet fully understood. Further research is needed to elucidate the molecular and cellular mechanisms by which magnetic nanoparticles exert their effects, as well as their potential to interact with other signaling pathways and affect tissue regeneration in vivo.

Conclusions

There are many studies on the application of MNPs alone or in the structure of nanocomposite scaffolds in TERM. A lot of these studies have also investigated and confirmed the physical and biological effects of MNPs in the microenvironment of stem cells. In particular, the increasing current advanced research in BTE and growing amounts of experimental data point to the high potential of MNPs in the repair and healing of bone injuries. According to the conducted studies, incorporation and homogeneous distribution of MNPs in scaffolds leads to biological effects on cell activity and plays a role in their fate toward osteogenic differentiation. In this regard, many researchers investigated the molecular mechanisms and signaling pathways related to the osteogenic differentiation of stem cells and related or helpful factors in bone healing, especially angiogenesis, in the presence of MNPs or scaffolds containing them. Finally, the activation of molecular mechanisms and signaling pathways related to bone regeneration, as well as the creation of angiogenic responses and the formation of blood vessels and capillary tubes have been confirmed in recent years. Therefore, according to the present review, scaffolds reinforced with MNPs have been very helpful and effective in the reconstruction and healing of damaged bones. It is expected that in the near future, great achievements will be made available in the field of bone reconstruction using magnetic nanocomposite products. To achieve this important goal in bone tissue engineering, biosafety issues in the field of magnetization strategies should be further investigated. Although so far in vivo studies of magnetic scaffolds have not shown any serious toxic and inflammatory effects, these studies need to be completely conducted in vivo and over a longer period of time.

Availability of data and materials

Not applicable.

Abbreviations

MNPs:

Magnetic nanoparticles

MSCs:

Mesenchymal stem cells

BTE:

Bone tissue engineering

EMFs:

Electromagnetic fields

SMFs:

Static magnetic fields

SPIONs:

Superparamagnetic iron oxide nanoparticles

MRI:

Magnetic resonance imaging

ECM:

Extracellular matrix

FDA:

Food and Drug Administration

MRS:

Magnetic responsive scaffolds

ALP:

Alkaline phosphatase

PLLA:

Poly (l-lactic acid)

CPC:

Calcium phosphate cement

hDPSCs:

Human dental pulp stem cells

HA:

Hydroxyapatite

CZF:

Cobalt-zinc ferrite

PCL:

Poly (ε-caprolactone)

h-BMSCs:

Human bone marrow mesenchymal stem cells

PVDF:

Poly (vinylidene fluoride)

PGA:

Poly (glycolic acid)

SVF:

Stromal vascular fraction

BMP2:

Bone morphogenetic protein 2

MAPK:

Classic mitogen-activated protein kinase

INZEB2 :

Long non-coding RNA

BV/TV:

Bone volume per tissue volume

TEM:

Transmission electron microscope

IHC:

Immunohistochemistry

Runx2:

Runt-related transcription factor 2

GSK3:

Glycogen synthase kinase-3

GO:

Graphene oxide

PEMF:

Pulsed electromagnetic field

PKA:

Protein kinase A

JNK:

C-Jun N-terminal kinase

VEGF:

Vascular endothelial growth factor

References

  1. Gordeladze JO, et al. Bone tissue engineering: state of the art, challenges, and prospects. Tissue Engineering for Artificial Organs: Regenerative Medicine, Smart Diagnostics and Personalized Medicine. Wiley, 2017: p. 525–51.

  2. Banimohamad-Shotorbani B, et al. The efficiency of PCL/HAp electrospun nanofibers in bone regeneration: a review. J Med Eng Technol. 2021;45(7):511–31.

    PubMed  Google Scholar 

  3. Lyons JG, et al. Nanostructured biomaterials for bone regeneration. Front Bioeng Biotechnol. 2020;8:922.

    PubMed  PubMed Central  Google Scholar 

  4. Del Bakhshayesh AR, et al. High efficiency biomimetic electrospun fibers for use in regenerative medicine and drug delivery: a review. Mater Chem Phys. 2022;125785.

  5. Xu H-Y, Gu N. Magnetic responsive scaffolds and magnetic fields in bone repair and regeneration. Front Mater Sci. 2014;8:20–31.

    Google Scholar 

  6. Augat P, Hollensteiner M, von Rüden C. The role of mechanical stimulation in the enhancement of bone healing. Injury. 2021;52:S78–83.

    PubMed  Google Scholar 

  7. Quittan M, et al. Clinical effectiveness of magnetic field therapy–a review of the literature. Acta Med Austriaca. 2000;27(3):61–8.

    CAS  PubMed  Google Scholar 

  8. Gujjalapudi M, et al. Effect of magnetic field on bone healing around endosseous implants–An in-vivo study. J Clin Diagnost Res: JCDR 2016;10(10):ZF01.

  9. Xiong F, et al. Cardioprotective activity of iron oxide nanoparticles. Sci Rep. 2015;5(1):1–8.

    CAS  Google Scholar 

  10. Henstock JR, et al. Remotely activated mechanotransduction via magnetic nanoparticles promotes mineralization synergistically with bone morphogenetic protein 2: applications for injectable cell therapy. Stem Cells Transl Med. 2014;3(11):1363–74.

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Wang Q, et al. Response of MAPK pathway to iron oxide nanoparticles in vitro treatment promotes osteogenic differentiation of hBMSCs. Biomaterials. 2016;86:11–20.

    CAS  PubMed  Google Scholar 

  12. Hu S, et al. Enhanced bone regeneration and visual monitoring via superparamagnetic iron oxide nanoparticle scaffold in rats. J Tissue Eng Regen Med. 2018;12(4):e2085–98.

    CAS  PubMed  Google Scholar 

  13. Zhao X, et al. Active scaffolds for on-demand drug and cell delivery. Proc Natl Acad Sci. 2011;108(1):67–72.

    CAS  PubMed  Google Scholar 

  14. Singh D, et al. Bench-to-bedside translation of magnetic nanoparticles. Nanomedicine. 2014;9(4):501–16.

    CAS  PubMed  Google Scholar 

  15. Delyagina E, et al. Magnetic targeting strategies in gene delivery. Nanomedicine. 2011;6(9):1593–604.

    CAS  PubMed  Google Scholar 

  16. Paun IA, et al. 3D biomimetic magnetic structures for static magnetic field stimulation of osteogenesis. Int J Mol Sci. 2018;19(2):495.

    PubMed  PubMed Central  Google Scholar 

  17. Zhang N, et al. Magnetic nanocomposite hydrogel for potential cartilage tissue engineering: synthesis, characterization, and cytocompatibility with bone marrow derived mesenchymal stem cells. ACS Appl Mater Interfaces. 2015;7(37):20987–98.

    CAS  PubMed  Google Scholar 

  18. Širc J, et al. Morphological characterization of nanofibers: methods and application in practice. J Nanomater. 2012;2012:1–14.

    Google Scholar 

  19. Sapountzi E, et al. Recent advances in electrospun nanofiber interfaces for biosensing devices. Sensors. 2017;17(8):1887.

    PubMed  PubMed Central  Google Scholar 

  20. Fan Q-L, et al. Solvent-free atom transfer radical polymerization for the preparation of poly (poly (ethyleneglycol) monomethacrylate)-grafted Fe3O4 nanoparticles: synthesis, characterization and cellular uptake. Biomaterials. 2007;28(36):5426–36.

    CAS  PubMed  Google Scholar 

  21. Wu J, Hong Y. Enhancing cell infiltration of electrospun fibrous scaffolds in tissue regeneration. Bioactive Mater. 2016;1(1):56–64.

    Google Scholar 

  22. Sadeghzadeh H, et al. PCL/Col I-based magnetic nanocomposite scaffold provides an osteoinductive environment for ADSCs in osteogenic cues-free media conditions. Stem Cell Res Ther. 2022;13(1):1–18.

    Google Scholar 

  23. Guo B, Ma PX. Synthetic biodegradable functional polymers for tissue engineering: a brief review. Sci China Chem. 2014;57(4):490–500.

    CAS  Google Scholar 

  24. Vacanti JP, Vacanti CA. The history and scope of tissue engineering. In: Principles of tissue engineering. Elsevier; 2014. p. 3–8.

    Google Scholar 

  25. Asadi N, et al. Nanocomposite electrospun scaffold based on polyurethane/polycaprolactone incorporating gold nanoparticles and soybean oil for tissue engineering applications. J Bionic Eng. 2023;1–11.

  26. Stevens MM. Biomaterials for bone tissue engineering. Mater Today. 2008;11(5):18–25.

    CAS  Google Scholar 

  27. Valizadeh N, et al. Enhanced osteogenic differentiation and mineralization of human dental pulp stem cells using Prunus amygdalus amara (bitter almond) incorporated nanofibrous scaffold. J Mech Behav Biomed Mater. 2023;142: 105790.

    CAS  PubMed  Google Scholar 

  28. Bharadwaz A, Jayasuriya AC. Recent trends in the application of widely used natural and synthetic polymer nanocomposites in bone tissue regeneration. Mater Sci Eng, C. 2020;110:110698.

    CAS  Google Scholar 

  29. Langer R, Tirrell DA. Designing materials for biology and medicine. Nature. 2004;428(6982):487–92.

    CAS  PubMed  Google Scholar 

  30. Rahmani Del Bakhshayesh A, et al. Recent advances in nano‐scaffolds for tissue engineering applications: toward natural therapeutics. Wiley Interdisciplinary Reviews: Nanomedicine and Nanobiotechnology, 2023;p. e1882.

  31. Kestell AE, DeLorey GT. Nanoparticles: properties, classification, characterization, and fabrication. Incorporated: Nova Science Publishers; 2009.

    Google Scholar 

  32. Jeevanandam J, et al. Review on nanoparticles and nanostructured materials: history, sources, toxicity and regulations. Beilstein J Nanotechnol. 2018;9(1):1050–74.

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Shang L, Nienhaus K, Nienhaus GU. Engineered nanoparticles interacting with cells: size matters. J Nanobiotechnol. 2014;12(1):1–11.

    Google Scholar 

  34. Pisane KL, Despeaux EC, Seehra MS. Magnetic relaxation and correlating effective magnetic moment with particle size distribution in maghemite nanoparticles. J Magn Magn Mater. 2015;384:148–54.

    CAS  Google Scholar 

  35. Nikzamir M, Akbarzadeh A, Panahi Y. An overview on nanoparticles used in biomedicine and their cytotoxicity. Jf Drug Deliv Sci Technol. 2021;61:102316.

    CAS  Google Scholar 

  36. Altavilla C, Ciliberto E. Inorganic nanoparticles: synthesis, applications, and perspectives. Boca Raton: CRC Press; 2017.

    Google Scholar 

  37. Siddiqi KS, ur Rahman A, Husen A. Biogenic fabrication of iron/iron oxide nanoparticles and their application. Nanoscale Res Lett. 2016;11:1–13.

    Google Scholar 

  38. Samrot AV, et al. A review on synthesis, characterization and potential biological applications of superparamagnetic iron oxide nanoparticles. Curr Res Green Sustain Chem. 2021;4:100042.

    CAS  Google Scholar 

  39. Sadeghzadeh H, et al. The effects of nanoencapsulated curcumin-Fe3O4 on proliferation and hTERT gene expression in lung cancer cells. Anti-Cancer Agents in Medicinal Chemistry (Formerly Current Medicinal Chemistry-Anti-Cancer Agents), 2017;17(10):1363–1373.

  40. Shi D, et al. Photo-fluorescent and magnetic properties of iron oxide nanoparticles for biomedical applications. Nanoscale. 2015;7(18):8209–32.

    CAS  PubMed  Google Scholar 

  41. Dong Q, et al. A molecular approach to magnetic metallic nanostructures from metallopolymer precursors. Chem Soc Rev. 2018;47(13):4934–53.

    CAS  PubMed  Google Scholar 

  42. Yang L, Liu H. Stimuli-responsive magnetic particles and their applications in biomedical field. Powder Technol. 2013;240:54–65.

    CAS  Google Scholar 

  43. Kalia S, et al. Magnetic polymer nanocomposites for environmental and biomedical applications. Colloid Polym Sci. 2014;292(9):2025–52.

    CAS  Google Scholar 

  44. Li X, et al. Current investigations into magnetic nanoparticles for biomedical applications. J Biomed Mater Res A. 2016;104(5):1285–96.

    CAS  PubMed  Google Scholar 

  45. Gupta AK, Gupta M. Synthesis and surface engineering of iron oxide nanoparticles for biomedical applications. Biomaterials. 2005;26(18):3995–4021.

    CAS  PubMed  Google Scholar 

  46. Jia Y, et al. Co-encapsulation of magnetic Fe3O4 nanoparticles and doxorubicin into biodegradable PLGA nanocarriers for intratumoral drug delivery. Int J Nanomed. 2012;7:1697.

    CAS  Google Scholar 

  47. Bonfield W, et al. Hydroxyapatite reinforced polyethylene–a mechanically compatible implant material for bone replacement. Biomaterials. 1981;2(3):185–6.

    CAS  PubMed  Google Scholar 

  48. Pina S, Oliveira JM, Reis RL. Natural-based nanocomposites for bone tissue engineering and regenerative medicine: a review. Adv Mater. 2015;27(7):1143–69.

    CAS  PubMed  Google Scholar 

  49. Oliveira JM, et al. Macroporous hydroxyapatite scaffolds for bone tissue engineering applications: physicochemical characterization and assessment of rat bone marrow stromal cell viability. J Biomed Mater Res A: Off J Soc Biomater Jpn Soc Biomater Austral Soc Biomater Korean Soc Biomater. 2009;91(1):175–86.

    Google Scholar 

  50. Hou Q, Grijpma DW, Feijen J. Porous polymeric structures for tissue engineering prepared by a coagulation, compression moulding and salt leaching technique. Biomaterials. 2003;24(11):1937–47.

    CAS  PubMed  Google Scholar 

  51. Liapis A, Pim M, Bruttini R. Research and development needs and opportunities in freeze drying. Drying Technol. 1996;14(6):1265–300.

    CAS  Google Scholar 

  52. van de Witte P, et al. Phase separation processes in polymer solutions in relation to membrane formation. J Membr Sci. 1996;117(1–2):1–31.

    Google Scholar 

  53. Dehghani F, Annabi N. Engineering porous scaffolds using gas-based techniques. Curr Opin Biotechnol. 2011;22(5):661–6.

    CAS  PubMed  Google Scholar 

  54. Abdelaal OA, Darwish SM. Fabrication of tissue engineering scaffolds using rapid prototyping techniques. Int J Indus Manuf Eng. 2011;5(11):2310–8.

    Google Scholar 

  55. Chae T, et al. Novel biomimetic hydroxyapatite/alginate nanocomposite fibrous scaffolds for bone tissue regeneration. J Mater Sci Mater Med. 2013;24(8):1885–94.

    CAS  PubMed  Google Scholar 

  56. Russo A, et al. Bone regeneration in a rabbit critical femoral defect by means of magnetic hydroxyapatite macroporous scaffolds. J Biomed Mater Res B Appl Biomater. 2018;106(2):546–54.

    CAS  PubMed  Google Scholar 

  57. Huang J, et al. Preparation and biocompatibility of diphasic magnetic nanocomposite scaffold. Mater Sci Eng C. 2018;87:70–7.

    CAS  Google Scholar 

  58. Zhao Y, et al. Magnetic bioinspired micro/nanostructured composite scaffold for bone regeneration. Colloids Surf B. 2019;174:70–9.

    CAS  Google Scholar 

  59. Xia Y, et al. Injectable calcium phosphate scaffold with iron oxide nanoparticles to enhance osteogenesis via dental pulp stem cells. Artif Cells Nanomed Biotechnol. 2018;46(sup1):423–33.

    CAS  PubMed  Google Scholar 

  60. Xia Y, Chen H, Zhang F. Injectable calcium phosphate scaffold with iron oxide nanoparticles to enhance osteogenesis via dental pulp stem cells. Artif Cells Nanomed Biotechnol. 2018;46:423–33.

    CAS  PubMed  Google Scholar 

  61. Li Y, et al. Adaptive materials based on iron oxide nanoparticles for bone regeneration. ChemPhysChem. 2018;19(16):1965–79.

    CAS  PubMed  Google Scholar 

  62. Hou R, et al. Magnetic nanohydroxyapatite/PVA composite hydrogels for promoted osteoblast adhesion and proliferation. Colloids Surf, B. 2013;103:318–25.

    CAS  Google Scholar 

  63. Ito A, et al. Tissue engineering using magnetite nanoparticles and magnetic force: heterotypic layers of cocultured hepatocytes and endothelial cells. Tissue Eng. 2004;10(5–6):833–40.

    CAS  PubMed  Google Scholar 

  64. Wu Y, et al. A novel calcium phosphate ceramic–magnetic nanoparticle composite as a potential bone substitute. Biomed Mater. 2010;5(1): 015001.

    Google Scholar 

  65. Meng J, et al. Paramagnetic nanofibrous composite films enhance the osteogenic responses of pre-osteoblast cells. Nanoscale. 2010;2(12):2565–9.

    CAS  PubMed  Google Scholar 

  66. Shan D, et al. Electrospun magnetic poly (l-lactide)(PLLA) nanofibers by incorporating PLLA-stabilized Fe3O4 nanoparticles. Mater Sci Eng, C. 2013;33(6):3498–505.

    CAS  Google Scholar 

  67. Wang H-T, et al. Development and biocompatibility tests of electrospun poly-l-lactide nanofibrous membranes incorporating oleic acid-coated Fe3O4. J Polym Eng. 2014;34(3):241–5.

    CAS  Google Scholar 

  68. Huang H-M, et al. Static magnetic fields up-regulate osteoblast maturity by affecting local differentiation factors. Clin Orthopaedics Relat Res®. 2006;447:201–8.

  69. Wójcik-Piotrowicz K, et al. Cell viability modulation through changes of Ca2+-dependent signalling pathways. Prog Biophys Mol Biol. 2016;121(1):45–53.

    PubMed  Google Scholar 

  70. Zhu Y, et al. Protein corona of magnetic hydroxyapatite scaffold improves cell proliferation via activation of mitogen-activated protein kinase signaling pathway. ACS Nano. 2017;11(4):3690–704.

    CAS  PubMed  Google Scholar 

  71. Wang J, et al. The effects of pulsed electromagnetic field on the functions of osteoblasts on implant surfaces with different topographies. Acta Biomater. 2014;10(2):975–85.

    CAS  PubMed  Google Scholar 

  72. Tampieri A, et al. Mimicking natural bio-mineralization processes: a new tool for osteochondral scaffold development. Trends Biotechnol. 2011;29(10):526–35.

    CAS  PubMed  Google Scholar 

  73. Arjmand M, et al. Osteogenic differentiation potential of mesenchymal stem cells cultured on nanofibrous scaffold improved in the presence of pulsed electromagnetic field. J Cell Physiol. 2018;233(2):1061–70.

    CAS  PubMed  Google Scholar 

  74. Panseri S, et al. Innovative magnetic scaffolds for orthopedic tissue engineering. J Biomed Mater Res A. 2012;100(9):2278–86.

    CAS  PubMed  Google Scholar 

  75. Cunha C, et al. Evaluation of the effects of a moderate intensity static magnetic field application on human osteoblast-like cells. Am J Biomed Eng. 2012;2:263–8.

    Google Scholar 

  76. Lai K, et al. Superparamagnetic nano-composite scaffolds for promoting bone cell proliferation and defect reparation without a magnetic field. RSC Adv. 2012;2(33):13007–17.

    CAS  Google Scholar 

  77. Hai-Yan X. NG Magnetic responsive scaffolds and magnetic fields in bone repair and regeneration. Front Mater Sci. 2014;8:20–31.

    Google Scholar 

  78. Yun H-M, et al. Magnetic nanocomposite scaffolds combined with static magnetic field in the stimulation of osteoblastic differentiation and bone formation. Biomaterials. 2016;85:88–98.

    CAS  PubMed  Google Scholar 

  79. Santos LJ, Reis RL, Gomes ME. Harnessing magnetic-mechano actuation in regenerative medicine and tissue engineering. Trends Biotechnol. 2015;33(8):471–9.

    CAS  PubMed  Google Scholar 

  80. Shubayev VI, Pisanic TR II, Jin S. Magnetic nanoparticles for theragnostics. Adv Drug Deliv Rev. 2009;61(6):467–77.

    CAS  PubMed  PubMed Central  Google Scholar 

  81. Mancarella S, et al. Polymer-coated magnetic nanoparticles for curcumin delivery to cancer cells. Macromol Biosci. 2015;15(10):1365–74.

    CAS  PubMed  Google Scholar 

  82. Osaci M, Cacciola M. Influence of the magnetic nanoparticle coating on the magnetic relaxation time. Beilstein J Nanotechnol. 2020;11(1):1207–16.

    CAS  PubMed  PubMed Central  Google Scholar 

  83. Liu Y, et al. Facile surface functionalization of hydrophobic magnetic nanoparticles. J Am Chem Soc. 2014;136(36):12552–5.

    CAS  PubMed  PubMed Central  Google Scholar 

  84. Sensenig R, et al. Magnetic nanoparticle-based approaches to locally target therapy and enhance tissue regeneration in vivo. Nanomedicine. 2012;7(9):1425–42.

    CAS  PubMed  Google Scholar 

  85. Cao J, et al. Notch signaling pathway promotes osteogenic differentiation of mesenchymal stem cells by enhancing BMP9/Smad signaling. Int J Mol Med. 2017;40(2):378–88.

    CAS  PubMed  PubMed Central  Google Scholar 

  86. Liu W, et al. Canonical Wnt signaling differently modulates osteogenic differentiation of mesenchymal stem cells derived from bone marrow and from periodontal ligament under inflammatory conditions. Biochimica et Biophysica Acta (BBA)-General Subjects 2014; 1840(3):1125–34.

  87. Salasznyk RM, et al. Focal adhesion kinase signaling pathways regulate the osteogenic differentiation of human mesenchymal stem cells. Exp Cell Res. 2007;313(1):22–37.

    CAS  PubMed  Google Scholar 

  88. Wang H, et al. Enhanced osteogenesis of bone marrow stem cells cultured on hydroxyapatite/collagen I scaffold in the presence of low-frequency magnetic field. J Mater Sci Mater Med. 2019;30:1–12.

    Google Scholar 

  89. Ross CL, et al. The effect of low-frequency electromagnetic field on human bone marrow stem/progenitor cell differentiation. Stem Cell Res. 2015;15(1):96–108.

    PubMed  PubMed Central  Google Scholar 

  90. Wang N, et al. Functions and applications of metallic and metallic oxide nanoparticles in orthopedic implants and scaffolds. J Biomed Mater Res B Appl Biomater. 2021;109(2):160–79.

    CAS  PubMed  Google Scholar 

  91. Chen G, et al. 3D scaffolds with different stiffness but the same microstructure for bone tissue engineering. ACS Appl Mater Interfaces. 2015;7(29):15790–802.

    CAS  PubMed  Google Scholar 

  92. Liu C, et al. Matrix stiffness regulates the differentiation of tendon-derived stem cells through FAK-ERK1/2 activation. Exp Cell Res. 2018;373(1–2):62–70.

    CAS  PubMed  Google Scholar 

  93. Lambertini E, et al. Osteogenic differentiation of human MSCs: Specific occupancy of the mitochondrial DNA by NFATc1 transcription factor. Int J Biochem Cell Biol. 2015;64:212–9.

    CAS  PubMed  Google Scholar 

  94. Hwang J-H, et al. Extracellular matrix stiffness regulates osteogenic differentiation through MAPK activation. PLoS ONE. 2015;10(8): e0135519.

    PubMed  PubMed Central  Google Scholar 

  95. Sun M, et al. Extracellular matrix stiffness controls osteogenic differentiation of mesenchymal stem cells mediated by integrin α5. Stem Cell Res Ther. 2018;9(1):1–13.

    Google Scholar 

  96. Lopes HB, et al. Participation of integrin β3 in osteoblast differentiation induced by titanium with nano or microtopography. J Biomed Mater Res A. 2019;107(6):1303–13.

    CAS  PubMed  PubMed Central  Google Scholar 

  97. Gao X, et al. A comparison of bone regeneration with human mesenchymal stem cells and muscle-derived stem cells and the critical role of BMP. Biomaterials. 2014;35(25):6859–70.

    CAS  PubMed  PubMed Central  Google Scholar 

  98. Chen G, Yang L, Lv Y. Cell-free scaffolds with different stiffness but same microstructure promote bone regeneration in rabbit large bone defect model. J Biomed Mater Res A. 2016;104(4):833–41.

    CAS  PubMed  Google Scholar 

  99. Aiyelabegan HT, Sadroddiny E. Fundamentals of protein and cell interactions in biomaterials. Biomed Pharmacother. 2017;88:956–70.

    CAS  PubMed  Google Scholar 

  100. Xie J, et al. Substrate elasticity regulates adipose-derived stromal cell differentiation towards osteogenesis and adipogenesis through β-catenin transduction. Acta Biomater. 2018;79:83–95.

    CAS  PubMed  Google Scholar 

  101. He Y, et al. Enhanced osteogenic differentiation of human bone–derived mesenchymal stem cells in 3-dimensional printed porous titanium scaffolds by static magnetic field through up-regulating Smad4. FASEB J. 2019;33(5):6069–81.

    CAS  PubMed  Google Scholar 

  102. Wang Q, et al. Magnetic iron oxide nanoparticles accelerate osteogenic differentiation of mesenchymal stem cells via modulation of long noncoding RNA INZEB2. Nano Res. 2017;10(2):626–42.

    Google Scholar 

  103. Yong Y, et al. Electromagnetic fields promote osteogenesis of rat mesenchymal stem cells through the PKA and ERK1/2 pathways. J Tissue Eng Regen Med. 2016;10(10):E537–45.

    CAS  PubMed  Google Scholar 

  104. Wu S, et al. Pulsed electromagnetic field induces Ca2+-dependent osteoblastogenesis in C3H10T1/2 mesenchymal cells through the Wnt-Ca2+/Wnt-β-catenin signaling pathway. Biochem Biophys Res Commun. 2018;503(2):715–21.

    CAS  PubMed  Google Scholar 

  105. Kim E-C, et al. Static magnetic fields promote osteoblastic/cementoblastic differentiation in osteoblasts, cementoblasts, and periodontal ligament cells. J Periodontal Implant Sci. 2017;47(5):273–91.

    CAS  PubMed  PubMed Central  Google Scholar 

  106. Hashimoto Y, Kawasumi M, Saito M. Effect of static magnetic field on cell migration. Electr Eng Jpn. 2007;160(2):46–52.

    Google Scholar 

  107. Miyakoshi J. Effects of static magnetic fields at the cellular level. Prog Biophys Mol Biol. 2005;87(2–3):213–23.

    CAS  PubMed  Google Scholar 

  108. Diomede F, et al. Functional relationship between osteogenesis and angiogenesis in tissue regeneration. Int J Mol Sci. 2020;21(9):3242.

    CAS  PubMed  PubMed Central  Google Scholar 

  109. Izadpanah M, et al. Melatonin and endothelial cell-loaded alginate-fibrin hydrogel promoted angiogenesis in rat cryopreserved/thawed ovaries transplanted to the heterotopic sites. J Biol Eng. 2023;17(1):1–13.

    Google Scholar 

  110. Sapir Y, et al. Magnetically actuated alginate scaffold: a novel platform for promoting tissue organization and vascularization. In: Cardiac Tissue Engineering. Springer; 2014. p. 83–95.

    Google Scholar 

  111. Sapir Y, et al. The promotion of in vitro vessel-like organization of endothelial cells in magnetically responsive alginate scaffolds. Biomaterials. 2012;33(16):4100–9.

    CAS  PubMed  PubMed Central  Google Scholar 

  112. Daňková J, et al. Highly efficient mesenchymal stem cell proliferation on poly-ε-caprolactone nanofibers with embedded magnetic nanoparticles. Int J Nanomed. 2015;10:7307.

    Google Scholar 

  113. De Santis R, et al. Towards the design of 3D fiber-deposited poly (-caprolactone)/iron-doped hydroxyapatite nanocomposite magnetic scaffolds for bone regeneration. J Biomed Nanotechnol. 2015;11(7):1236–46.

    PubMed  Google Scholar 

  114. Ganesh N, et al. Magnetic resonance functional nano-hydroxyapatite incorporated poly (caprolactone) composite scaffolds for in situ monitoring of bone tissue regeneration by MRI. Tissue Eng A. 2014;20(19–20):2783–94.

    CAS  Google Scholar 

  115. Chen H, et al. Magnetic cell–scaffold interface constructed by superparamagnetic IONP enhanced osteogenesis of adipose-derived stem cells. ACS Appl Mater Interfaces. 2018;10(51):44279–89.

    CAS  PubMed  Google Scholar 

  116. Shuai C, et al. A magnetic micro-environment in scaffolds for stimulating bone regeneration. Mater Des. 2020;185:108275.

    CAS  Google Scholar 

  117. Zhang J, et al. 3D-printed magnetic Fe 3 O 4/MBG/PCL composite scaffolds with multifunctionality of bone regeneration, local anticancer drug delivery and hyperthermia. Journal of materials chemistry B. 2014;2(43):7583–95.

    CAS  PubMed  Google Scholar 

  118. Perez RA, Patel KD, Kim H-W. Novel magnetic nanocomposite injectables: calcium phosphate cements impregnated with ultrafine magnetic nanoparticles for bone regeneration. RSC Adv. 2015;5(18):13411–9.

    CAS  Google Scholar 

  119. Xia Y, et al. Novel magnetic calcium phosphate-stem cell construct with magnetic field enhances osteogenic differentiation and bone tissue engineering. Mater Sci Eng, C. 2019;98:30–41.

    CAS  Google Scholar 

  120. Lu J-W, et al. Magnetic nanoparticles modified-porous scaffolds for bone regeneration and photothermal therapy against tumors. Nanomed: Nanotechnol Biol Med. 2018;14(3):811–22.

    CAS  Google Scholar 

  121. Hou C, et al. Electromagnetic and mechanical properties of Fe3O4-coated amorphous carbon nanotube/polyvinyl chloride composites. Compos Interfaces. 2016;23(9):901–7.

    CAS  Google Scholar 

  122. Pan YH, et al. Fabrication of Fe3O4/PLLA composites for use in bone tissue engineering. Polym Compos. 2017;38(12):2881–8.

    CAS  Google Scholar 

  123. D’Amora U, et al. 3D additive-manufactured nanocomposite magnetic scaffolds: effect of the application mode of a time-dependent magnetic field on hMSCs behavior. Bioactive Mater. 2017;2(3):138–45.

    Google Scholar 

  124. Xia Y, et al. Magnetic field and nano-scaffolds with stem cells to enhance bone regeneration. Biomaterials. 2018;183:151–70.

    CAS  PubMed  Google Scholar 

  125. Torgbo S, Sukyai P. Fabrication of microporous bacterial cellulose embedded with magnetite and hydroxyapatite nanocomposite scaffold for bone tissue engineering. Mater Chem Phys. 2019;237:121868.

    CAS  Google Scholar 

  126. Li J, et al. Fabrication and biocompatibility of core–shell structured magnetic fibrous scaffold. J Biomed Nanotechnol. 2019;15(3):500–6.

    CAS  PubMed  Google Scholar 

  127. Aliramaji S, Zamanian A, Mozafari M. Super-paramagnetic responsive silk fibroin/chitosan/magnetite scaffolds with tunable pore structures for bone tissue engineering applications. Mater Sci Eng C. 2017;70:736–44.

    CAS  Google Scholar 

  128. Hughes S, El Haj AJ, Dobson J. Magnetic micro-and nanoparticle mediated activation of mechanosensitive ion channels. Med Eng Phys. 2005;27(9):754–62.

    PubMed  Google Scholar 

  129. Huang D-M, et al. The promotion of human mesenchymal stem cell proliferation by superparamagnetic iron oxide nanoparticles. Biomaterials. 2009;30(22):3645–51.

    CAS  PubMed  Google Scholar 

  130. Singh RK, et al. Potential of magnetic nanofiber scaffolds with mechanical and biological properties applicable for bone regeneration. PLoS ONE. 2014;9(4):e91584.

    PubMed  PubMed Central  Google Scholar 

  131. Moradian E, et al. Fabrication and physicochemical characterization of a novel magnetic nanocomposite scaffold: electromagnetic field effect on biological properties. Mater Sci Eng C. 2020;116:111222.

    CAS  Google Scholar 

  132. Mushtaq A, et al. Magnetic hydroxyapatite nanocomposites: the advances from synthesis to biomedical applications. Mater Des. 2021;197:109269.

    CAS  Google Scholar 

  133. Mosig AS. Organ-on-chip models: new opportunities for biomedical research. Future Sci. 2016;FSO130.

  134. Grainger DW. Cell-based drug testing; this world is not flat. Adv Drug Deliv Rev. 2014;69:vii–xi.

    PubMed  Google Scholar 

  135. Zimmerling A, et al. 3D printing PCL/nHA bone scaffolds: Exploring the influence of material synthesis techniques. Biomater Res. 2021;25(1):1–12.

    Google Scholar 

  136. Mattei G, Vozzi G. CFD modelling of a mixing chamber for the realisation of functionally graded scaffolds. Comput Chem Eng. 2016;84:43–8.

    CAS  Google Scholar 

  137. Abpeikar Z, et al. Influence of pore sizes in 3D-scaffolds on mechanical properties of scaffolds and survival, distribution, and proliferation of human chondrocytes. Mech Adv Mater Struct. 2022;29(26):4911–22.

    CAS  Google Scholar 

  138. Oftadeh R, et al. Biomechanics and mechanobiology of trabecular bone: a review. J Biomech Eng. 2015;137(1).

  139. McHale MK, Bergmann NM, West JL. Histogenesis in three-dimensional scaffolds. In: Principles of Regenerative Medicine. Elsevier; 2019. p. 661–74.

    Google Scholar 

  140. Brito-Pereira R, et al. Silk fibroin-magnetic hybrid composite electrospun fibers for tissue engineering applications. Compos B Eng. 2018;141:70–5.

    CAS  Google Scholar 

  141. Hao L, et al. Synergistic osteogenesis promoted by magnetically actuated nano-mechanical stimuli. Nanoscale. 2019;11(48):23423–37.

    CAS  PubMed  Google Scholar 

  142. Liu Z, et al. A magnetically responsive nanocomposite scaffold combined with Schwann cells promotes sciatic nerve regeneration upon exposure to magnetic field. Int J Nanomed. 2017;12:7815.

    CAS  Google Scholar 

  143. Filippi M, et al. Magnetic nanocomposite hydrogels and static magnetic field stimulate the osteoblastic and vasculogenic profile of adipose-derived cells. Biomaterials. 2019;223:119468.

    CAS  PubMed  Google Scholar 

  144. Yang C, et al. Non-invasive monitoring of in vivo bone regeneration based on alkaline phosphatase-responsive scaffolds. Chem Eng J. 2021;408:127959.

    CAS  Google Scholar 

  145. Chang W-J, et al. Development and testing of X-ray imaging-enhanced poly-l-lactide bone screws. PLoS ONE. 2015;10(10):e0140354.

    PubMed  PubMed Central  Google Scholar 

  146. Lai W-Y, et al. In vivo investigation into effectiveness of Fe3O4/PLLA nanofibers for bone tissue engineering applications. Polymers. 2018;10(7):804.

    PubMed  PubMed Central  Google Scholar 

  147. Shabatina TI, et al. Magnetic nanoparticles for biomedical purposes: modern trends and prospects. Magnetochemistry. 2020;6(3):30.

    CAS  Google Scholar 

  148. Liu XL, et al. Magnetic nanomaterials for advanced regenerative medicine: the promise and challenges. Adv Mater. 2019;31(45):1804922.

    CAS  Google Scholar 

  149. Dasari A, Xue J, Deb S. Magnetic nanoparticles in bone tissue engineering. Nanomaterials. 2022;12(5):757.

    CAS  PubMed  PubMed Central  Google Scholar 

  150. Van de Walle A, et al. Magnetic nanoparticles in regenerative medicine: what of their fate and impact in stem cells? Mater Today Nano. 2020;11:100084.

    Google Scholar 

Download references

Acknowledgements

The authors thank the Department of Tissue Engineering, Faculty of Advanced Medical Science of Tabriz University for all supports provided.

Funding

This is a report of the database from a Ph.D. thesis registered in Tabriz University of Medical Sciences with the Number 64724 (Ethical code: IR.TBZMED.VCR.REC.1399.042).

Author information

Authors and Affiliations

Authors

Contributions

HS contributed to the investigation, visualization, and writing original draft, figure preparation, and project management. HD-M contributed to the research, writing, and editing of the initial draft. AR-D contributed to studying and writing original draft. DM contributed investigation and writing—review. AM contributed to writing the initial draft, figures designing, editing, figures preparation, supervision, and project administration. All authors participated in the finalization of the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Ahmad Mehdipour.

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.

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

Sadeghzadeh, H., Dianat-Moghadam, H., Del Bakhshayesh, A.R. et al. A review on the effect of nanocomposite scaffolds reinforced with magnetic nanoparticles in osteogenesis and healing of bone injuries. Stem Cell Res Ther 14, 194 (2023). https://doi.org/10.1186/s13287-023-03426-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13287-023-03426-0

Keywords