Skip to main content

Emerging nano-scale delivery systems for the treatment of osteoporosis

Abstract

Osteoporosis is a pathological condition characterized by an accelerated bone resorption rate, resulting in decreased bone density and increased susceptibility to fractures, particularly among the elderly population. While conventional treatments for osteoporosis have shown efficacy, they are associated with certain limitations, including limited drug bioavailability, non-specific administration, and the occurrence of adverse effects. In recent years, nanoparticle-based drug delivery systems have emerged as a promising approach for managing osteoporosis. Nanoparticles possess unique physicochemical properties, such as a small size, large surface area-to-volume ratio, and tunable surface characteristics, which enable them to overcome the limitations of conventional therapies. These nanoparticles offer several advantages, including enhanced drug stability, controlled release kinetics, targeted bone tissue delivery, and improved drug bioavailability. This comprehensive review aims to provide insights into the recent advancements in nanoparticle-based therapy for osteoporosis. It elucidates the various types of nanoparticles employed in this context, including silica, polymeric, solid lipid, and metallic nanoparticles, along with their specific processing techniques and inherent properties that render them suitable as potential drug carriers for osteoporosis treatment. Furthermore, this review discusses the challenges and future suggestions associated with the development and translation of nanoparticle drug delivery systems for clinical use. These challenges encompass issues such as scalability, safety assessment, and regulatory considerations. However, despite these challenges, the utilization of nanoparticle-based drug delivery systems holds immense promise in revolutionizing the field of osteoporosis management by enabling more effective and targeted therapies, ultimately leading to improved patient outcomes.

Introduction

Bone, being a metabolically dynamic tissue, serves as the foundational framework of the human skeletal system. It encompasses three fundamental functions of paramount significance. Firstly, bone acts as a reservoir for diverse elements, including magnesium, phosphate, and bicarbonate, thus playing a pivotal role in the maintenance of calcium homeostasis [1]. Secondly, it assumes a critical role in providing protection against internal injuries. Lastly, bone is essential for adult hematopoiesis, the process of blood cell formation [2]. To accomplish these functions, bone tissue must uphold a delicate equilibrium between its resorption and formation, known as bone remodeling [3].

Disruption of the intricate bone remodeling processes can give rise to a spectrum of disorders, leading to detrimental effects on the skeletal system and resulting in compromised mobility and potentially life-threatening consequences. Notable skeletal disorders include osteosarcoma, non-union bone defects, osteoarthritis, osteoporosis (OP), and bone metastatic cancers [4, 5]. Among these, OP is the most prevalent metabolic bone disease that primarily affects the elderly and postmenopausal women [6, 7]. OP is described by the World Health Organization as a “progressive systemic skeletal disease characterized by low bone mass and microarchitectural deterioration of the bone tissue, resulting in increased bone fragility and susceptibility to fracture” [8]. OP predominantly affects vulnerable regions of the skeleton, leading to decreased bone density [9, 10]. Moreover, OP induces structural weakening of the vertebrae, hip joint, and carpal bones due to aberrant remodeling of trabecular and cortical bones [6, 11, 12].

Multiple research endeavors have been dedicated to developing treatment options for OP. However, striking a delicate balance between drug efficacies and minimizing adverse effects remains a formidable challenge. There exists an imperative to explore innovative approaches that enable the safe and efficient delivery of drugs to the bone tissue, thereby enhancing drug efficacy while mitigating side effects [13, 14]. To address the immediate needs of OP patients, an array of interventions, including amputation surgery, chemotherapy, radiation therapy, and drug injection, have been employed [13]. While effective in managing OP, these methods are not suitable for long-term and localized therapies, and they are associated with a range of side effects. These include broad tissue distribution [15], suboptimal targeting efficiency leading to off-target effects [16], limited drug half-life, inadequate bioavailability [14], and inadequate availability of bone graft sources [17]. Furthermore, these approaches pose risks of infection and uncontrolled drug release [17, 18]. To mitigate these side effects, drug delivery systems (DDSs) based on biocompatible materials have been developed. Table 1 provides an overview of materials employed in DDSs for OP treatment. Although these systems have demonstrated enhanced therapeutic efficacy in addressing OP, further research is required to overcome the challenges elucidated in Table 1.

Table 1 Pros and cons of various drugs used in the treatment of osteoporosis. Conventional drug delivery systems for osteoporosis have limitations including low selectivity, poor bioavailability, short half-life, potential side effects, invasive administration routes, lack of targeted delivery, and limited control over drug release. Addressing these limitations is a current area of research for improved therapeutic approaches

Nanomedicine, an expeditiously progressing discipline within the realm of material science, presents a plethora of advantages such as drug side effect reduction and biomimicking capabilities [19,20,21]. The exploration of nanomedicine has been underway since the late 1970s, and after the late 1980s, three pivotal strategies have emerged as primary drivers in the domain of nano-based drug delivery systems (DDSs): (1) PEGylation, (2) active targeting, and (3) the enhanced permeation and retention (EPR) effect [22]. PEGylation, a technique involving the conjugation of poly(ethylene glycol) to drugs, was initially investigated in the late 1960s to enhance drug stability and prolong circulation time within the organism [23]. Active targeting, which enables the attachment of targeting molecules such as ligands and antibodies to drugs, has been feasible since the late 1950s [24]. The discovery of the EPR effect in 1984 by Hiroshi Maeda of Kumamoto University facilitated the formation of polymer-conjugated DDSs through diverse biomaterial conjugations, thereby harnessing the phenomenon [25]. These three strategies have significantly propelled the advancement of research in nano-based DDSs, delivering substantial advantages in the field of drug delivery. However, further investigation is indispensable to effectively address the challenges associated with these systems.

Nanoparticles (NPs) are being used as an alternative method for bone-targeted treatment. Because of their small size and similarity to the components found in tissues [26], the nanoparticle materials can be delivered to specific tissues, organelles, or cells where the medicine will be released [27]. The advantages of NPs 1are that they provide a large capacity of the drug concerning size [28, 29], improve solubility [30], provide drug stability [31], reduce adverse effects [32], and improve transport for drug internalization in specific organelles [33]. Another advantage is that the NPs, which are composed of calcium phosphate, gold, and nanodiamonds, can help to activate functions in the cells, improve mineralization, and stimulate bone growth [32]. Thus, nanotechnology can overcome the limitations of conventional bone therapy, such as adverse effects and poor penetration to skeletal lesions.

The employment of nano-based DDSs offers the potential to enhance drug solubility and stability within the human body by conferring a biocompatible protective shield [34,35,36,37]. Moreover, the drug-loading capacity can be meticulously controlled through size adjustments of the DDSs [38, 39]. These DDSs possess inherent characteristics that facilitate the regulation of drug-release rates [40,41,42,43]. Recent investigations have explored diverse nanomaterials in the pursuit of advancing drug delivery, and in this review, we aim to present a concise overview of current research endeavors pertaining to nano-based DDSs.

Nano-based approaches for the treatment of OP

Osteoporosis is a systemic skeletal disorder that exhibits diminished bone strength and an elevated susceptibility to fractures during routine activities. This condition is characterized by a reduction in bone mineral density (BMD) and deterioration of the bone microarchitecture, leading to compromised structural integrity. OP not only undermines the physical well-being of individuals but also significantly impacts their overall quality of life [59]. In a state of normal health, the dynamic balance between OCs and OBs ensures efficient bone remodeling, facilitating the repair of microdamage caused by routine activities. OCs are responsible for the breakdown of bone tissue, while OBs play a vital role in its formation. However, in the case of OP, there is an imbalance in this process, favoring OC activity over OB function. Consequently, the trabecular region, in particular, experiences a loss of bone mass. This reduction in trabecular connectivity renders the bone more susceptible to brittleness and significantly heightens the risk of fractures [60].

Considering the complexities associated with OP, alternative strategies involving biomaterials, particularly nanomaterials, have been explored to address this condition and promote bone regeneration. In this context, nano-based DDSs hold promise as a potential solution. Therefore, to improve the limitations posed by the conventional DDSs, nano-based DDSs could offer a potential solution (Fig. 1).

Fig. 1
figure 1

A schematic illustration depicting the disadvantages and advantages between conventional treatment methods and nanoscale-based treatment in addressing OP. Drugs like bisphosphates and hormones have been used to treat osteoporosis with treatment approaches like oral and intranasal delivery; however, these exhibit numerous drawbacks like long-term side effects, low drug solubility, and poor drug stability. To overcome these limitations, employing a nano-based DDS emerges as a potential solution. Nano-based DDSs provide various advantages like controlled and sustained drug release, improved targeting, and reduced systemic toxicity.MSC: Mesenchymal stem cells; OB: Osteoblast; OC: Osteoclast

Nanomaterials are characterized as materials with dimensions ranging from 1 to 100 nanometers. In this submicron scale, the properties of materials undergo significant changes attributed to quantum effects and the amplified surface-to-volume ratio [61]. These unique characteristics make nanomaterials particularly promising for applications in the field of OP treatment and bone tissue engineering. Some of the main approaches will be described next.

Inorganic nanomaterials

Nanomaterials have garnered significant attention in the realm of bone repair owing to their distinctive osteoconductive properties. They exhibit potential as carriers for localized drug delivery, enabling the administration of diverse therapeutic agents. These nanomaterials possess customizable pore structures, adjustable pore sizes, and large surface areas, rendering them suitable for accommodating therapeutic entities ranging from proteins and genes to small molecular drugs. Moreover, these NPs can be further modified to regulate drug loading and influence the behavior of bone cells and immune cells [62,63,64,65].

Silica NPs

Silica, a biocompatible element, exhibits numerous biological advantages and has demonstrated the potential to enhance the osteoconductivity of hydroxyapatite (HA) bone scaffolds. Nevertheless, the underlying mechanism by which silica modulates skeletal development remains largely unexplored [66, 67]. To synthesize silica NPs, the sol-gel method was employed utilizing a tetraethoxysilane (TEOS) solution. The sol-gel process involved the hydrolysis of TEOS molecules, followed by condensation reactions facilitated by water and alcohol, which promoted nucleation and subsequent growth. As a result, the conversion of TEOS facilitated the formation of a silica network, ultimately yielding silica-based NPs [68].

Beck et al. investigated the in vitro effects of 50 nm fluorescent silica-based NPs on the differentiation of OBs and OCs. The study findings unveiled that these silica NPs exhibited the capacity to diminish the formation of multinucleated cells (RAW 264.7) in a concentration-dependent manner (25–100 µg/mL) without inducing apoptosis in OC precursors. Notably, the silica-based NPs also stimulated the differentiation of pre-OBs (MC3T3-E1) into mature OBs, and this effect was observed to be concentration-dependent [69].

Peptides, which have high biocompatibility and specific tissue-targeting abilities, have long been used for several biomedical applications [70, 71]. In one study, to overcome some of the limitations of salmon calcitonin, such as its short half-life, it was loaded into pentapeptide-decorated silica NPs (SiO2-Pep@sCT). It was found that the alkaline phosphatase activity (ALP) of MC3T3-E1 cells was significantly higher (2 times and 1.4 times) in the SiO2-Pep@sCT group than in the SiO2 and salmon calcitonin groups, respectively, indicating that the external negative charge of the NPs and salmon calcitonin stimulated the differentiation of OBs. The number of calcified nodules was also higher in the SiO2-Pep@sCT group. In vivo bone parameters, such as BMD (1.2 times), the bone surface/bone volume ratio and bone volume/total volume ratio were all increased in the SiO2-Pep@sCT group. Therefore, silica NPs can be used as a potential drug carrier in OP and other skeletal diseases [72].

Mesoporous bioactive glasses (MBG) have been used for orthopedic and dental applications due to their strong bonding properties with the bone tissue. Moreover, they are biodegradable, which enables controlled drug release [73].

Estradiol (E2) use is a crucial therapeutic method for women with postmenopausal OP [74]. However, long-term treatment has disadvantages. In a previous study, mesoporous bioactive glass NPs (MBGNPs) were used to encapsulate E2. Moreover, they were modified with β-cyclodextrin (CD-MBGNPs) to improve their affinity for E2. These particles were then electrospun with silk fibroin (SF) to create a nanofibrous mesh (E2@CD-MBGNPs/SF). Further characterization of the NPs revealed a burst release in the first 48 h, followed by further release after day 13. The ALP activity of the cells (MC3T3-E1) was increased after treatment with E2@CD-MBGNPs/SF on day 7. In the case of OCs, the E2@CD-MBGNPs/SF group could reduce the total DNA amount, indicating the inhibiting nature of the nanofiber. Tartrate-resistant acid phosphatase (TRAP) activity was found to be significantly suppressed in the E2@CD-MBGNPs/SF group. Additionally, the presence of TRAP + multinucleated cells and the actin-ring formation was minimal in the E2@CD-MBGNPs/SF group versus the others. These observations all suggest that E2@CD-MBGNPs/SF is a potential anti-osteoporotic drug [75].

Cerium (Ce) is an element with diverse biological functions, including potent antioxidant properties that make it suitable for applications such as radiation protection, cardiovascular diseases, and neurological disorders [76,77,78,79,80]. In a previous study, mesoporous silica NPs used to encapsulate nanoceria (Ce@MSNs). Ce@MSNs were found to be toxic at concentrations exceeding 100 µg/mL but were internalized by MC3T3-E1 cells after 24 hours, allowing direct interaction with mitochondria to reduce oxidative stress. The ALP activity and calcification level of MC3T3-E1 cells were significantly increased in the Ce@MSNs group at 100 µg/mL compared to that of the MSNs group. Ce@MSNs were also found to decrease the number of multinucleated cells (RAW 264.7) when co-cultured with MC3T3-E1 cells, indicating their potential as anti-osteoclastic agents [81].

Long non-coding RNAs (lncRNAs) are a class of RNA with no protein-coding ability, and their role in bone formation is not well-studied. However, lnc-ob1 and ORLNC1 have been identified as crucial for bone formation [82, 83]. In a previous study, bioactive glass NPs (BGN) were used to induce bone marrow-derived mesenchymal stem cells (BM-MSCs) to secrete extracellular vesicles (EVs). It was reported that the EVs isolated after exposure to BGN successfully attenuated OC differentiation. Parameters associated with OC activity, such as the formation of multinucleated cells, TRAP activity, and the expression of NFATc1, were significantly downregulated after treatment with BGN-EVs [84].

Titanium nanotubes

Titanium nanotubes exhibit significant potential as materials for bone-tissue engineering, owing to their advantageous physicochemical properties that facilitate the adhesion and proliferation of OBs. Furthermore, they have been extensively investigated for their capacity to serve as carriers for osteogenic substances that encourage bone formation and enhance the differentiation of OBs. Promising outcomes from preclinical studies have demonstrated the exceptional osseointegration properties of these nanotubes, thereby establishing them as highly suitable contenders for orthopedic implants and therapies aimed at bone regeneration [85,86,87,88].

The primary methodology employed for fabricating titanium nanotubes is electrochemical anodizing. This process involves the preparation of titanium and platinum sheets, which are subsequently immersed in an electrolyte solution. The titanium sheet, acting as the substrate for nanotube growth, is connected to the positive electrode, while the platinum sheet is connected to the negative electrode. In this electrochemical configuration, the primary oxidation of titanium occurs at the anode, specifically at the interface between the metal and the oxide layer. Concurrently, reactions such as hydrogen evolution and oxygen reduction take place at the platinum electrode. By applying a continuous voltage, metal ions migrate outward through the oxide-electrolyte interface, while oxygen molecules migrate inward from the electrolyte-oxide interface to the surface of the metal-oxide. These migration processes result in the formation of titanium crystals, ultimately leading to the fabrication of titanium nanotubes [89].

Mu et al. utilized TiO2 arrays as carriers for raloxifene, and subsequently coated them with a hybrid multilayered coating composed of chitosan and alendronate-grafted hyaluronic acid (TNT/Ral/LBL-Aln). Through in vitro experiments, it was observed that TNT/Ral/LBL-Aln exhibited a twofold increase in ALP expression and a 1.4 times increase in mineralization. Additionally, it demonstrated the ability to inhibit OC differentiation. In vivo studies further demonstrated that the administration of raloxifene led to an increase in bone formation and a decrease in the level of TRAP activity, resulting in reduced trabecular space. These findings suggest that raloxifene not only suppresses OC differentiation but also enhances bone formation, thereby positioning it as a promising therapeutic option for OP [90].

Icariin (ICA), the primary active component found in Epimedium, has been identified as a promoter of bone formation. Furthermore, it has been observed to support fracture healing by mitigating oxidative stress [91]. Strontium (Sr) is another osteogenic component known for its exceptional biocompatibility [92]. In one particular study, a TiO2 nanotube surface was coated with Sr and ICA to investigate their effects on the cell adhesion, proliferation, and osteogenic differentiation of pre-OBs, as well as bone formation surrounding titanium implants. The group treated with TiO2 + Sr + ICA demonstrated a significantly higher proliferative and adhesive effect on MC3T3-E1 cells. Moreover, ALP activity increased by 1.7 times, and mineralization was found to be the highest in the TiO2 + Sr + ICA group, followed by the TiO2 + Sr and TiO2 groups. While new bone formation occurred around the implants in all three groups in vivo, the TiO2 + Sr + ICA group exhibited a greater bone volume relative to the total tissue volume (BV/TV) [93].

Calcitonin is a small peptide hormone that has been reported to inhibit OC-mediated bone resorption. Calcitonin also targets Wnt10b in OCs and promotes bone formation [94]. Studies have shown that calcitonin gene-related peptide (CGRP) immobilized onto TiO2 nanotubes (TNT-CGRP) can affect the differentiation of OBs and OCs in vitro. TNT-CGRP can indicate better cell adhesion on the implant, with better cell spreading and stress fiber formation. ALP activity was greatly noticed in the TNT-CGRP group, which was more than what was displayed in the other groups. In addition, mineralization was also observed to be the highest in this group. Osteogenic-related gene expression, such as collagen type I (ColI), Runt-related transcription factor 2 (Runx2), Osteopontin (Opn), and Osteoprotegerin (Opg), was greater in the TNT-CGRP group than in the Ti group in the 7-day culture. The mRNA levels of Col1, Runx2, Opn, and Opg were 3.0-, 2.2-, 2.1-, and 2.0-fold higher than those of the Ti group, respectively. In the case of OC differentiation, TRAP activity was lowest in the TNT-CGRP group in comparison with other groups. The expression levels of OC-related genes, such as vacuolar (H+) ATPase, matrix metalloproteinase 9, TRAP, and cathepsin K, were decreased after treatment with TNT-CGRP. Therefore, TNT-CGRP could improve bone formation while decreasing OC differentiation [95].

Hydroxyapatite NPs (HA NPs)

Hydroxyapatite (HA) is a natural inorganic material found in the bones and teeth of humans and vertebrates [96]. The chemical structure of HA is mainly known as Ca10(PO4)6(OH)2. HA is a biocompatible, osteoconductive, and bioactive ceramic widely recognized for its ability to establish direct bonds with living tissues. This property is attributed to its inherent high affinity towards collagen type I, making HA an exceptionally reliable candidate for bone formation [97]. In the field of bone regeneration, HA scaffolds and materials have been extensively utilized. Recently, HA nanoparticles have garnered attention due to their large surface area, which facilitates enhanced protein adsorption and cell interaction. The production of HA nanoparticles can be achieved through both dry-synthesis and wet-synthesis methods. Dry synthesis methods typically yield well-crystallized and stoichiometric products but require high temperatures and lengthy treatment times. On the other hand, wet-synthesis methods can produce nanoparticles at relatively low temperatures, but the resulting crystallinity and calcium-to-phosphorus (Ca/P) ratio are comparatively lower [98].

Zoledronic acid (ZOL), a bisphosphonate (BP) with anti-resorptive properties, is commonly employed in the treatment of OP [99]. In a recent study focused on investigating the potential synergistic effects of hydroxyapatite (HA) nanoparticles loaded with ZOL (HNLZ) on OP, promising results were obtained. In the HNLZ group, the serum levels of bone-specific ALP, procollagen type I N-terminal propeptide, osteocalcin, and TRACP 5b were significantly lower compared to those of the sham-operated group [100].

The potential of targeted delivery for OP has remained in its infancy. This study utilized salmon calcitonin (SCT)-loaded HA NPs (SCT-HAP-NPs) for the treatment of sublingual OP as a non-invasive therapy. It was observed that the SCT-HAP-NPs deeply penetrated the tissue through the stratified squamous epithelial layer via the basement membrane. Additionally, the SCT-HAP-NPs could significantly lower serum calcium and phosphorous levels followed by a significant decrease in the resorption pits [101].

Research has shown that targeted delivery of BP for OP is an area that requires further investigation. In a study using risedronate/zinc-hydroxyapatite NPs (ZnHA-NPs), it was found that ZnHA-NPs attenuated the increase in serum levels of bone-specific ALP and preserved the structural integrity of the cortical and trabecular bones. The study also showed that ZnHA-NPs successfully regulated the level of TRAP-5b, indicating the action of OC-mediated bone resorption [102].

Zhao et al. investigated the use of HA bioceramics composed of a micro whiskered scaffold strengthened with multiple layers of releasable HA NPs (nwHA) as a treatment option for OP. The study showed that the nwHA group induced new bone formation and had significantly higher expressions of the genes responsible for bone formation, ATP2A2 and FGF23, upon implantation in critical-sized femur defects in osteoporotic rats [103].

Hwang et al. tried to achieve the bone-specific dual delivery of both a drug and a mineral. To do this, hydroxyapatite nanoparticles (HA NPs) were modified with alendronate (Aln) and coated with multilayers of poly(allylamine) (PAA) and alginate (ALG). The resulting compound, Alen-LBL-HA NPs, was synthesized and investigated for its effects on bone-related cells. In vitro studies using MC3T3-E1 cells demonstrated that Alen-LBL-HA NPs exhibited high proliferation rates and significantly increased ALP activity. These findings suggest that Alen-LBL-HA NPs possess properties that enhance cell proliferation and promote bone formation [104].

Additionally, calcium-rich hydroxyapatite nanoparticles (CRHNPs) were examined for their impact on BM-MSCs in the context of OP. The study revealed that CRHNPs promoted the proliferation of BM-MSCs while reducing apoptosis. Furthermore, the expression levels of osteogenic markers, such as Runx2 and OPN, were significantly increased by 1.6-fold and 1.5-fold, respectively, compared to those of the control groups. These results indicate that the use of HA NPs may hold promise as an approach for the treatment of OP [105].

Metallic NPs

Tissue engineering extensively employs metal nanoparticles. Materials like Au, Ag, Fe, Al, Ni, Cu, Zr, and magnetic nanoparticles (MNPs) have been extensively investigated for this purpose. Although previous research highlighted the toxicity of certain metal NPs, such as Ni, it has now been established that when used in appropriate sizes and doses, metal NPs offer numerous benefits. Additionally, metallic NPs commonly possess desirable traits like high surface areas and antibacterial properties [106, 107].

Magnetic NPs (MNPs)

MNPs, which predominantly consist of iron oxide II and III, are composite crystals composed of magnetic elements such as Fe, Ni, or Co. Ni NPs exhibit desirable properties such as good Curie temperatures, chemical stability, and high saturation magnetization, making them excellent candidates for drug delivery systems. Cobalt ferrite NPs, on the other hand, possess exceptional mechanical hardness, wear resistance, ease of synthesis, and electrical insulation properties, which make them promising agents for various medical applications, including magnetic drug delivery. Additionally, MNPs can also serve as contrast agents for magnetic resonance imaging. Furthermore, MNPs can be utilized to induce thermolysis by exposing cells to radiofrequencies, offering a potential avenue for hyperthermia-based therapies [108,109,110,111].

Over time, diverse strategies have been developed to synthesize MNPs with control over their size, morphology, stability, and biocompatibility. These approaches can be broadly categorized into physical, chemical, and biological methods. Physical methods include techniques such as ball-milling, laser evaporation, and wire explosion. Chemical methods encompass coprecipitation, thermal decomposition, microemulsion synthesis, hydrothermal synthesis, and sol-gel methods. On the other hand, biological methods utilize the byproducts of plants or microorganisms and have gained recognition for their ability to synthesize MNPs [112].

Research has been conducted on fabricating BP-conjugated MNPs (BP/Dex/Fe3O4) to inhibit OC activation. Bis/Dex/Fe3O4 has a high binding affinity to bone grafts and decreases the activation of OCs by BP. Continuous radiofrequency exposure to BP/Dex/Fe3O4 induced thermolysis of OCs while having no effect on the survival of osteoblasts [113].

Tran et al. have shown that hydroxyapatite-coated Fe3O4 NPs (HA-IONP) significantly upregulates ALP, collagen, and calcium in OBs. The results indicate that HA-IONP promotes OB differentiation. The authors found that HA-IONP adsorbed a large amount of fibronectin and consequently enhanced the function of OBs and upregulated genes related to OB differentiation [114].

In another study, Li et al. developed HA-coated superparamagnetic iron oxide NPs (SPIO@HA) with a core − shell structure for targeting both osteoclastogenesis and osteogenesis. SPIO@HA exhibited chemical stability and low cytotoxicity in in vitro experiments. It promoted the differentiation of MSCs to OBs while inhibiting OC formation and downregulated genes related to osteoclastic differentiation. Additionally, SPIO@HA prevented bone loss and increased BMD in the OVX mouse model [115].

Extracellular vesicles (EVs) derived from MSC (MSC-EVs) can deliver therapeutic targets for various diseases [116]. Despite this, the isolation and detection of EVs still exhibit several technical drawbacks such as limited sensitivity and time consumption [117]. In a previous study, gold-coated magnetic nanoparticles (GMNPs) were used to load EVs. The surfaces of GMNPs were decorated with a Fe3O4@SiO2 core and a silica shell with PEG-aldehyde (CHO) to examine its role in diabetic osteoporosis (DO). Microarray analysis revealed that OP-associated miR-150-5p was differentially expressed. To establish models of OP, rats were injected with streptozotocin, and bone tissue analysis confirmed the reduced expression of miR-150-5p. Subsequently, a combination of GMNPs and anti-CD63 formed GMNPE, which was then co-cultured with OBs. The reintroduction of miR-150-5p facilitated osteogenesis in the OBs. GMNPE played a role in enriching EVs in the bone tissues of the rats with OP. The miR-150-5p carried by BMSC-EVs targeted MMP14, thereby activating the Wnt/β-catenin pathway. This activation then enhanced the proliferation and maturation of OBs. Additionally, GMNPE improved the delivery of miR-150-5p via EVs, effectively regulating the MMP14/Wnt/β-catenin axis and promoting osteogenesis [118].

Gold NPs

Gold NPs (GNPs) have been widely developed for therapeutic application of biologic therapies, including DDS of drugs and genes, photographic agents, photothermal therapies, biosensors, and diagnostic reagents. GNPs are highly influenced by the physical and chemical characteristics of their synthesis such as reaction temperature, stirring rate, and the ratio of gold to the reducing agent. To prepare GNPs, many researchers developed GNP fabrication methods such as green synthesis using plants or bacteria, the Turkevich-Frens method, Brust-Schiffrin method, Martin method, and Seeding-Growth method. GNPs have long been known as osteoinductive agents that inhibit OC formation [119,120,121,122,123,124,125,126]. High amounts of GNPs in the body can be toxic; therefore, it is necessary to modify the surface of these particles so that they reach their target sites.

BPs have been used as therapeutics against osteoporosis [127]. BPs, such as alendronate, have been conjugated with GNPs (GNPs-ALD) to specifically target OCs and inhibit their differentiation, and thereby, bone resorption. GNPs-ALD was also successful in inhibiting OC differentiation. It was found that 20 µM of GNPs-ALD fully inhibited the formation of RANKL-induced osteoclastogenesis. In addition, GNPs-ALD at 20 µM could inhibit the formation of TRAP + multinuclear cells, compared to the positive control group. The expression of OC-specific genes like OSCAR, c-Fos, and NFATc1 were also significantly reduced when the cells were treated with 20 µM of GNPs-ALD. In addition, in vivo data showed that all OVX groups, except GNPs-ALD, had a lower trabecular bone volume [128].

β-cyclodextrin-conjugated GNPs with curcumin (CUR-CGNPs) as an inclusion complex have been used to determine their effects on receptor activator of nuclear factor-κb ligand (RANKL)-induced osteoclastogenesis in bone-marrow–derived macrophages. In addition, CUR-CGNPs could significantly reduce the number of TRAP + multinuclear cells. The real-time PCR analysis showed that OC marker genes, such as c-Fos, NFATc1, and OSCAR, were also significantly downregulated after treatment with CUR-CGNPs (Fig. 2A). While RANKL stimulated actin-ring formation, which allows OCs to resorb bone, CUR, CGNPs, and CUR-CGNPs reduced actin-ring formation. Particularly, CUR-CGNPs showed almost no actin-ring formation, implying that CUR-CGNPs are the most useful therapeutic agents (Fig. 2B). In vivo data showed that CGNPs and CUR-CGNPs increased the bone density with a smaller trabecular number, in comparison with the OVX group. The BMD was also found to be high in the CUR-CGNPs treatment group [129].

Fig. 2
figure 2

Effect of CUR-CGNPs on OC. (A) β-cyclodextrin-conjugated GNPs with curcumin inhibits the osteoclastic differentiation of bone-derived macrophages. (B) Immunofluorescence analysis of bone-derived macrophages incubated in osteoclastic differentiation medium with CUR, CGNPs, or CUR-CGNPs for F-actin expression (RANKL-induced actin-ring formation). GNP: Gold nanoparticles; BMSCs: Bone marrow derived mesenchymal stem cells; CUR: Curcumin; CGNPs: β-cyclodextrin-conjugated GNPs; CUR-CGNPs: Curcumin loaded β-cyclodextrin-conjugated GNPs; RANKL: Receptor activator of nuclear factor κB. (B) is reprinted (adapted) with permission from Heo, D.N., Ko, W.K., Moon, H.J., Kim, H.J., Lee, S.J., Lee, J.B., Bae, M.S., Yi, J.K., Hwang, Y.S., Bang, J.B., and Kim, E.C., 2014. Inhibition of osteoclast differentiation by gold nanoparticles functionalized with cyclodextrin curcumin complexes. ACS nano, 8(12), pp. 12,049–12,062. Copyright 2014 American Chemical Society

Nah et al. investigated the effect of vitamin-D–conjugated GNPs (VGNPs) on RANKL-induced osteoclastogenesis. PEG-containing sulfhydryl groups were used to attach vitamin D to the surface of the GNPs. The VGNPs were successfully internalized by bone marrow macrophages, and it was found that 20 µM of VGNPs could decrease the formation of TRAP + multinuclear cells. Moreover, VGNPs downregulated genes related to OC differentiation, such as TRAP, OSCAR, NFATc1, and c-Fos [130].

The transcription factor c-myb is a member of the myeloblastosis (MYB) family and is crucial for cell differentiation, survival, death, and proliferation [131,132,133]. Several studies have investigated its role in osteogenesis and odontogenesis. In one study, chitosan-gold nanoparticles conjugated with plasmid DNA/c-myb (Ch-GNPs/c-myb) were found to upregulate c-myb expression and stimulate osteogenesis on titanium surfaces in MC3T3-E1 cells. Ch-GNPs/c-myb also inhibited OC differentiation in bone marrow-derived macrophages, significantly reducing the number of TRAP + multinucleated cells [134].

Polymeric NPs

Polymeric nanoparticles exhibit superior stability within the gastrointestinal (GI) tract when compared to alternative colloidal carriers. This enhanced stability enables them to shield encapsulated drugs from the harsh conditions encountered in the GI environment. The utilization of diverse polymeric materials allows for the deliberate adjustment of physicochemical attributes, pharmacokinetic properties, and biological behaviors of nanoparticles. Furthermore, the surface of these particles can be conveniently modified through the adsorption or chemical grafting of specific molecules, such as polyethylene glycol (PEG), poloxamers, and bioactive compounds. These modifications offer opportunities to tailor the surface properties of NPs, thereby influencing their interaction with biological entities and improving their therapeutic potential [135, 136].

Poly lactic-co-glycolic acid nanoparticles (PLGA NPs)

PLGA, a hydrophobic biopolymer renowned for its exceptional biodegradability and biocompatibility characteristics, has emerged as a prominent candidate in the realm of biomedical applications. Remarkably, PLGA generates a biocompatible byproduct that can be efficiently eliminated via metabolic pathways, thereby bolstering its safety profile and rendering it a preferred choice in clinical therapeutic interventions. Functioning as an exemplary drug carrier, PLGA offers the ability to encapsulate a diverse array of pharmaceutical agents, thus conferring remarkable benefits such as enhanced bioavailability and sustained release of the encapsulated drug. This advantageous attribute holds significant importance and has propelled the widespread utilization of PLGA as a versatile platform for DDSs [137, 138].

PLGA NPs were synthesized using an emulsion method, specifically through two types of emulsions: (1) water/oil emulsion and (2) water/oil/water emulsion. Initially, PLGA powder was dissolved in an organic solvent such as chloroform, dimethyl sulfoxide, or dichloromethane. Subsequently, the PLGA solution was combined with a water-based solution, resulting in the formation of an emulsion. The mixed solution was then subjected to stirring while the organic solvent underwent evaporation. Finally, the remaining PLGA nanoparticles were obtained through the process of centrifugation [139].

The therapeutic efficacy of simvastatin (SIM)-loaded tetracycline-mediated PEG-PLGA (TC-PEG-PLGA) micelles was evaluated through their administration in osteoporotic rats, thereby delineating their potential impact. Intriguingly, the TC-PEG-PLGA micelles exhibited a remarkable augmentation in mineralization, with a remarkable two-fold increase observed in comparison to the negative control, specifically in MC3T3-E1 cells. Furthermore, the experimental findings unveiled that the administration of these micelles elicited a notable induction of osteogenesis in the rat model of osteoporosis. Remarkably, the bone mineral content of the rats subjected to TC-PEG-PLGA was significantly higher, with a 1.2-fold increase compared to that of the control groups, as documented in a previous study [140].

Gene delivery is an efficient method for treating OP [141]. In this approach, PLGA nanocapsules loaded with a PEI-RANK siRNA complex are used to suppress OC differentiation. The PLGA nanocapsules can be successfully internalized by RAW 264.7 cells, whose OC differentiation is consequently inhibited. Moreover, the nanocapsules can effectively downregulate RANK, leading to a significant reduction (50%) in the RANK mRNA level in OC precursor cells. The TRAP enzyme is also downregulated in the group exposed to the nanocapsules, indicating that PLGA nanocapsules loaded with a PEI-RANK siRNA complex can effectively inhibit OC differentiation [142].

17β- estradiol (E2) has several anabolic effects in maintaining the structural integrity of the bone. Postmenopausal OP is a serious issue among women for which there are very limited therapies [143]. In a study on transdermal DDSs, E2-loaded PLGA NPs were used for OP therapy. As E2 is affected by the first-pass hepatic metabolism, the study investigated the effect of iontophoresis on the dermal permeabilization of E2. It was found that the E2-loaded PLGA NPs significantly increased the dermal permeabilization of E2 upon applying iontophoresis. Furthermore, E2-loaded PLGA NPs increased the BMD of cancellous bone when compared with the OVX mouse control group [144].

Gelatin NPs

Gelatin is a promising material for various biomedical applications due to its biocompatibility, biodegradability, and non-toxic nature [145, 146]. In addition, using gelatin composites, such as NPs, microparticles, 3D scaffolds, and electrospun nanofibers, improved mechanical properties. Due to the release pattern of bioactive molecules, they can be controlled depending on the cross-linking density of gelatin. Gelatin NPs have been widely used as drug and gene carriers [147]. The most common method to fabricate gelatin NPs is cross-linking by glutaraldehyde (GA), genipin, and carbodiimide/N-hydroxysuccinimide. Briefly, gelatin and drugs in the aqueous phase (salt water or alcohol) are homogenized with the oil phase (olive oil, polymethyl methacrylate, or paraffin oil) and then cross-linked with GA or genipin. To collect gelatin NPs, water, and the oil solvent are removed by evaporation, filtration, centrifugation, and lyophilization [148].

In a study conducted by Yang et al., a polydopamine-coated porous titanium scaffold was designed to be integrated with zoledronic acid (ZOL)-loaded gelatin NPs to investigate their effects on osteogenesis and osteoclastogenesis. From the findings illustrated in Fig. 3, it was discovered that OBs displayed a notable augmentation in their morphological elongation and the presence of filamentous filopodia when exposed to concentrations ranging from 1 µmol/L to 50 µmol/L. However, in the groups exposed to 100 µmol/L and 500 µmol/L, the number of cells adhering to the scaffold diminished, and they exhibited signs of atrophy and reduced pseudopodium formation. In addition, there was no significant difference in OC attachment to the scaffold at concentrations of 1–10 µmol/L. The osteogenic effect of ZOL-loaded NPs (ALP activity and expression of genes such as Runx2 and ALP) was highest at 50 µmol/L, and the number of mature OCs (multinucleated cells) decreased at this concentration [149].

Fig. 3
figure 3

SEM image showing the cell attachment and proliferation of OBs on different concentrations of ZOL-loading scaffolds. OBs exhibited increased morphological elongation and filamentous filopodia at concentrations of 1 µmol/L to 50 µmol/L (Fig. 6). However, at concentrations of 100 µmol/L and 500 µmol/L, cell attachment to the scaffold decreased, resulting in atrophy and reduced pseudopodium formation. OB: Osteoblast; OC: Osteoclast. Reproduced with permission from Yang et al. [73]. (Copyright 2020, IOP Publishing Ltd.)

Strontium ranelate (SR) has the advantage of promoting bone formation and inhibiting bone resorption. However, high doses of SR may cause heart and kidney-related complications with frequent intake [150,151,152]. In a previous study, a cross-linking strategy was employed, which included enzyme-crosslinking using tyrosinase and physical folding to achieve SR-loaded gelatin NP/silk fibroin aerogel (S/G-Sr/MT). S/G-Sr/MT showed the highest level of ALP activity. In addition, the gene expression levels of the osteogenic markers ALP, Runx2, Col1, and Osterix were 10.0-, 6.0-, 1.5-, and 10.0-folds higher than those of the S/G group, respectively. Although the in vivo bone parameters (BV/TV, trabecular number [Tb.N], trabecular separation [Tb.Sp], and BMD) of the S/G-Sr/MT group were similar to those of the other groups, the S/G-Sr/MT group showed the highest bone union rate at each testing time point. Additionally, S/G-Sr/MT was also able to downregulate the activity of TRAP, a marker of bone resorption [153].

Chitosan NPs

Chitosan has been successfully used in many fields related to human health, pharmaceuticals, and the environment. Its biodegradability has paved the way for its numerous applications in various fields as a treatment option [154]. Furthermore, chitosan nanoparticles (NPs) play a significant role in protecting drugs from enzymatic degradation and minimizing adverse effects on non-targeted tissues or cells. These nanoparticles possess a positively charged surface and exhibit controlled and sustained drug release characteristics [155]. The fabrication of chitosan NPs commonly involves methods such as emulsification and crosslinking. Additionally, chitosan NPs can be prepared using alternative approaches including reversed micelles, phase inversion precipitation, and emulsion-droplet coalescence [156].

In a study on glucocorticoid-induced OP rats by Alshubaily et al., the anti-osteoporotic effect of Shilajit-loaded chitosan NPs was investigated. The study used nanochitosan (NCT) and NCT conjugated with a shilajit water extract (SWE) (NCT-SWE) as the main test groups. The results showed that NCT-SWE was very efficient in enhancing the levels of calcium, phosphorus, osteocalcin, and calcitonin. It was also successful in reducing hydrogen peroxide levels, thereby maintaining the level of antioxidants. Therefore, NCT-SWE could decrease oxidative stress and upregulate biomarkers of bone formation [157].

Santhosh et al. found that the treatment efficiency of BPs, such as risedronate, against osteoporosis was increased by functionalizing risedronate with chitosan NPs (RISCN). The results showed that RISCN has more affinity toward human farnesyl diphosphate synthase (FDPS) in the mevalonate pathway, thereby blocking the process of bone resorption by inhibiting osteoclastogenesis. Thus, RISCN is considered highly target-specific in treating OP [158].

BMP-2, a growth factor recognized for its ability to induce bone formation, faces limitations when administered as an injection due to its short half-life and poor retention efficiency [159, 160]. To address this challenge, researchers have developed an innovative approach using a dual-function injectable fibrin gel (Fg) combined with semisynthetic sulfated chitosan NPs (SCS-NPs) loaded with recombinant human BMP-2 (rhBMP-2). In a previous study, the Fg loaded with 20 mg of SCS-NPs and 5 µg of rhBMP-2 demonstrated a remarkable threefold increase in the gene expression of key markers for bone formation, including type 1 collagen (Col-1), Osterix (Osx), and Runx2. Moreover, this formulation exhibited accelerated bone formation in vivo [161].

Nanogels

Nanogels have been utilized as carriers for various drugs and are classified based on the bonds involved in polymer network formation. Chemically cross-linked nanogels are polymer chains that are covalently bonded, while physically crosslinked nanogels are formed by weaker interactions, such as hydrogen bonds, electrostatic interactions, and hydrophobic interactions [162].

Chemical crosslinking serves as the primary approach for synthesizing nanogels, encompassing techniques like emulsion polymerization, controlled/living radical polymerization, click chemistry, and photo-crosslinking. This method involves utilizing low molecular weight monomers, short polymers, or polymer precursors with reactive end groups. Initiating materials, such as initiators, catalysts, and crosslinking agents, play a crucial role in the process. In nanogel formation, radicals are generated by the initiator through various means such as heat, hydrogen ions, or light. These radicals initiate reactions with monomers or polymers, resulting in the formation of monomer radicals or polymer radicals. Subsequently, the monomer or polymer radicals engage with other monomers or polymers, establishing covalent bonds. Furthermore, radicals derived from previously formed covalent bonds continue the chain of reactions, leading to the creation of additional covalent bonds. Ultimately, this radical-induced polymerization reaction culminates in the formation of nanogels. In contrast, the physical crosslinking method for nanogel formation relies on supramolecular polymers or biomolecules. These materials facilitate the spontaneous aggregation or self-assembly of nanostructures without the necessity of crosslinking agents. The driving forces in this process include ionic and hydrophilic-hydrophobic interactions, as well as Van der Waals and hydrogen bonds. Notably, during the formation of physically cross-linked nanogels, properties such as size, morphology, and strength undergo alterations influenced by factors such as ionic strength, temperature, and pH [163, 164].

Glucocorticoid-induced OP is the most prevalent cause of secondary OP, where exposure to glucocorticoids worsens the risk of fracture and bone loss. In one study, a transdermal nanoemulsion gel formulation was developed, and lovastatin was loaded into the nano-sized globules of the nanoemulsion to promote skin layer entry while avoiding liver metabolism. It was discovered that biomechanical strength testing indicated better strength and load-bearing capacity, and anabolic markers for bone formation were also found to be elevated, while levels of bone resorptive markers decreased [165].

Zhang et al. developed a combination nanogel scaffold (N.E) fabricated from p(N-isopropylacrylamide-co-butyl methacrylate) and a strontium MBG scaffold (Sr-MBGS). N.E can retain large amounts of water and harden at body temperature, making it suitable for transplantation into living cells. The combination of N.E and Sr-MBGS particles was found to serve as an excellent delivery vehicle for primary OBs delivered directly to the scaffold site for tissue regeneration [166].

Nanogel-mediated peptide delivery has also been investigated as a treatment option for bone loss. In one study, a cholesterol-bearing pullulan (CHP) nanogel was used to encapsulate the W9-peptide, TNF-α, and RANKL antagonist. A decrease in BMD was significantly prevented in the CHP-W9 peptide-injected groups. Additionally, histomorphometric analysis of the proximal tibiae showed a significant prevention in changes to bone resorption parameters, such as a decrease in the number of mature OCs [167].

Guo et al. synthesized a raloxifene HCL-loaded solid NP (RAL-SLNs) decorated gel to alleviate the effects of OP. The successful loading of RAL-SLNs into the gel allowed for its retention for a long period of time, along with the convenience of transdermal delivery. Biochemical analysis of the ALP and calcium levels in OP-induced rat models showed that the levels were significantly higher after treatment with RAL-SLNs [168].

Polyurethane nanomicelles

Polyurethane, a polymer composed of organic units linked by carbamate bonds, offers excellent biocompatibility, low cytotoxicity, and mechanical flexibility, making it ideal for drug encapsulation and long-term stability [169]. In addition, polyurethane can form a micelle structure through its self-assembly property in an aqueous solution. Thus, the polyurethane micelles are composed of a hydrophilic shell and a hydrophobic core, allowing them to be used as a DDS for encapsulating either hydrophilic or hydrophobic drugs. Also, polyurethane micelles can have a target effect on specific tissues by surface modification with antibodies and peptides [170, 171].

Peptide targeting is a promising approach in its early stages of development. Researchers have used polyurethane nanomicelles modified by the peptide ASP8 (PU-ASP8) to deliver anti-miR214 specifically to OCs, reducing miR214 efficiency by 80%, increasing bone mass, and decreasing trabecular spacing with an increase in BMD (Fig. 4) [172]. Similarly, SDSSD-modified polyurethane (PU) nanomicelles encapsulating anti-miR-214 were developed to target OBs, which resulted in an 80% reduction in miR-214 levels in OBs after SDSSD-PU-anti-miR-214 treatment. Targeting specific cell types in OP treatment may be a promising avenue for future research to improve drug targeting efficiency. SDSSD-PU-anti-miR-214 was also successful in improving the bone mass density along with inducing a higher mineral apposition rate with high miRNA stability [173].

Fig. 4
figure 4

PU-ASP8-modified polyurethane nanomicelles targeting OCs. ASP8 has a higher affinity towards crystallized hydroxyapatite surfaces (resorbed bone surface). After the PU-ASP8 reaches the bone microenvironment, it undergoes cellular uptake by OCs, leading to an 80% reduction in the expression of miR-214. This substantial decrease in miR-214 levels plays a crucial role in improving bone health by promoting an increase in bone mass and a simultaneous decrease in trabecular space. PU-polyurethane; ASP8-eight repeating sequences of aspartate; OC- osteoclast

Lipid-based NPs

Lipid-based nanocarriers have gained attention as potential vehicles for OP drugs due to their high biocompatibility, biodegradability, and ability to release drugs in a controlled manner after various routes of administration [174,175,176,177,178]. Liposomal formulations and solid-core micelles are the most widely studied lipid-based NPs, with surface modifications improving their therapeutic outcomes such as long-circulation, tissue-targeted effect, and pH-sensitivity. The different approaches used for the fabrication of lipid-based NPs are high-pressure homogenization, emulsion, solvent evaporation, solvent diffusion, and ultrasonication [179].

In a study investigating the effect of quercetin on bone health, quercetin-based solid lipid NPs (QSLNs) were formulated. QSLNs were found to restore trabecular and bone mineral density in ovariectomized mice compared to the control group. Additionally, QSLNs significantly downregulated the osteoclastogenic genes RANK, TRAP, and c-Fos, indicating the potential of these NPs as anti-osteoporotic therapeutics [180].

Simvastatin (SIM) is commonly used to treat high cholesterol levels. However, recent studies have suggested that SIM can increase bone formation through BMP-2 [181]. In a study investigating the effect of SIM on bone formation, SIM was encapsulated in lipid NPs with aspartic oligopeptide 6 (ASP6) moieties, which were grafted onto the NPs to target bone formation (SIM/ASP6-LNPs) (Fig. 5A). The formation of mineralized nodules between SIM/LNPs and SIM/ASP6-LNPs indicated that SIM was responsible for the mineralization in both groups at an equal concentration of 10− 7 M. Furthermore, 29% of tetramethylindotricarbocyanine iodide (DiR)-loaded ASP6-LNPs were found in the femur and tibia, suggesting the bone-targeting specificity of the NPs. In vivo bone parameters such as BMD and BV/TV were increased in comparison with the sham-operated group, indicating the potential of SIM/ASP6-LNPs to target bone and promote its formation (Fig. 5B) [182].

Fig. 5
figure 5

A schematic presentation of the improved efficacy of simvastatin-loaded LNPs in treating osteoporosis. (A) SIM was enclosed within lipid nanoparticles (NPs) that were further modified with aspartic oligopeptide 6 (ASP6) moieties. The purpose of incorporating ASP6 onto the NPs was to facilitate targeted delivery and enhance the affinity of the resulting formulation, known as SIM/ASP6-LNPs, towards bone formation processes. (B) Characterization of the LNPs, ASP6-LNPs, SIM/LNPs, and SIM/ASP6-LNPs. Left: TEM images. Right: size distribution based on DLS. Scale bar: 100 nm. (C) Distribution of DiR-loaded LNPs and ASP6-LNPs in ICR mice. Fluorescence image of the major organs (the heart, liver, spleen, lung, and kidney) and the femur and tibia 48 h after the tail-vein administration of the nanoparticles. LNPs: Lipid NPs; ASP6-LNPs: Lipid NPs with aspartic oligopeptide; SIM/LNPs: Lipid NPs delivering SIM; SIM/ASP6-LNPs: SIM encapsulating LNPs with ASP6; TEM: Transmission electron microscopy; DLS: Dynamic light scattering; DiR: Tetramethylindotricarbocyanine iodide; ICR: Institute of Cancer Research. (B) and (C) were reproduced with the permission from Tao et al. [111]. (Copyright 2020, The Royal Society of Chemistry)

Cathepsin K is an enzyme secreted by OC to digest collagen and other bone-matrix proteins. Inhibiting or neutralizing the expression of cathepsin K may suppress bone resorption [183]. In one study, the cathepsin K inhibitor, odanacatib, was loaded into PLGA-derived lipid hybrid NPs, which were modified with bone-specific polyaspartic acid [(ASP)8 DNPs] to investigate its effect on OP. Here, the binding affinity of (ASP)8-DNPs was two times higher than that of the DNPs. Odanacatib-loaded (ASP)8-DNPs were found to have the least TRAP activity and could downregulate OC-related genes, such as Cathepsin K, TRAP, and RANK. Along with inhibiting OC activity, osteogenesis-related genes, such as Runx2 and ALP, were also upregulated after the treatment with odanacatib-loaded (ASP)8-DNPs. Several in vivo bone parameters such as BMD and BV/TV were higher following the odanacatib-loaded (ASP)8-DNPs treatment than with the odanacatib-loaded DNPs treatment [184].

Prior scientific investigations have conclusively demonstrated that MSCs possess the remarkable ability to perceive signals from damaged tissues and exhibit directed migration towards the site of injury. This migration process is facilitated by the intricate interplay between chemo-attractant stromal cell-derived factor 1 (SDF-1) and its specific receptor, chemokine receptor type 4 (CXCR4). Notably, experimental inhibition of CXCR4 has been observed to effectively impede the migratory response of MSCs. Harnessing the inherent homing and engraftment capacities instigated by the SDF-1/CXCR4 axis holds great promise for therapeutic applications in regenerative medicine and tissue repair [185,186,187]. In a previous study, the BMSC secretomes were loaded into PLGA NPs and were further modified by encapsulating the NP inside the CXCR-overexpressed HMEC membrane (MSC-Sec/CXCR4 NPs). MSC-Sec/CXCR4 NPs at a concentration of 106 were able to successfully inhibit RANKL-induced OC differentiation simultaneously, promoting the osteogenic capacity and proliferation of MSCs. The levels of osteocalcin and BMP-2 were found to increase in the OVX rat model. Injection with MSC-Sec/CXCR4 NPs increased the bone volume while decreasing bone resorption [188]. An overall explanation of the nanocarrier types with their drugs along with their effects have been described in Table 2. The impact of different nanocarriers loaded with various drugs on bone has been illustrated in Fig. 6.

Table 2 Effects of several nano-based drug delivery systems in treating osteoporosis. In nano-based DDSs, the bare NPs can affect bone regeneration by forming a bone matrix or promoting mineralization. In addition, drug-encapsulated NPs can increase the stability of the drug in vivo and decrease side effects from excessive drug delivery. Recently, there have been many studies for the bone tissue targeting of NPs by conjugation and hybridization with peptides, BPs, and lipid-based NPs.
Fig. 6
figure 6

A schematic representation of different drug-loaded nanocarriers and their effects. An extensive repertoire of NP formulations, encompassing diverse categories such as inorganic, metallic, polymeric, and lipid nanoparticles, presents a versatile array of options for loading pharmacological agents aimed at combatting OP. These NPs, functioning as potent drug carriers, facilitate precise and targeted delivery, thereby conferring therapeutic benefits that encompass the dual capacity to foster bone formation or impede OC differentiation, thus mitigating the deleterious effects of OP. NP: Nanoparticles; RANK: Receptor activator of nuclear κB; EVs: Extracellular vesicles; HA: Hydroxyapatite; NFATc1: Nuclear factor of activated T-cells; BMP-2: Bone morphogenetic protein; SIM: Simvastatin; BMD: Bone mass density; TRAP: Tartrate-resistant acid phosphatase; OC: Osteoclast; OB: Osteoblast; ALP: Alkaline phosphatase; lncRNA: Long non-coding ribonucleic acid

Discussion

OP is a skeletal disorder that worsens with age, making it crucial to develop new delivery systems to tackle the issue with maximum efficiency. Application of biomaterials for bone regeneration is transforming the lives of patients by reducing off-target effects and increasing therapeutic efficiency. New methods for drug delivery are being developed worldwide, offering promising drugs through material-based DDSs.

Anti-osteoporotic drugs aim to rebalance bone metabolism by promoting OB differentiation or inhibiting OC differentiation. This review explained several nano DDSs that are based on targeted or non-targeted approaches for the treatment of OP. Materials such as Bis, Sr, and HA have been formulated in combination with other bone-targeting molecules to increase the efficiency of osteogenesis. In some cases, bone-targeting peptides together with other osteoconductive materials are used for efficient targeting with fewer side-effects. Oral administration of several osteoporotic drugs leads to poor bioavailability and drug retention in the body. However, targeted bone therapy eliminates these problems to a certain extent. This review shed light on different types of nano-based DDSs. Although these methods are well-established by in vitro and in vivo studies, further research is still needed.

Numerous advances have been made in the field of nano-based drug delivery for OP.

  1. a)

    Enhanced targeting strategies: Surface modifications, such as the use of ligands or antibodies, enable active targeting of specific cells or receptors in osteoporotic bone, improving drug delivery efficiency.

  2. b)

    Controlled and sustained release: This allows for prolonged drug exposure and therapeutic effects, reducing the frequency of drug administration.

  3. c)

    Combination therapies: The co-delivery of multiple therapeutic agents, such as anti-resorptive drugs and bone-stimulating factors, within a single nanocarrier provides synergistic effects and comprehensive treatment of osteoporosis.

  4. d)

    Advanced imaging and diagnostics: NPs with imaging capabilities have been developed, enabling non-invasive monitoring of drug distribution, bone health, and treatment response. These imaging techniques provide valuable insights for personalized treatment optimization.

  5. e)

    Smart nanocarriers: These are nanocarriers that respond to specific stimuli, such as pH changes or enzymatic activity in the bone microenvironment. These nanocarriers can be engineered to release drugs precisely at the desired sites, enhancing therapeutic effectiveness.

  6. f)

    Theranostic approaches: This integrates both therapeutic and diagnostic functionalities. These systems can simultaneously deliver therapeutic agents while providing real-time imaging and monitoring the treatment response.

  7. g)

    Biomimetic nanocarriers: These mimic the body’s inner environment, offering improved biocompatibility and facilitating targeted drug delivery to osteoporotic bone.

Conclusion and future perspectives

OP is a bone-associated disorder characterized by a loss of bone mass and increased susceptibility of fractures. Currently, the primary treatment for osteoporosis is the use of drugs to slow the rate of bone loss or promote bone formation. However, these drugs are inadequate in their performances with serious long-term side effects.

Nano-based DDSs have the potential to revolutionize the treatment of OP. These systems include the use of nanoscale particles to deliver the drug to the bone. By using this system, drugs can be delivered to specific sites within the body with increased precision and efficacy, reducing side effects and improving patient outcomes. In the future, nano-based DDSs for OP may become the preferred method for treating this condition.

In the context of translating scientific findings from basic research into practical applications and solutions, the goal is to ensure the safer administration of drugs into the body system. In pre-clinical studies, the stability, drug release kinetics, and targeting efficiency of the nano-based DDSs in cell cultures and animal models of OP are assessed. Once a promising nano-based delivery system is identified, its pharmacokinetic properties like absorption, distribution, metabolism, and excretion from the body are studied. Next, the safety and toxicity evaluation is conducted to investigate its side effects and adverse effects. The nano-based DDS is then subjected to clinical study involving experimentation on human participants with OP after which the researchers seek regulatory approval from health authorities, such as the Food and Drug Administration (FDA). After its approval, the nano-based DDS can be commercialized and made available to the patients.

Even though the different delivery systems in this review have been well researched, certain challenges still persist for their clinical translation. Some of those are mentioned below.

  1. a)

    Safety concerns: NPs should be thoroughly evaluated for their potential toxicity, biocompatibility, and long-term effects on various organs or tissues.

  2. b)

    Scale-up and manufacturing: Scaling up the production of nano-based DDSs while maintaining quality control and reproducibility can be challenging. Cost-effective manufacturing processes need to be developed for large-scale production and commercialization.

  3. c)

    Targeting specificity: Achieving precise targeting of NPs to bone tissues while minimizing off-target effects remains a challenge. Improving the specificity and selectivity of NP accumulation in osteoporotic bone is crucial for optimizing therapeutic outcomes.

  4. d)

    Pharmacokinetics and biodistribution: Understanding the pharmacokinetics, clearing mechanisms, biodegradation pathways, and biodistribution of NPs in the body is important for determining the optimal dosing regimen and predicting potential side effects.

  5. e)

    Clinical translation: Bridging the gap between preclinical studies and clinical translation is a significant challenge. Comprehensive and well-designed clinical trials are needed to assess the safety, efficacy, and long-term effects of nano-based delivery in human patients.

  6. f)

    Immunogenicity and immune response: Immunological reactions may impact the efficacy of the delivered drugs and the overall safety profile of the nano-based delivery system.

  7. g)

    Regulatory considerations: Complying with regulatory requirements and obtaining necessary approvals for nano-based DDSs is a crucial step in their clinical translation. Meeting regulatory standards for safety, efficacy, and quality control for their successful implementation in patient care is necessary.

By addressing these challenges, nano-based drug delivery systems for OP can contribute to therapeutic advancements, offering more targeted, safe, and effective treatments for patients. Some of the possible suggestions are listed below.

  1. a)

    Improved bone targeting: Further research is needed to enhance the specificity and selectivity of nanocarriers for bone targeting. Exploring new targeting ligands or receptors specific to osteoporotic bone can improve the accumulation of NPs at the desired sites.

  2. b)

    Combination nanotherapies: Investigating synergistic combinations of NPs, including drugs, gene therapies, or stem cell therapies, can lead to more efficient treatment approaches by targeting multiple aspects of OP.

  3. c)

    Personalized medicine: Integrating personalized medicine approaches, such as genetic profiling or individualized bone quality assessments, can guide the optimization of nano-based DDSs. Tailoring treatments to individual patient characteristics can improve treatment efficacy and minimize adverse effects.

  4. d)

    Clinical translation and validation: Conducting well-designed clinical trials is essential for validating the safety, efficacy, and long-term effects of the delivery system in human patients. Close collaboration between researchers, clinicians, and regulatory agencies is necessary for successful clinical translation.

  5. e)

    Patient compliance and convenience: Developing delivery systems like long-acting implants or transdermal patches can enhance treatment adherence and simplify administration procedures.

  6. f)

    Bioinformatics and computational modeling: Utilizing these models to predict and optimize the behavior of nano-based delivery systems can guide the design process and accelerate the development of effective therapies.

In conclusion, nano-based DDSs hold great promise for addressing the challenges associated with OP treatment. The unique properties of nanomaterials, such as their small size, controlled release capabilities, and targeted delivery potential, offer exciting opportunities for enhancing therapeutic outcomes in this debilitating condition. By utilizing nanotechnology, it is possible to improve the bioavailability, stability, and efficacy of therapeutic agents for OP. Moreover, nano DDSs have the potential to minimize off-target effects, reduce systemic toxicity, and enable personalized medicine approaches. Even though there are still many limitations associated with nano-based DDSs, the solution can be achieved by more innovative designs and extensive studies, as highlighted in Fig. 7. While further research and development are needed, the advancements in nano-based drug delivery for OP offer a hopeful path towards more effective treatments and improved quality of life for patients.

Fig. 7
figure 7

Limitations of nano-based drug delivery systems and their potential solutions. Nano-based drug delivery systems have shown immense potential in revolutionizing drug delivery for various medical conditions, including osteoporosis. However, they are not without their limitations such as safety concerns and immune responses. It is crucial to identify these limitations and explore potential solutions to overcome them for the successful implementation of nano-based DDSs. NPs-Nanoparticles

Data Availability

Not applicable.

Abbreviations

ALG:

Alginate

Aln:

Alendronate

ALP:

Alkaline phosphatase

ASP:

Aspartic oligopeptide

ATP2A2:

Sarcoplasmic/endoplasmic reticulum calcium ATPase2

BP:

Bisphosphonate

BMD:

Bone mineral density

BM-MSCs:

Bone marrow-derived mesenchymal stem cells

BMP:

Bone morphogenetic protein-2

CD:

β-cyclodextrin

Ce:

Ceria

CGRP:

Calcitonin gene-related peptide

CRHNPs:

Calcium-rich hydroxyapatite nanoparticles

CUR:

Curcumin

CXCR4:

Chemokine receptor type 4

DDSs:

Drug delivery systems

DNA:

Deoxyribonucleic acid

DO:

Diabetic osteoporosis

E2:

17β estradiol

EPR:

Enhanced permeation and retention

EVs:

Extracellular vesicles

FDPS:

Farnesyl diphosphate synthase

FGF23:

Fibroblast growth factor 23

GI:

Gastrointestinal tract

HA:

Hydroxyapatite

HMEC:

Human microvascular endothelial cells

HNLZ:

Hydroxyapatite nanoparticles loaded with zoledronic acid

ICA:

Icariin

lncRNAs:

Long non-coding RNAs

MBGNPs:

Mesoporous bioactive glass nanoparticles

MBGS:

Mesoporous bioactive glass scaffold

MNP:

Magnetic nanoparticle

MSC:

Mesenchymal stem cell

MYB:

Myeloblastosis

NCT:

Nanochitosan

NPs:

Nanoparticles

OB:

Osteoblast

OC:

Osteoclast

OP:

Osteoporosis

OVX:

Ovarectomized

PAA:

Poly allylamine

PEG:

Polyethyelene glycol

PEI:

Polyethylenimine

PLGANPs:

Poly lactic-co-glycolic acid

QSLNs:

Quercetin-based solid lipid nanoparticles

Ral:

Raloxifene

RANKL:

Receptor activator of nuclear factor kappa-B ligand

RhBMP-2:

Recombinant human bone morphogenetic protein-2

RNA:

Ribonucleic acid

SCT:

Salmon calcitonin

SDF-1:

Stromal cell-derived factor 1

SF:

Silk fibroin

SIM:

Simvastatin

SPIO@HA:

Hydroxyapatite-coated superparamagnetic iron oxide

Sr:

Strontium

SWE:

Shilajit water extract

Tb.N:

Trabecular number

Tb.Sp:

Trabecular separation

TC:

Tetracycline

TEOS:

Tetraethoxysilane solution

TiO2 :

Titanium dioxide

TNT:

Titanium dioxide nanotubes

TRAP:

Tartrate-resistant acid phosphatase

VGNPs:

Vitamin-D-conjugated gold nanoparticles

ZOL:

Zoledronic acid

References

  1. Balogh E, Paragh G, Jeney V. Influence of iron on bone homeostasis. Pharmaceuticals. 2018;11(4):107.

    Article  CAS  Google Scholar 

  2. Rodan GA. Introduction to bone biology. Volume 13. Bone; 1992. pp. S3–S6.

  3. Nakahama KI. Cellular communications in bone homeostasis and repair. Cell Mol Life Sci. 2010;67:4001–9.

    Article  CAS  Google Scholar 

  4. Katsumi H, Yamashita S, Morishita M, Yamamoto A. Bone-targeted drug delivery systems and strategies for treatment of bone metastasis. Chem Pharm Bull. 2020;68(7):560–6.

    Article  CAS  Google Scholar 

  5. Gu W, Wu C, Chen J, Xiao Y. 2013. Nanotechnology in the targeted drug delivery for bone diseases and bone regeneration. Int J Nanomed, pp.2305–17.

  6. Pouresmaeili F, Kamalidehghan B, Kamarehei M, Goh YM. A comprehensive overview on osteoporosis and its risk factors. Therapeutics and clinical risk management; 2018. pp. 2029–49.

  7. Park YS, Lee JY, Suh JS, Jin YM, Yu Y, Kim HY, Park YJ, Chung CP, Jo I. Selective osteogenesis by a synthetic mineral inducing peptide for the treatment of osteoporosis. Biomaterials. 2014;35(37):9747–54.

    Article  CAS  Google Scholar 

  8. Johnston CB, Dagar M. Osteoporosis in older adults. Med Clin. 2020;104(5):873–84.

    Google Scholar 

  9. Wright NC, Looker AC, Saag KG, Curtis JR, Delzell ES, Randall S, Dawson-Hughes B. The recent prevalence of osteoporosis and low bone mass in the United States based on bone mineral density at the femoral neck or lumbar spine. J Bone Miner Res. 2014;29(11):2520–6.

    Article  Google Scholar 

  10. Hoque J, Shih YRV, Zeng Y, Newman H, Sangaj N, Arjunji N, Varghese S. 2021. Bone targeting nanocarrier-assisted delivery of adenosine to combat osteoporotic bone loss. Biomaterials, 273, p.120819.

  11. Li L, Wang Z. 2018. Ovarian aging and osteoporosis. Aging and aging-related diseases: mechanisms and interventions, pp.199–215.

  12. Ahn TK, Kim KT, Joshi HP, Park KH, Kyung JW, Choi UY, Sohn S, Sheen SH, Shin DE, Lee SH, Han IB. Therapeutic potential of tauroursodeoxycholic acid for the treatment of osteoporosis. Int J Mol Sci. 2020;21(12):4274.

    Article  CAS  Google Scholar 

  13. Cheng H, Chawla A, Yang Y, Li Y, Zhang J, Jang HL, Khademhosseini A. Development of nanomaterials for bone-targeted drug delivery. Drug Discovery Today. 2017;22(9):1336–50.

    Article  CAS  Google Scholar 

  14. Lin Y, Villacanas MG, Zou H, Liu H, Carcedo IG, Wu Y, Sun B, Wu X, Prasadam I, Monteiro MJ, Li L. Calcium-bisphosphonate nanoparticle platform as a prolonged nanodrug and bone-targeted delivery system for bone diseases and cancers. ACS Appl Bio Mater. 2021;4(3):2490–501.

    Article  CAS  Google Scholar 

  15. Rajani R, Schaefer L, Scarborough MT, Gibbs CP. Giant cell tumors of the foot and ankle bones: high recurrence rates after surgical treatment. J Foot Ankle Surg. 2015;54(6):1141–5.

    Article  Google Scholar 

  16. Baines CR, McGuiness W, O’Rourke GA. An integrative review of skin assessment tools used to evaluate skin injury related to external beam radiation therapy. J Clin Nurs. 2017;26(7–8):1137–44.

    Article  Google Scholar 

  17. Mura S, Nicolas J, Couvreur P. Stimuli-responsive nanocarriers for drug delivery. Nat Mater. 2013;12(11):991–1003.

    Article  CAS  Google Scholar 

  18. Figueiredo A, Silva O, Cabrita S. 2012. Inflammatory reaction post implantation of bone graft materials. Exp Pathol Health Sci, 6(1), pp.15 – 8.

  19. Lems WF, Raterman HG. Critical issues and current challenges in osteoporosis and fracture prevention. An overview of unmet needs. Therapeutic Adv Musculoskelet Disease. 2017;9(12):299–316.

    Article  Google Scholar 

  20. Bhattacharyya A, Janarthanan G, Kim T, Taheri S, Shin J, Kim J, Bae HC, Han HS, Noh I. Modulation of bioactive calcium phosphate micro/nanoparticle size and shape during in situ synthesis of photo-crosslinkable gelatin methacryloyl based nanocomposite hydrogels for 3D bioprinting and tissue engineering. Biomaterials Res. 2022;26(1):1–17.

    Article  Google Scholar 

  21. Zhang Z, Zhao J, Chen Z, Wu H, Wang S. A molybdenum-based nanoplatform with multienzyme mimicking capacities for oxidative stress-induced acute liver injury treatment. Inorganic Chemistry Frontiers; 2023.

  22. Hoffman AS. The origins and evolution of “controlled” drug delivery systems. J Controlled Release. 2008;132(3):153–63.

    Article  CAS  Google Scholar 

  23. Davis FF. The origin of pegnology. Adv Drug Deliv Rev. 2002;54(4):457–8.

    Article  CAS  Google Scholar 

  24. Ruoslahti E. The RGD story: a personal account. Matrix biology: journal of the International Society for Matrix Biology. 2003;22(6):459–65.

    Article  CAS  Google Scholar 

  25. Dai L, Liu KF, Si CL, He J, Lei JD, Guo LQ. A novel self-assembled targeted nanoparticle platform based on carboxymethylcellulose co-delivery of anticancer drugs. J Mater Chem B. 2015;3(32):6605–17.

    Article  CAS  Google Scholar 

  26. Nguyen TBL, Min YK, Lee BT. Nanoparticle biphasic calcium phosphate loading on gelatin-pectin scaffold for improved bone regeneration. Tissue Eng Part A. 2015;21(7–8):1376–87.

    Article  CAS  Google Scholar 

  27. Ma X, Gong N, Zhong L, Sun J, Liang XJ. Future of nanotherapeutics: targeting the cellular sub-organelles. Biomaterials. 2016;97:10–21.

    Article  CAS  Google Scholar 

  28. Parent M, Baradari H, Champion E, Damia C, Viana-Trecant M. Design of calcium phosphate ceramics for drug delivery applications in bone diseases: a review of the parameters affecting the loading and release of the therapeutic substance. J Controlled Release. 2017;252:1–17.

    Article  CAS  Google Scholar 

  29. Melville AJ, Rodríguez-Lorenzo LM, Forsythe JS. Effects of calcination temperature on the drug delivery behaviour of Ibuprofen from hydroxyapatite powders. J Mater Science: Mater Med. 2008;19:1187–95.

    CAS  Google Scholar 

  30. Matsumoto T, Okazaki M, Inoue M, Yamaguchi S, Kusunose T, Toyonaga T, Hamada Y, Takahashi J. Hydroxyapatite particles as a controlled release carrier of protein. Biomaterials. 2004;25(17):3807–12.

    Article  CAS  Google Scholar 

  31. Lin K, Wu C, Chang J. Advances in synthesis of calcium phosphate crystals with controlled size and shape. Acta Biomater. 2014;10(10):4071–102.

    Article  CAS  Google Scholar 

  32. Uskokovic V, Batarni SS, Schweicher J, King A, Desai TA. Effect of calcium phosphate particle shape and size on their antibacterial and osteogenic activity in the delivery of antibiotics in vitro. ACS Appl Mater Interfaces. 2013;5(7):2422–31.

    Article  CAS  Google Scholar 

  33. Feng B, Weng J, Yang BC, Qu SX, Zhang XD. Characterization of surface oxide films on titanium and adhesion of osteoblast. Biomaterials. 2003;24(25):4663–70.

    Article  CAS  Google Scholar 

  34. Patra JK, Das G, Fraceto LF, Campos EVR, Rodriguez-Torres MDP, Acosta-Torres LS, Diaz-Torres LA, Grillo R, Swamy MK, Sharma S, Habtemariam S. Nano based drug delivery systems: recent developments and future prospects. J Nanobiotechnol. 2018;16(1):1–33.

    Article  Google Scholar 

  35. Dai L, Liu KF, Si CL, He J, Lei JD, Guo LQ. A novel self-assembled targeted nanoparticle platform based on carboxymethylcellulose co-delivery of anticancer drugs. J Mater Chem B. 2015;3(32):6605–17.

    Article  CAS  Google Scholar 

  36. Liechty WB, Kryscio DR, Slaughter BV, Peppas NA, Reis C. CP. Recent advances in drug delivery systems. J Biomater Nanobiotechnol. 2011;2:510.

  37. Ahn GY, Kim SE, Yun TH, Choi I, Park D, Choi SW. Enhanced osteogenic differentiation of alendronate-conjugated nanodiamonds for potential osteoporosis treatment. Biomaterials Res. 2021;25:1–11.

    Article  Google Scholar 

  38. Martinho N, Damgé C, Reis CP. Recent advances in drug delivery systems. J biomaterials Nanobiotechnol. 2011;2(05):510.

    Article  CAS  Google Scholar 

  39. Gan L, Wang J, Jiang M, Bartlett H, Ouyang D, Eperjesi F, Liu J, Gan Y. Recent advances in topical ophthalmic drug delivery with lipid-based nanocarriers. Drug Discovery Today. 2013;18(5–6):290–7.

    Article  CAS  Google Scholar 

  40. Xiao H, Li C, Dai Y, Cheng Z, Hou Z, Lin J. Inorganic nanocarriers for platinum drug delivery. Mater Today. 2015;18(10):554–64.

    Article  Google Scholar 

  41. Emami F, Yazdi M, S.J. and, Na DH. Poly (lactic acid)/poly (lactic-co-glycolic acid) particulate carriers for pulmonary drug delivery. J Pharm Invest. 2019;49:427–42.

    Article  CAS  Google Scholar 

  42. Lam PL, Wong WY, Bian Z, Chui CH, Gambari R. Recent advances in green nanoparticulate systems for drug delivery: efficient delivery and safety concern. Nanomedicine. 2017;12(4):357–85.

    Article  CAS  Google Scholar 

  43. Jeong J, Shim JH, Koo BM, Choy YB, Heo CY. Synergistic effect of Whitlockite Scaffolds Combined with Alendronate to promote bone regeneration. Tissue Eng Regenerative Med. 2022;19(1):83–92.

    Article  CAS  Google Scholar 

  44. Leutner M, Matzhold C, Bellach L, Deischinger C, Harreiter J, Thurner S, Klimek P, Kautzky-Willer A. Diagnosis of osteoporosis in statin-treated patients is dose-dependent. Ann Rheum Dis. 2019;78(12):1706–11.

    Article  CAS  Google Scholar 

  45. Borciani G, Ciapetti G, Vitale-Brovarone C, Baldini N. 2022. Strontium functionalization of biomaterials for bone tissue engineering purposes: a biological point of view. Materials, 15(5), p.1724.

  46. Lee NH, Kang MS, Kim TH, Yoon DS, Mandakhbayar N, Jo SB, Kim HS, Knowles JC, Lee JH, Kim HW. 2021. Dual actions of osteoclastic-inhibition and osteogenic-stimulation through strontium-releasing bioactive nanoscale cement imply biomaterial-enabled osteoporosis therapy. Biomaterials, 276, p.121025.

  47. Dou C, Li J, He J, Luo F, Yu T, Dai Q, Chen Y, Xu J, Yang X, Dong S. Bone-targeted pH-responsive cerium nanoparticles for anabolic therapy in osteoporosis. Bioactive Mater. 2021;6(12):4697–706.

    Article  CAS  Google Scholar 

  48. Kou Y, Li C, Yang P, Li D, Lu X, Liu H, Li M. 2021. The W9 peptide inhibits osteoclastogenesis and osteoclast activity by downregulating osteoclast autophagy and promoting osteoclast apoptosis. J Mol Histol, pp.1–12.

  49. Liang W, Zhuo X, Tang Z, Wei X, Li B. Calcitonin gene-related peptide stimulates proliferation and osteogenic differentiation of osteoporotic rat-derived bone mesenchymal stem cells. Mol Cell Biochem. 2015;402:101–10.

    Article  CAS  Google Scholar 

  50. Rey JRC, Cervino EV, Rentero ML, Crespo EC, Álvaro AO, Casillas M. Raloxifene: mechanism of action, effects on bone tissue, and applicability in clinical traumatology practice. open Orthop J. 2009;3:14.

    Article  Google Scholar 

  51. Lewiecki EM. Bisphosphonates for the treatment of osteoporosis: insights for clinicians. Therapeutic Adv chronic disease. 2010;1(3):115–28.

    Article  CAS  Google Scholar 

  52. Erlacher L, Kettenbach J, Kiener H, Graninger W, Kainberger F, Pietschmann P. Salmon calcitonin and calcium in the treatment of male osteoporosis: the effect on bone mineral density. Wiener klinische Wochenschrift. 1997;109(8):270–4.

    CAS  Google Scholar 

  53. Onuora S. Gene therapy counteracts bone loss in osteoporosis. Nat Rev Rheumatol. 2019;15(9):513–3.

    Article  Google Scholar 

  54. Manolagas SC, O’brien CA, Almeida M. 2013. The role of estrogen and androgen receptors in bone health and disease. Nature Reviews Endocrinology, 9(12), p.699.

  55. Hatefi M, Ahmadi MRH, Rahmani A, Dastjerdi MM, Asadollahi K. Effects of curcumin on bone loss and biochemical markers of bone turnover in patients with spinal cord injury. Volume 114. World neurosurgery; 2018. pp. e785–91.

  56. Wang Z, Wang D, Yang D, Zhen W, Zhang J, Peng S. The effect of icariin on bone metabolism and its potential clinical application. Volume 29. Osteoporosis International; 2018. pp. 535–44.

  57. Wong SK, Chin KY, Ima-Nirwana S. 2020. Quercetin as an agent for protecting the bone: a review of the current evidence. International Journal of Molecular Sciences, 21(17), p.6448.

  58. Chapurlat RD. Odanacatib: a review of its potential in the management of osteoporosis in postmenopausal women. Therapeutic Adv Musculoskelet disease. 2015;7(3):103–9.

    Article  CAS  Google Scholar 

  59. Ehrlich PJ, Lanyon LE. 2002. Mechanical strain and bone cell function: a review. Osteoporosis international, 13(9), p.688.

  60. Parfitt AM. Trabecular bone architecture in the pathogenesis and prevention of fracture. Am J Med. 1987;82(1):68–72.

    Article  CAS  Google Scholar 

  61. Karunaratne DN. 2007. Nanotechnology in medicine. J Natl Sci Foundation Sri Lanka, 35(3).

  62. Yan X, Yu C, Zhou X, Tang J, Zhao D. Highly ordered mesoporous bioactive glasses with superior in vitro bone-forming bioactivities. Angew Chem Int Ed. 2004;43(44):5980–4.

    Article  CAS  Google Scholar 

  63. Mora-Raimundo P, Lozano D, Benito M, Mulero F, Manzano M, Vallet‐Regí M. 2021. Osteoporosis remission and new bone formation with mesoporous silica nanoparticles. Advanced Science, 8(16), p.2101107.

  64. Xu C, Lei C, Huang L, Zhang J, Zhang H, Song H, Yu M, Wu Y, Chen C, Yu C. Glucose-responsive nanosystem mimicking the physiological insulin secretion via an enzyme–polymer layer-by-layer coating strategy. Chem Mater. 2017;29(18):7725–32.

    Article  CAS  Google Scholar 

  65. Xu C, Lei C, Yu C. Mesoporous silica nanoparticles for protein protection and delivery. Front Chem. 2019;7:290.

    Article  CAS  Google Scholar 

  66. Gibson IR, Best SM, Bonfield W. Chemical characterization of silicon-substituted hydroxyapatite. J Biomedical Mater Research: Official J Soc Biomaterials Japanese Soc Biomaterials Australian Soc Biomaterials. 1999;44(4):422–8.

    Article  CAS  Google Scholar 

  67. Keeting PE, Oursler MJ, Wiegand KE, Bonde SK, Spelsberg TC, Riggs BL. Zeolite a increases proliferation, differentiation, and transforming growth factor β production in normal adult human osteoblast-like cells in vitro. J Bone Miner Res. 1992;7(11):1281–9.

    Article  CAS  Google Scholar 

  68. Rahman IA, Padavettan V. 2012. Synthesis of silica nanoparticles by sol-gel: size-dependent properties, surface modification, and applications in silica-polymer nanocomposites—a review. Journal of nanomaterials, 2012, pp.8–8.

  69. Beck Jr GR, Ha SW, Camalier CE, Yamaguchi M, Li Y, Lee JK, Weitzmann MN. 2012. Bioactive silica-based nanoparticles stimulate bone-forming osteoblasts, suppress bone-resorbing osteoclasts, and enhance bone mineral density in vivo. Nanomedicine: Nanotechnology, Biology and Medicine, 8(6), pp.793–803.

  70. Qi GB, Gao YJ, Wang L, Wang H. 2018. Self-assembled peptide‐based nanomaterials for biomedical imaging and therapy. Advanced Materials, 30(22), p.1703444.

  71. Spicer CD, Jumeaux C, Gupta B, Stevens MM. Peptide and protein nanoparticle conjugates: versatile platforms for biomedical applications. Chem Soc Rev. 2018;47(10):3574–620.

    Article  CAS  Google Scholar 

  72. Yu P, Chen Y, Wang Y, Liu Y, Zhang P, Guo Q, Li S, Xiao H, Xie J, Tan H, Li J. Pentapeptide-decorated silica nanoparticles loading salmon calcitonin for in vivo osteoporosis treatment with sustained hypocalcemic effect. Mater Today Chem. 2019;14:100189.

    Article  CAS  Google Scholar 

  73. Hench LL, Paschall HA. Direct chemical bond of bioactive glass-ceramic materials to bone and muscle. J Biomed Mater Res. 1973;7(3):25–42.

    Article  CAS  Google Scholar 

  74. Gambacciani M, Vacca F. Postmenopausal osteoporosis and hormone replacement therapy. Minerva Med. 2004;95(6):507–20.

    CAS  Google Scholar 

  75. Wang D, Steffi C, Wang Z, Kong CH, Lim PN, Shi Z, San Thian E, Wang W. Beta-cyclodextrin modified mesoporous bioactive glass nanoparticles/silk fibroin hybrid nanofibers as an implantable estradiol delivery system for the potential treatment of osteoporosis. Nanoscale. 2018;10(38):18341–53.

    Article  CAS  Google Scholar 

  76. Pinna A, Lasio B, Piccinini M, Marmiroli B, Amenitsch H, Falcaro P, Tokudome Y, Malfatti L, Innocenzi P. Combining top-down and bottom-up routes for fabrication of mesoporous titania films containing ceria nanoparticles for free radical scavenging. ACS Appl Mater Interfaces. 2013;5(8):3168–75.

    Article  CAS  Google Scholar 

  77. Tarnuzzer RW, Colon J, Patil S, Seal S. Vacancy engineered ceria nanostructures for protection from radiation-induced cellular damage. Nano Lett. 2005;5(12):2573–7.

    Article  CAS  Google Scholar 

  78. Asati A, Kaittanis C, Santra S, Perez JM. pH-tunable oxidase-like activity of cerium oxide nanoparticles achieving sensitive fluorigenic detection of cancer biomarkers at neutral pH. Anal Chem. 2011;83(7):2547–53.

    Article  CAS  Google Scholar 

  79. Pagliari F, Mandoli C, Forte G, Magnani E, Pagliari S, Nardone G, Licoccia S, Minieri M, Di Nardo P, Traversa E. Cerium oxide nanoparticles protect cardiac progenitor cells from oxidative stress. ACS Nano. 2012;6(5):3767–75.

    Article  CAS  Google Scholar 

  80. Pinna A, Malfatti L, Galleri G, Manetti R, Cossu S, Rocchitta G, Migheli R, Serra PA, Innocenzi P. Ceria nanoparticles for the treatment of Parkinson-like diseases induced by chronic manganese intoxication. RSC Adv. 2015;5(26):20432–9.

    Article  CAS  Google Scholar 

  81. Pinna A, Baghbaderani MT, Hernández VV, Naruphontjirakul P, Li S, McFarlane T, Hachim D, Stevens MM, Porter AE, Jones JR. Nanoceria provides antioxidant and osteogenic properties to mesoporous silica nanoparticles for osteoporosis treatment. Acta Biomater. 2021;122:365–76.

    Article  CAS  Google Scholar 

  82. Sun Y, Cai M, Zhong J, Yang L, Xiao J, Jin F, Xue H, Liu X, Liu H, Zhang Y, Jiang D. The long noncoding RNA lnc-ob1 facilitates bone formation by upregulating Osterix in osteoblasts. Nat Metabolism. 2019;1(4):485–96.

    Article  CAS  Google Scholar 

  83. Yang L, Li Y, Gong R, Gao M, Feng C, Liu T, Sun Y, Jin M, Wang D, Yuan Y, Yan G. The long non-coding RNA-ORLNC1 regulates bone mass by directing mesenchymal stem cell fate. Mol Ther. 2019;27(2):394–410.

    Article  CAS  Google Scholar 

  84. Yang Z, Liu X, Zhao F, Yao M, Lin Z, Yang Z, Liu C, Liu Y, Chen X, Du C. Bioactive glass nanoparticles inhibit osteoclast differentiation and osteoporotic bone loss by activating lncRNA NRON expression in the extracellular vesicles derived from bone marrow mesenchymal stem cells. Biomaterials. 2022;283:121438.

    Article  CAS  Google Scholar 

  85. Brammer KS, Frandsen CJ, Jin S. TiO2 nanotubes for bone regeneration. Trends Biotechnol. 2012;30(6):315–22.

    Article  CAS  Google Scholar 

  86. Brammer KS, Oh S, Cobb CJ, Bjursten LM, van der Heyde H, Jin S. Improved bone-forming functionality on diameter-controlled TiO2 nanotube surface. Acta Biomater. 2009;5(8):3215–23.

    Article  CAS  Google Scholar 

  87. Oh S, Daraio C, Chen LH, Pisanic TR, Finones RR, Jin S. 2006. Significantly accelerated osteoblast cell growth on aligned TiO2 nanotubes. Journal of Biomedical Materials Research Part A: an Official Journal of the Society for Biomaterials, The Japanese Society for Biomaterials, and The Australian Society for Biomaterials and the Korean Society for Biomaterials, 78(1), pp.97–103.

  88. Bjursten LM, Rasmusson L, Oh S, Smith GC, Brammer KS, Jin S. Titanium dioxide nanotubes enhance bone bonding in vivo. J Biomedical Mater Res Part A: Official J Soc Biomaterials. 2010;92(3):1218–24. The Japanese Society for Biomaterials, and The Australian Society for Biomaterials and the Korean Society for Biomaterials.

    Google Scholar 

  89. Batool SA, Salman Maqbool M, Javed MA, Niaz A, Rehman MAU. A review on the fabrication and characterization of Titania Nanotubes Obtained via Electrochemical anodization. Surfaces. 2022;5(4):456–80.

    Article  CAS  Google Scholar 

  90. Mu C, Hu Y, Huang L, Shen X, Li M, Li L, Gu H, Yu Y, Xia Z, Cai K. 2018. Sustained raloxifene release from hyaluronan-alendronate-functionalized titanium nanotube arrays capable of enhancing osseointegration in osteoporotic rabbits. Materials Science and Engineering: C, 82, pp.345–353.

  91. Zhao J, Ohba S, Komiyama Y, Shinkai M, Chung UI, Nagamune T. Icariin: a potential osteoinductive compound for bone tissue engineering. Tissue Eng Part A. 2010;16(1):233–43.

    Article  CAS  Google Scholar 

  92. Yang J, Fang K, Xu K, Shen X, Xu X. 2023. Effect of zinc or copper doping on corrosion resistance and anti-oxidative stress of strontium-based micro-arc oxidation coatings on titanium. Applied Surface Science, 626, p.157229.

  93. Zhu Y, Zheng T, Wen LM, Li R, Zhang YB, Bi WJ, Feng XJ, Qi MC. Osteogenic capability of strontium and icariin-loaded TiO2 nanotube coatings in vitro and in osteoporotic rats. J Biomater Appl. 2021;35(9):1119–31.

    Article  CAS  Google Scholar 

  94. Hsiao CY, Chen TH, Chu TH, Ting YN, Tsai PJ, Shyu JF. 2020. Calcitonin induces bone formation by increasing expression of Wnt10b in osteoclasts in ovariectomy-induced osteoporotic rats. Frontiers in Endocrinology, 11, p.613.

  95. Lai M, Yan X, Shen K, Tang Q, Fang X, Zhang C, Zhu Z, Hou Y. 2020. The effect of calcitonin gene-related peptide functionalized TiO2 nanotubes on osteoblast and osteoclast differentiation in vitro. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 600, p.124899.

  96. Hengst V, Oussoren C, Kissel T, Storm G. 2007. Bone targeting potential of bisphosphonate-targeted liposomes: Preparation, characterization and hydroxyapatite binding in vitro. International journal of pharmaceutics, 331(2), pp.224–227.

  97. Lara-Ochoa S, Ortega-Lara W, Guerrero-Beltrán CE. 2021. Hydroxyapatite nanoparticles in drug delivery: physicochemistry and applications. Pharmaceutics, 13(10), p.1642.

  98. Okada M, Furuzono T. 2012. Hydroxylapatite nanoparticles: fabrication methods and medical applications. Science and technology of advanced materials, 13(6), p.064103.

  99. Oryan A, Sahvieh S. Effects of bisphosphonates on osteoporosis: focus on zoledronate. Volume 264. Life Sciences; 2021. p. 118681.

  100. Khajuria DK, Razdan R, Mahapatra DR. 2015. Development, in vitro and in vivo characterization of zoledronic acid functionalized hydroxyapatite nanoparticle based formulation for treatment of osteoporosis in animal model. European Journal of Pharmaceutical Sciences, 66, pp.173–183.

  101. Kotak DJ, Devarajan PV. 2020. Bone targeted delivery of salmon calcitonin hydroxyapatite nanoparticles for sublingual osteoporosis therapy (SLOT). Nanomedicine: Nanotechnology, Biology and Medicine, 24, p.102153.

  102. Khajuria DK, Disha C, Vasireddi R, Razdan R, Mahapatra DR. 2016. Risedronate/zinc-hydroxyapatite based nanomedicine for osteoporosis. Materials Science and Engineering: C, 63, pp.78–87.

  103. Zhao R, Shang T, Yuan B, Zhu X, Zhang X, Yang X. Osteoporotic bone recovery by a bamboo-structured bioceramic with controlled release of hydroxyapatite nanoparticles. Bioactive Mater. 2022;17:379–93.

    Article  CAS  Google Scholar 

  104. Hwang SJ, Lee JS, Ryu TK, Kang RH, Jeong KY, Jun DR, Koh JM, Kim SE, Choi SW. Alendronate-modified hydroxyapatite nanoparticles for bone-specific dual delivery of drug and bone mineral. Macromol Res. 2016;24(7):623–8.

    Article  CAS  Google Scholar 

  105. Chen M, Lin Y, Wang H, Wang D. Function of calcium-rich hydroxyapatite nanoparticles on proliferation and differentiation of bone marrow mesenchymal stem cells (BMSCs) in osteoporosis. Mater Express. 2022;12(2):214–9.

    Article  CAS  Google Scholar 

  106. Eivazzadeh-Keihan R, Chenab KK, Taheri-Ledari R, Mosafer J, Hashemi SM, Mokhtarzadeh A, Maleki A, Hamblin MR. 2020. Recent advances in the application of mesoporous silica-based nanomaterials for bone tissue engineering. Materials Science and Engineering: C, 107, p.110267.

  107. Maleki A, Ravaghi P, Movahed H. Green approach for the synthesis of carboxycoumarins by using a highly active magnetically recyclable nanobiocomposite via sustainable catalysis. Micro & Nano Letters. 2018;13(5):591–4.

    Article  CAS  Google Scholar 

  108. Figuerola A, Di Corato R, Manna L, Pellegrino T. From iron oxide nanoparticles towards advanced iron-based inorganic materials designed for biomedical applications. Pharmacol Res. 2010;62(2):126–43.

    Article  CAS  Google Scholar 

  109. Mornet S, Vasseur S, Grasset F, Veverka P, Goglio G, Demourgues A, Portier J, Pollert E, Duguet E. Magnetic nanoparticle design for medical applications. Prog Solid State Chem. 2006;34(2–4):237–47.

    Article  CAS  Google Scholar 

  110. Hou Y, Gao S. Monodisperse nickel nanoparticles prepared from a monosurfactant system and their magnetic properties. J Mater Chem. 2003;13(7):1510–2.

    Article  CAS  Google Scholar 

  111. Sun S, Murray CB. Synthesis of monodisperse cobalt nanocrystals and their assembly into magnetic superlattices. J Appl Phys. 1999;85(8):4325–30.

    Article  CAS  Google Scholar 

  112. Ali A, Shah T, Ullah R, Zhou P, Guo M, Ovais M, Tan Z, Rui Y. 2021. Review on recent progress in magnetic nanoparticles: Synthesis, characterization, and diverse applications. Frontiers in chemistry, 9, p.629054.

  113. Lee MS, Su CM, Yeh JC, Wu PR, Tsai TY, Lou SL. 2016. Synthesis of composite magnetic nanoparticles Fe3O4 with alendronate for osteoporosis treatment. International Journal of Nanomedicine, 11, p.4583.

  114. Tran N, Webster TJ. Increased osteoblast functions in the presence of hydroxyapatite-coated iron oxide nanoparticles. Acta Biomater. 2011;7(3):1298–306.

    Article  CAS  Google Scholar 

  115. Li M, Fu S, Cai Z, Li D, Liu L, Deng D, Jin R, Ai H. Dual regulation of osteoclastogenesis and osteogenesis for osteoporosis therapy by iron oxide hydroxyapatite core/shell nanocomposites. Regenerative Biomaterials. 2021;8(5):p.rbab027.

    Article  Google Scholar 

  116. Rani S, Ryan AE, Griffin MD, Ritter T. Mesenchymal stem cell-derived extracellular vesicles: toward cell-free therapeutic applications. Mol Ther. 2015;23(5):812–23.

    Article  CAS  Google Scholar 

  117. Oliveira-Rodríguez M, Serrano-Pertierra E, García AC, López-Martín S, Yañez-Mo M, Cernuda-Morollón E, Blanco-López M. Point-of-care detection of extracellular vesicles: sensitivity optimization and multiple-target detection. Biosens Bioelectron. 2017;87:38–45.

    Article  Google Scholar 

  118. Xu C, Wang Z, Liu Y, Wei B, Liu X, Duan K, Zhou P, Xie Z, Wu M, Guan J. Extracellular vesicles derived from bone marrow mesenchymal stem cells loaded on magnetic nanoparticles delay the progression of diabetic osteoporosis via delivery of miR-150-5p. Cell Biology and Toxicology; 2022. pp. 1–18.

  119. Li H, Pan S, Xia P, Chang Y, Fu C, Kong W, Yu Z, Wang K, Yang X, Qi Z. Advances in the application of gold nanoparticles in bone tissue engineering. J Biol Eng. 2020;14:1–15.

    Article  CAS  Google Scholar 

  120. De Souza CD, Nogueira BR, Rostelato MEC. Review of the methodologies used in the synthesis gold nanoparticles by chemical reduction. J Alloys Compd. 2019;798:714–40.

    Article  Google Scholar 

  121. Tiwari PM, Vig K, Dennis VA, Singh SR. Functionalized gold nanoparticles and their biomedical applications. Nanomaterials. 2011;1(1):31–63.

    Article  CAS  Google Scholar 

  122. Vial S, Reis RL, Oliveira JM. Recent advances using gold nanoparticles as a promising multimodal tool for tissue engineering and regenerative medicine. Curr Opin Solid State Mater Sci. 2017;21(2):92–112.

    Article  CAS  Google Scholar 

  123. Vieira S, Vial S, Reis RL, Oliveira JM. Nanoparticles for bone tissue engineering. Biotechnol Prog. 2017;33(3):590–611.

    Article  CAS  Google Scholar 

  124. Sperling RA, Gil PR, Zhang F, Zanella M, Parak WJ. Biological applications of gold nanoparticles. Chem Soc Rev. 2008;37(9):1896–908.

    Article  CAS  Google Scholar 

  125. Dykman LA, Khlebtsov NG. Gold nanoparticles in biology and medicine: recent advances and prospects. Acta Naturae (англоязычная версия). 2011;3(2):34–55.

    Article  CAS  Google Scholar 

  126. Yi C, Liu D, Fong CC, Zhang J, Yang M. Gold nanoparticles promote osteogenic differentiation of mesenchymal stem cells through p38 MAPK pathway. ACS Nano. 2010;4(11):6439–48.

    Article  CAS  Google Scholar 

  127. Cramer JA, Gold DT, Silverman SL, Lewiecki EM. A systematic review of persistence and compliance with bisphosphonates for osteoporosis. Volume 18. Osteoporosis International; 2007. pp. 1023–31.

  128. Lee D, Heo DN, Kim HJ, Ko WK, Lee SJ, Heo M, Bang JB, Lee JB, Hwang DS, Do SH, Kwon IK. Inhibition of osteoclast differentiation and bone resorption by bisphosphonate-conjugated gold nanoparticles. Sci Rep. 2016;6(1):1–11.

    Google Scholar 

  129. Heo DN, Ko WK, Moon HJ, Kim HJ, Lee SJ, Lee JB, Bae MS, Yi JK, Hwang YS, Bang JB, Kim EC. Inhibition of osteoclast differentiation by gold nanoparticles functionalized with cyclodextrin curcumin complexes. ACS Nano. 2014;8(12):12049–62.

    Article  CAS  Google Scholar 

  130. Nah H, Lee D, Lee JS, Lee SJ, Heo DN, Lee YH, Bang JB, Hwang YS, Moon HJ, Kwon IK. Strategy to inhibit effective differentiation of RANKL-induced osteoclasts using vitamin D-conjugated gold nanoparticles. Appl Surf Sci. 2020;527:146765.

    Article  CAS  Google Scholar 

  131. Greene LA, Liu DX, Troy CM, Biswas SC. 2007. Cell cycle molecules define a pathway required for neuron death in development and disease. Biochimica et Biophysica Acta (BBA)-Molecular basis of Disease, 1772(4), pp.392–401.

  132. Matalová E, Buchtová M, Tucker AS, Bender TP, Janečková E, Lungová V, Balková S, Šmarda J. Expression and characterization of c-Myb in prenatal odontogenesis. Dev Growth Differ. 2011;53(6):793–803.

    Article  Google Scholar 

  133. Lungova V, Buchtova M, Janeckova E, Tucker AS, Knopfova L, Smarda J, Matalova E. Localization of c-MYB in differentiated cells during postnatal molar and alveolar bone development. Eur J Oral Sci. 2012;120(6):495–504.

    Article  Google Scholar 

  134. Takanche JS, Kim JE, Kim JS, Lee MH, Jeon JG, Park IS, Yi HK. Chitosan-gold nanoparticles mediated gene delivery of c-myb facilitates osseointegration of dental implants in ovariectomized rat. Artif Cells Nanomed Biotechnol. 2018;46(sup3):S807–17.

    Article  CAS  Google Scholar 

  135. Galindo-Rodriguez SA, Allemann E, Fessi H, Doelker E. 2005. Polymeric nanoparticles for oral delivery of drugs and vaccines: a critical evaluation of in vivo studies. Crit reviews™ therapeutic drug carrier Syst, 22(5).

  136. Tosi G, Bortot B, Ruozi B, Dolcetta D, Vandelli MA, Forni F, Severini GM. Potential use of polymeric nanoparticles for drug delivery across the blood-brain barrier. Curr Med Chem. 2013;20(17):2212–25.

    Article  CAS  Google Scholar 

  137. Danhier F, Ansorena E, Silva JM, Coco R, Le Breton A, Préat V. PLGA-based nanoparticles: an overview of biomedical applications. J Controlled Release. 2012;161(2):505–22.

    Article  CAS  Google Scholar 

  138. Soppimath KS, Aminabhavi TM, Kulkarni AR, Rudzinski WE. Biodegradable polymeric nanoparticles as drug delivery devices. J Controlled Release. 2001;70(1–2):1–20.

    Article  CAS  Google Scholar 

  139. Hernández-Giottonini KY, Rodríguez-Córdova RJ, Gutiérrez-Valenzuela CA, Peñuñuri-Miranda O, Zavala-Rivera P, Guerrero-Germán P, Lucero-Acuña A. PLGA nanoparticle preparations by emulsification and nanoprecipitation techniques: Effects of formulation parameters. RSC Adv. 2020;10(8):4218–31.

    Article  Google Scholar 

  140. Xie Y, Liu C, Huang H, Huang J, Deng A, Zou P, Tan X. Bone-targeted delivery of simvastatin-loaded PEG-PLGA micelles conjugated with tetracycline for osteoporosis treatment. Drug Delivery and Translational Research. 2018;8(5):1090–102.

    Article  CAS  Google Scholar 

  141. Sun X, Guo Q, Wei W, Robertson S, Yuan Y, Luo X. 2019. Current progress on microRNA-based gene delivery in the treatment of osteoporosis and osteoporotic fracture. International journal of endocrinology, 2019.

  142. Sezlev Bilecen D, Uludag H, Hasirci V. Development of PEI-RANK siRNA complex loaded PLGA nanocapsules for the treatment of osteoporosis. Tissue Eng Part A. 2019;25(1–2):34–43.

    Article  CAS  Google Scholar 

  143. Vandenput L, Boonen S, Van Herck E, Swinnen JV, Bouillon R, Vanderschueren D. Evidence from the aged orchidectomized male rat model that 17β-estradiol is a more effective bone‐sparing and anabolic agent than 5α‐dihydrotestosterone. J Bone Miner Res. 2002;17(11):2080–6.

    Article  CAS  Google Scholar 

  144. Takeuchi I, Kobayashi S, Hida Y, Makino K. Estradiol-loaded PLGA nanoparticles for improving low bone mineral density of cancellous bone caused by osteoporosis: application of enhanced charged nanoparticles with iontophoresis. Colloids Surf B. 2017;155:35–40.

    Article  CAS  Google Scholar 

  145. Ding D, Zhu Z, Liu Q, Wang J, Hu Y, Jiang X, Liu B. Cisplatin-loaded gelatin-poly (acrylic acid) nanoparticles: synthesis, antitumor efficiency in vivo and penetration in tumors. Eur J Pharm Biopharm. 2011;79(1):142–9.

    Article  CAS  Google Scholar 

  146. Song J, Leeuwenburgh SC. Sustained delivery of biomolecules from gelatin carriers for applications in bone regeneration. Therapeutic Delivery. 2014;5(8):943–58.

    Article  CAS  Google Scholar 

  147. Bello AB, Kim D, Kim D, Park H, Lee SH. Engineering and functionalization of gelatin biomaterials: from cell culture to medical applications. Tissue Eng Part B: Reviews. 2020;26(2):164–80.

    Article  CAS  Google Scholar 

  148. Yasmin R, Shah M, Khan SA, Ali R. Gelatin nanoparticles: a potential candidate for medical applications. Nanatechnol Reviews. 2017;6(2):191–207.

    Article  CAS  Google Scholar 

  149. Yang XJ, Wang FQ, Lu CB, Zou JW, Hu JB, Yang Z, Sang HX, Zhang Y. 2020. Modulation of bone formation and resorption using a novel zoledronic acid loaded gelatin nanoparticles integrated porous titanium scaffold: an in vitro and in vivo study. Biomedical Materials, 15(5), p.055013.

  150. Hankenson KD, Dishowitz M, Gray C, Schenker M. Angiogenesis in bone regeneration. Injury. 2011;42(6):556–61.

    Article  Google Scholar 

  151. Reginster JY. Cardiac concerns associated with strontium ranelate. Exp Opin Drug Saf. 2014;13(9):1209–13.

    Article  CAS  Google Scholar 

  152. D’haese PC, Schrooten I, Goodman WG, Cabrera WE, Lamberts LV, Elseviers MM, Couttenye MM, De Broe ME. Increased bone strontium levels in hemodialysis patients with osteomalacia. Kidney Int. 2000;57(3):1107–14.

    Article  Google Scholar 

  153. Li D, Chen K, Duan L, Fu T, Li J, Mu Z, Wang S, Zou Q, Chen L, Feng Y, Li Y. Strontium ranelate incorporated enzyme-cross-linked gelatin nanoparticle/silk fibroin aerogel for osteogenesis in OVX-induced osteoporosis. ACS Biomaterials Science & Engineering. 2019;5(3):1440–51.

    Article  CAS  Google Scholar 

  154. Sivanesan I, Gopal J, Muthu M, Shin J, Mari S, Oh J. 2021. Green Synthesized Chitosan/Chitosan Nanoforms/Nanocomposites for Drug Delivery Applications. Polymers, 13(14), p.2256.

  155. Mohammed MA, Syeda JT, Wasan KM, Wasan EK. 2017. An overview of chitosan nanoparticles and its application in non-parenteral drug delivery. Pharmaceutics, 9(4), p.53.

  156. Yanat M, Schroën K. Preparation methods and applications of chitosan nanoparticles; with an outlook toward reinforcement of biodegradable packaging. Reactive and Functional Polymers. 2021;161:104849.

    Article  CAS  Google Scholar 

  157. Alshubaily FA, Jambi EJ. 2022. Correlation between Antioxidant and Anti-Osteoporotic Activities of Shilajit Loaded into Chitosan Nanoparticles and Their Effects on Osteoporosis in Rats. Polymers, 14(19), p.3972.

  158. Santhosh S, Mukherjee D, Anbu J, Murahari M, Teja BV. Improved treatment efficacy of risedronate functionalized chitosan nanoparticles in osteoporosis: formulation development, in vivo, and molecular modelling studies. J Microencapsul. 2019;36(4):338–55.

    Article  Google Scholar 

  159. Chen X, Wang S, Zhang X, Yu Y, Wang J, Liu C. 2022. Dual-function injectable fibrin gel incorporated with sulfated chitosan nanoparticles for rhBMP-2-induced bone regeneration. Applied Materials Today, 26, p.101347.

  160. Poynton AR, Lane JM. Safety profile for the clinical use of bone morphogenetic proteins in the spine. Spine. 2002;27(16S):S40–8.

    Article  Google Scholar 

  161. Chen X, Wang S, Zhang X, Yu Y, Wang J, Liu C. 2022. Dual-function injectable fibrin gel incorporated with sulfated chitosan nanoparticles for rhBMP-2-induced bone regeneration. Applied Materials Today, 26, p.101347.

  162. Bone HG, Hosking D, Devogelaer J-P, Tucci JR, Emkey RD, Tonino RP. Ten years’ experience with alendronate for osteoporosis in postmenopausal women. N Engl J Med. 2004;350(12):1189–99. Jose Adolfo Rodriguez-Portales.

  163. Mauri E, Giannitelli SM, Trombetta M, Rainer A. 2021. Synthesis of nanogels: Current trends and future outlook. Gels, 7(2), p.36.

  164. Matsumoto NM, González-Toro DC, Chacko RT, Maynard HD, Thayumanavan S. Synthesis of nanogel–protein conjugates. Polym Chem. 2013;4(8):2464–9.

    Article  CAS  Google Scholar 

  165. Kaur R, Ajitha M. Formulation of transdermal nanoemulsion gel drug delivery system of lovastatin and its in vivo characterization in glucocorticoid induced osteoporosis rat model. J Drug Deliv Sci Technol. 2019;52:968–78.

    Article  CAS  Google Scholar 

  166. Zhang Q, Chen X, Geng S, Wei L, Miron RJ, Zhao Y, Zhang Y. 2017. Nanogel-based scaffolds fabricated for bone regeneration with mesoporous bioactive glass and strontium: in vitro and in vivo characterization. Journal of Biomedical Materials Research Part A, 105(4), pp.1175–1183.

  167. Alles N, Soysa NS, Hussain MA, Tomomatsu N, Saito H, Baron R, Morimoto N, Aoki K, Akiyoshi K, Ohya K. Polysaccharide nanogel delivery of a TNF-α and RANKL antagonist peptide allows systemic prevention of bone loss. Eur J Pharm Sci. 2009;37(2):83–8.

    Article  CAS  Google Scholar 

  168. Guo Z, Qi P, Pei D, Zhang X. 2022. Raloxifene-loaded solid lipid nanoparticles decorated gel with enhanced treatment potential of osteoporosis. J Drug Deliv Sci Technol, 75,p.103733.

  169. Yu S, Ding J, He C, Cao Y, Xu W, Chen X. Disulfide cross-linked polyurethane micelles as a reduction‐triggered drug delivery system for cancer therapy. Adv Healthc Mater. 2014;3(5):752–60.

    Article  CAS  Google Scholar 

  170. Song N, Ding M, Pan Z, Li J, Zhou L, Tan H, Fu Q. Construction of targeting-clickable and tumor-cleavable polyurethane nanomicelles for multifunctional intracellular drug delivery. Biomacromolecules. 2013;14(12):4407–19.

    Article  CAS  Google Scholar 

  171. Guan Y, Su Y, Zhao L, Meng F, Wang Q, Yao Y, Luo J. 2017. Biodegradable polyurethane micelles with pH and reduction responsive properties for intracellular drug delivery. Materials Science and Engineering: C, 75, pp.1221–1230.

  172. Cai M, Yang L, Zhang S, Liu J, Sun Y, Wang X. A bone-resorption surface-targeting nanoparticle to deliver anti-miR214 for osteoporosis therapy. Int J Nanomed. 2017;12:7469.

    Article  CAS  Google Scholar 

  173. Sun Y, Ye X, Cai M, Liu X, Xiao J, Zhang C, Wang Y, Yang L, Liu J, Li S, Kang C. Osteoblast-targeting-peptide modified nanoparticle for siRNA/microRNA delivery. ACS Nano. 2016;10(6):5759–68.

    Article  CAS  Google Scholar 

  174. zur Mühlen A, Schwarz C, Mehnert W. Solid lipid nanoparticles (SLN) for controlled drug delivery–drug release and release mechanism. Eur J Pharm Biopharm. 1998;45(2):149–55.

    Article  Google Scholar 

  175. Luo Y, Chen D, Ren L, Zhao X, Qin J. Solid lipid nanoparticles for enhancing vinpocetine’s oral bioavailability. J Controlled Release. 2006;114(1):53–9.

    Article  CAS  Google Scholar 

  176. Manjunath K, Venkateswarlu V. Pharmacokinetics, tissue distribution and bioavailability of clozapine solid lipid nanoparticles after intravenous and intraduodenal administration. J Controlled Release. 2005;107(2):215–28.

    Article  CAS  Google Scholar 

  177. Chattopadhyay P, Shekunov BY, Yim D, Cipolla D, Boyd B, Farr S. Production of solid lipid nanoparticle suspensions using supercritical fluid extraction of emulsions (SFEE) for pulmonary delivery using the AERx system. Adv Drug Deliv Rev. 2007;59(6):444–53.

    Article  CAS  Google Scholar 

  178. Huang LE, Wang X, Cao H, Li L, Chow DHK, Tian L, Wu H, Zhang J, Wang N, Zheng L, Yao X. A bone-targeting delivery system carrying osteogenic phytomolecule icaritin prevents osteoporosis in mice. Biomaterials. 2018;182:58–71.

    Article  Google Scholar 

  179. Ganesan P, Narayanasamy D. Lipid nanoparticles: different preparation techniques, characterization, hurdles, and strategies for the production of solid lipid nanoparticles and nanostructured lipid carriers for oral drug delivery. Sustainable Chem Pharm. 2017;6:37–56.

    Article  Google Scholar 

  180. Ahmad N, Banala VT, Kushwaha P, Karvande A, Sharma S, Tripathi AK, Verma A, Trivedi R, Mishra PR. Quercetin-loaded solid lipid nanoparticles improve osteoprotective activity in an ovariectomized rat model: a preventive strategy for post-menopausal osteoporosis. RSC Adv. 2016;6(100):97613–28.

    Article  CAS  Google Scholar 

  181. Mundy G, Garrett R, Harris S, Chan J, Chen D, Rossini G, Boyce B, Zhao M, Gutierrez G. Stimulation of bone formation in vitro and in rodents by statins. Science. 1999;286(5446):1946–9.

    Article  CAS  Google Scholar 

  182. Tao S, Chen SQ, Zhou WT, Yu FY, Bao L, Qiu GX, Qiao Q, Hu FQ, Wang JW, Yuan H. A novel biocompatible, simvastatin-loaded, bone-targeting lipid nanocarrier for treating osteoporosis more effectively. RSC Adv. 2020;10(35):20445–59.

    Article  CAS  Google Scholar 

  183. Saftig P, Hunziker E, Everts V, Jones S, Boyde A, Wehmeyer O, Suter A, Figura KV. 2002. Functions of cathepsin K in bone resorption. Cellular Peptidases in Immune Functions and Diseases 2 (293–303). Springer, Boston, MA.

    Chapter  Google Scholar 

  184. Zeng Y, Shen Y, Wu S, Cai L, Wang Z, Cai K, Shen J, Yie KHR, Zhang H, Xu L, Liu J. Bone-targeting PLGA derived lipid drug delivery system ameliorates bone loss in osteoporotic ovariectomized rats. Volume 221. Materials & Design; 2022. p. 110967.

  185. Lapidot T, Kollet O. The essential roles of the chemokine SDF-1 and its receptor CXCR4 in human stem cell homing and repopulation of transplanted immune-deficient NOD/SCID and NOD/SCID/B2mnull mice. Leukemia. 2002;16(10):1992–2003.

    Article  CAS  Google Scholar 

  186. Chen Q, Zheng C, Li Y, Bian S, Pan H, Zhao X, Lu WW. Bone targeted delivery of SDF-1 via alendronate functionalized nanoparticles in guiding stem cell migration. ACS Appl Mater Interfaces. 2018;10(28):23700–10.

    Article  CAS  Google Scholar 

  187. Wynn RF, Hart CA, Corradi-Perini C, O’Neill L, Evans CA, Wraith JE, Fairbairn LJ, Bellantuono I. A small proportion of mesenchymal stem cells strongly expresses functionally active CXCR4 receptor capable of promoting migration to bone marrow. Blood. 2004;104(9):2643–5.

    Article  CAS  Google Scholar 

  188. Zhang C, Zhang W, Zhu D, Li Z, Wang Z, Li J, Mei X, Xu W, Cheng K, Zhong B. 2022. Nanoparticles functionalized with stem cell secretome and CXCR4-overexpressing endothelial membrane for targeted osteoporosis therapy. Journal of Nanobiotechnology, 20(1), p.35.

Download references

Funding

This work was supported by the National Research Foundation of Korea (NRF) grant, the Korean Fund for Regenerative Medicine (KFRM) grant, and the Korea Health Industry Development Institute (KHIDI) funded by the Korean government (MSIT and MOHW) (NRF-2022R1A2C3004850, NRF-2019M3A9H1032376, RS-2023-00208529, 21C0703L1, and HI23C0689000023).

Author information

Authors and Affiliations

Authors

Contributions

The review topic was conceived and designed by P.D.A., W.J.C., H.K., Y.A., and S.L. The manuscript was drafted by P.D.A., W.J.C., H.K., Y.A., A.B.; and it was revised and critically edited by B.J.K, Y.A., and S.L.

Corresponding author

Correspondence to Soo-Hong Lee.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

There are no conflicts to declare.

Additional information

Publisher’s Note

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

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary Material 1

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

Dayanandan, A.P., Cho, W.J., Kang, H. et al. Emerging nano-scale delivery systems for the treatment of osteoporosis. Biomater Res 27, 68 (2023). https://doi.org/10.1186/s40824-023-00413-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40824-023-00413-7

Keywords