You are viewing the site in preview mode

Skip to main content

Strategies for promoting neurovascularization in bone regeneration

Abstract

Bone tissue relies on the intricate interplay between blood vessels and nerve fibers, both are essential for many physiological and pathological processes of the skeletal system. Blood vessels provide the necessary oxygen and nutrients to nerve and bone tissues, and remove metabolic waste. Concomitantly, nerve fibers precede blood vessels during growth, promote vascularization, and influence bone cells by secreting neurotransmitters to stimulate osteogenesis. Despite the critical roles of both components, current biomaterials generally focus on enhancing intraosseous blood vessel repair, while often neglecting the contribution of nerves. Understanding the distribution and main functions of blood vessels and nerve fibers in bone is crucial for developing effective biomaterials for bone tissue engineering. This review first explores the anatomy of intraosseous blood vessels and nerve fibers, highlighting their vital roles in bone embryonic development, metabolism, and repair. It covers innovative bone regeneration strategies directed at accelerating the intrabony neurovascular system over the past 10 years. The issues covered included material properties (stiffness, surface topography, pore structures, conductivity, and piezoelectricity) and acellular biological factors [neurotrophins, peptides, ribonucleic acids (RNAs), inorganic ions, and exosomes]. Major challenges encountered by neurovascularized materials during their clinical translation have also been highlighted. Furthermore, the review discusses future research directions and potential developments aimed at producing bone repair materials that more accurately mimic the natural healing processes of bone tissue. This review will serve as a valuable reference for researchers and clinicians in developing novel neurovascularized biomaterials and accelerating their translation into clinical practice. By bridging the gap between experimental research and practical application, these advancements have the potential to transform the treatment of bone defects and significantly improve the quality of life for patients with bone-related conditions.

Background

Bone is essential to all mammalian species by functioning as a living organ that enables movement. Rehabilitation of bone defects that result from limb trauma, degenerative pathology, or tumor resection poses a challenge for physicians once they exceed a critical size [1,2,3,4]. In the world, as many as 2 million bone grafting procedures are performed annually. This number is expected to increase to over 3 million by 2030, as a result of the aging population and the rising prevalence of bone-related conditions [5].

Autologous bone grafts remain the gold standard for the treatment of bone defects because of their high success rate and compatibility with a patient’s own tissue, minimizing the risk of immune rejection. However, significant concerns persist, including the limited availability of donor bone, the need for a secondary surgical procedure to harvest the graft, and potential complications at the donor site, such as chronic pain, pathogenic infection, and structural weakness [6,7,8]. These issues drive the ongoing search for effective alternative treatments and the development of biomaterials that can mimic the biological and functional properties of natural bone tissue, while reducing the associated risks and limitations [9]. Nevertheless, most synthetic biomaterials only replicate the mechanical properties and macroscopic structures of bone. These biomaterials lack the ability to accurately mimic the complex bone microenvironment and functional units. They often fall short in promoting the same level of osteogenesis, vascularization, and integration with existing bone tissue. As a result, the healing process is slower and less effective. These limitations highlight the need for continued research and development of advanced biomaterials that can more closely emulate the biological and functional characteristics of natural bone.

Osteogenesis is not solely the work of bone cells, but involves the collaboration of multiple systems, including the vascular and nervous systems [10,11,12]. The general distribution of blood vessels and nerves in the body is shown in Fig. 1. The figure illustrates the widespread distribution of blood vessels and nerves in the teeth, jaws, and femurs. Bone cells, including osteoblasts, osteoclasts, and osteocytes, are central to the formation, resorption, remodeling, and maintenance of bone tissue. However, these processes are heavily influenced by the vascular system, which supplies the basic necessities, such as oxygen, nutrients, and growth factors, and provides a pathway to remove the waste products; Blood vessels also play a key role in the recruitment of osteoprogenitor cells to sites of bone formation and repair [13,14,15,16]. However, when designing bone repair materials, nerves have often been overlooked [17]. Nerves secrete neurotransmitters, neurotrophins, and neuropeptides that influence bone cell activity and differentiation [18,19,20]. Moreover, nerves guide angiogenesis and control the blood flow in intrabony blood vessels [21,22,23]. Consequently, the crosstalk between bone cells, endothelial cells (ECs), nerve cells, and immune cells creates a specific microenvironment to restore homeostasis and regulate tissue repair [24,25,26,27]. Many kinds of tissue-specific biomaterials with bioactive components that promote neurovascular regeneration have been developed in order to enhance the therapeutic efficacy. Recent studies have systematically introduced new advances in cellular crosstalk between bone and nervous system [28,29,30,31,32]. Nevertheless, few have comprehensively summarized the regulatory effect of the neurovascular system  in the bone microenvironment.

Fig. 1
figure 1

Distribution of blood vessels and nerves in different parts of the body, including the teeth, jaws, and femurs. a The inferior alveolar artery and vein within the mandible. b The inferior alveolar artery and vein branching out to the pulp chamber of the mandibular molar. c The inferior alveolar nerve within the mandible. d The inferior alveolar nerve branching out to the pulp chamber of the mandibular molar. e Distribution of blood vessels within the femur, which mainly consists of periosteal artery, nutrient artery and emissary vein. f Distribution of nerves within the femurs, including sensory nerves and sympathetic nerves

In this review, the anatomic structures and functions of neurovascular coupling in bone development, metabolism and repair are first introduced. This is followed by a discussion on the design strategy and mechanism of neurovascularized materials to promote osteogenesis. Then, future direction perspectives in this field are proposed to provide a platform for clinical translation. These therapies aim to improve patient outcomes by enhancing the integration and functionality of bone grafts, reducing healing times, and minimizing complications. By bridging the gap between experimental research and practical applications, these advancements have the potential to fundamentally revolutionize the treatment of bone defects, and to improve the quality of life for patients suffering from bone-related conditions.

Neurovascular coupling in bone development, metabolism and repair

The neurovascular system is extensively distributed throughout the skeleton, such as the periosteum, cortical bone, and bone marrow (Fig. 2) [11, 33,34,35]. However, the extent to which neurovascular coupling contributes to improvements in skeletal remodeling remains obscure. In this section, the role of neurovascular coupling in bone development, metabolism, and repair will be comprehensively discussed. By understanding these fundamental aspects, a clearer picture of how nerve-vessel interactions influence bone health may be formed, potentially guiding future research to address the existing knowledge gaps.

Fig. 2
figure 2

Distribution of the neurovascular system in bone. a The blood supply of a long bone. The marrow cavity contains a large central venous sinus, a dense network of medullary sinusoids, and longitudinal medullary arteries and their circumferential rami [33]. b A simplified schematic of the neuronal distribution in the mouse femur [34]. c A schematic illustrating the general pattern and course of the sensory nerve fibers and blood vessels in the periosteum and mineralized bone [11]. d A schematic of the morphology and distribution of type H and type L blood vessels. Arrowhead marks the entry of the arteriole through the cortical bone [35]. CGRP calcitonin gene-related peptide

Role of neurovascular coupling in bone development

Embryological anatomical studies show that neural tissue develops before bone tissue. Innervation first appears in the central part of the backbone, followed by the extension to the metaphysis. Subsequently, bone tissue forms canals surrounding neural pathways, thereby forming new bone [36]. Two bone formation processes including intramembranous ossification and endochondral ossification are present during embryogenesis [37].

Intramembranous ossification occurs during the development of flat bones including the skull, mandible, maxilla, and clavicle. Ossification begins with the condensation of mesenchymal stem cells (MSCs). These cells secrete growth factors such as vascular endothelial growth factor (VEGF)-A to promote the differentiation of MSCs into osteoprogenitors and osteoblasts. Ossification centers are eventually formed as blood vessels invade these centers to promote osteogenesis [38]. During endochondral ossification in long bones, chondrocytes become hypertrophic with a high expression level of VEGF. Blood vessels infiltrate the hypertrophic cartilage to induce the differentiation of osteoblasts and the formation of the bone marrow cavity [37]. The first capillary plexus continues to sprout longitudinally towards the ends of the bone. This causes the bone marrow cavity to expand from the center and form epiphyseal growth plates at both ends. Then ossified bone and bone marrow gradually replace the cartilage [39, 40]. In addition, the vascular system serves as an active secretor of many signaling molecules to regulate the growth, differentiation, patterning, homeostasis, and morphogenesis of developing tissues [41, 42]. Bone morphogenetic protein 2 (BMP-2) and BMP-4, released by vessel-forming cells, act on chondrocytes and osteoblasts [43]. In addition, factors including platelet-derived growth factor type BB (PDGF-BB), slit guidance ligand 3 (SLIT3), hypoxia-inducible factor 1-alpha (HIF-1α), and Notch signaling have been reported to be involved in osteogenesis-angiogenesis coupling [14].

Sensory and sympathetic nerve fibers are widely involved in both primary and secondary ossification processes. During postnatal development, interleukin (IL)-6 induced skeletal sympathetic cholinergic nerve fibers preserve the survival and function of osteocytes through the neurotrophic axis of the neurturin-GDNF-family receptor-α2 (NRTN-GFRα2) [44]. Meanwhile, perichondrial cells of the embryonic femur drive sensory nerve innervation of this bone element by secreting nerve growth factor (NGF). This cytokine promotes bone vascular invasion, progression of bone progenitor cell lineage, and eventually the formation of the primary ossification centers. These phenomena support a predominant role of skeletal innervation in bone development, which can affect the acquisition of peak bone mass.

Vascularization and innervation have many similarities in embryonic development and anatomy. Similar structures such as the growth cone in axons and tip cells in vessels orchestrated the chemotaxis of axons and blood vessels [45,46,47]. ECs secrete neurotropic factors such as artemin and neurotrophin 3 (NT3) to recruit and control innervation [48]. Sensory neurons and Schwann cells (SCs), in turn, secrete VEGF to regulate vessel growth in a spatiotemporally controlled manner [49].

Role of neurovascular coupling in bone metabolism and health

The nervous system plays a regulatory role in bone metabolism and health [50]. Receptors for specific neuropeptides are present extensively in skeletal cells. The central nervous system (CNS) influences bone mass through humoral mechanisms. These mechanisms include the regulation of plasma calcium, and signals from the thyrotropic, hypothalamic-pituitary-corticotropic, somatotropic, and gonadotropic axes [18, 51]. Moreover, peripheral nerves communicate with the skeleton to regulate bone metabolism through nerve-resident cells, neuropeptides, axon guidance factors, and neurotrophins [31, 36, 52, 53]. These biological factors act directly on different target cells or organs to control bone metabolism and formation (Table 1) [54,55,56,57,58,59,60,61,62,63,64]. The sympathetic nervous system is particularly important in osteogenesis, as well as in bone metabolism and remodeling. Sympathetic nerves are regulated by leptin from the hypothalamus, which in turn regulates the outgoing sympathetic nerves to control bone mass [65]. Conversely, pathological conditions of the bone can affect nerve innervation [66,67,68]. For example, bone tumors can release neuropeptides that activate and sensitize the bone nociceptors and different areas of the nervous system, resulting in pain [69, 70].

Table 1 Effects of neuro-associated molecules on bone regeneration

The vasculature plays a crucial role in the cross-talk between various cells within the bone tissue microenvironment. Type H vessels, found in the metaphysis and endosteum, are major regulators of bone metabolism. Any functional disruption in the Notch signaling pathway, which is a major regulatory pathway for type H vessels, reduces the abundance of ECs and type H vessels in the bone microenvironment [71, 72]. Factors such as PDGF-BB, SLIT2, SLIT3, HIF-1α, and VEGF have been shown to influence bone metabolism by affecting the formation of type H vessels [73,74,75]. HIF-1α acts as a cellular oxygen sensor and plays a key role in regulating bone homeostasis and angiogenesis [76]. HIF-1α has been reported to promote angiogenesis and osteogenic differentiation via the VEGF/Akt/mammalian target of rapamycin (mTOR) signaling pathway in adipose-derived stem cells [77]. Glucocorticoid-induced osteoporosis is treated by regulating the adenosine monophosphate-activated protein kinase/mTOR and HIF-1α/VEGF signaling pathways [78]. Matrix metalloproteinase (MMP)-2 inhibitor 1 induces osteogenesis differentiation of bone marrow mesenchymal stem cells (BMSCs) and promotes type H vessel angiogenesis to rescue osteoporosis through the HIF-1α signaling pathway [79]. MiR-26a-5p in extracellular vesicles derived from urine stem cells activates the HIF-1α/VEGF pathway by inhibiting histone deacetylase 4, promoting the differentiation of osteoblast progenitor cells and inhibiting osteoclast activity, and preventing diabetic osteoporosis [80]. SLIT2 and SLIT3, axon guidance molecules secreted by osteoclast lineages, function as local coupling factors that preserve bone balance and protect bone metabolism [61]. SLIT2 inhibits osteoclast differentiation and reduces the migration and fusion of preosteoclasts by suppressing recombinant cell division cycle protein 42 activity [81]. SLIT3 promotes bone formation and inhibits bone resorption through Robo receptors; it has strong therapeutic potential in metabolic bone diseases [82].

Evidence also suggests that neurovascular coupling has a significant impact  in bone metabolism. Calcitonin gene-related peptide (CGRP) is believed to promote bone formation partly due to its ability to dilate blood vessels and stimulate EC migration, promoting angiogenesis in bone remodeling [83]. In addition, substance P and neuropeptide Y secreted by sensory neurons can effectively activate ECs and promote angiogenesis. Sympathetic activation, however, may cause a decrease in type H vessels. MSCs can indirectly regulate EC angiogenesis through paracrine effects under conditions of sympathetic excitation [84]. These findings provide important evidence of the role of neurovascular coupling in bone metabolism. They offer a strong reference for understanding the treatment of bone-related diseases. Osteoporosis is a common metabolic bone disease in middle-aged and elderly populations. This condition is linked to Alzheimer’s disease, an age-related neurodegenerative disorder known for impairing memory and cognition. Alzheimer’s disease is associated with osteoporosis through an abnormal central serotonergic regulatory pathway [85, 86]. This pathway upregulates sympathetic nervous signaling, which in turn activates β-adrenergic receptor on bone cells, enhancing bone resorption [87,88,89]. Further elucidating these underlying mechanisms may provide new avenues for the prevention and treatment for bone diseases.

Role of neurovascular coupling in bone repair

Bone repair is a complex, multi-step process that involves inflammation, neurovascular network reconstruction, rapid bone mineralization, and remodeling [90,91,92,93,94]. A bone defect, such as a fracture, creates a microenvironment of injury. The disruption of oxygen due to insufficient blood supply produces a hypoxic environment, which is an important physiological signal in the process of bone repair. This hypoxic environment regulates the production of key biological factors by osteoblasts; these factors affect EC proliferation, determine cellular differentiation, and induce ECs to secrete osteogenic growth factors [95, 96]. The accompanying inflammation and revascularization are the most crucial phases of bone repair [97]. Platelets first appear at the site of injury to form a blood clot or hematoma. High levels of VEGF-A within the hematoma promote vascular invasion. The hematoma also serves as a template for temporary vascular bone scab formation. This is followed by cartilage scab formation to stabilize the fracture. The cartilage scab contains osteoblasts that promote bone formation, chondrocytes that contribute to new cartilage formation, and fibroblasts. The cartilage scab then matures into a hard bone scab, which is finally remodeled into mature bone, forming new trabeculae and cortical bone. In a murine femoral fracture model, inhibition of VEGF signaling by Fms-related receptor tyrosine kinase 1 (Flt-1), delayed cartilage turnover, disrupted conversion of soft cartilaginous callus to a hard bony callus, and impaired fracture healing [98]. This unstable, hypoxic environment results in indirect bone healing that resembles endochondral osteogenesis [38, 99, 100].

Nerve fibers play a critical regulatory role not only in bone metabolism but also in fracture repair. It is reported that the healing process of fracture in patients with brain trauma was significantly accelerated [101]. Accelerating fracture healing after CNS injury may be related to factors such as the local release of growth factors in the brain [102,103,104,105]. The sensory nerves in peripheral nerves dominate bone repair by transmitting information about local conditions of bone fracture to the CNS and allowing the perception of pain signals. This can elicit an appropriate neuroendocrine response to modulate bone turnover locally at the fracture site [106]. In addition, sensory nerves are directly involved in osteogenesis through the secretion of neuropeptides, which stimulate the corresponding receptors on bone cells [107]. In the early bone repair phase, NGF was highly expressed in both macrophages and periosteal MSCs at the fracture site [108]. NGF not only stimulates the ingrowth of CGRP+ sensory fibers and tyrosine hydroxylase-positive (TH+) sympathetic fibers to the fracture callus from the periosteum and bone marrow, but also recruits MSCs to migrate toward bone defects and enhance bone formation via the p75 signaling pathway [109, 110]. Many other neuro-associated molecules are involved in different stages of fracture healing (Fig. 3) [17]. This substantiates that neurogenesis precedes vascular growth during bone repair. The interplay between nerves and blood vessels is intricate, with both entities collaborating to facilitate bone repair. Mg nail implantation has been shown to enhance the repair of critical-sized bone defects during distraction osteogenesis via CGRP/focal adhesion kinase (FAK)/VEGF signaling axis. This pathway may serve as a major signaling mechanism linking sensory nerve and ECs [111]. Research has demonstrated that CGRP has an equal effect on migration and tube formation compared to VEGF. Hence CGRP is a potent pro-angiogenic growth factor during bone healing [112].

Fig. 3
figure 3

Expression of neuro-associated molecules and related events during bone healing. a The line chart displays the temporal sequence of events following a bone fracture. b Distinctive distribution of neuropeptides across the 4 phases of bone healing [17]. BDNF brain-derived neurotrophic factor, CGRP calcitonin gene-related peptide, NGF nerve growth nerve, NPY neuropeptide Y, Sema3A semaphorin 3A, SP substance P

In summary, neurovascular coupling plays a vital role in bone embryonic development, metabolism, and repair [72, 113]. Recently, there has been a growing body of research on how nerves and blood vessels affect bone regeneration [110, 114, 115]. However, studies that specifically address how neurovascular interactions influence bone-related conditions are still lacking. In addition, most research has focused on the effects of neuropeptides and cytokines. However, the clinical evidence of the application of neuropeptides and their receptor antagonists or agonists in conditions still has significant limitations [116]. Recently, cardiovascular research has introduced injectable and long-lasting CGRP analogues, which show antihypertensive effects, attenuate cardiac failure, and improve metabolic parameters in mice [117, 118]. Therefore, it is clinically important to investigate whether these novel CGRP analogues can not only be effective in treating metabolic and cardiac diseases, but also enhance innervated and vascularized bone regeneration, and whether their potential benefits outweigh the nociceptive effects known to be facilitated by CGRP. Investigating whether and how bone-organ axis regulates the intraosseous neurovascular system represents a promising area for future research. In the brain-bone axis, while the brain exerts dominant effects in bone metabolism, homeostasis, and disease progression, bones, in turn, signal to the brain to promote brain development and skeletal growth [12]. For example, psychological stress has been shown to induce changes in bone mass through activation of the hypothalamic–pituitary–adrenal axis, glucocorticoid signaling, and a blunted response to growth factors [119]. Estrogen and osteocalcin (OCN) are hormones involved in bone homeostasis. These hormones may also influence cognitive function by inhibiting neuronal apoptosis and activating the acute stress response through the inhibition of the peripheral nervous system [120,121,122].

Materials for regenerating innervated and vascularized bone

The interdependence between biomaterials and biology is recently recognized [123]. In this relation, altering the composition, physical properties, and structural characteristics of materials can affect the basic physiological functions of cells, including cell migration, proliferation, and differentiation. In response, cells secrete extracellular matrix (ECM) components to remodel tissues, maintaining their mechanical properties and architecture [124]. This further illustrates the impact of material properties on biological functions at different levels (i.e., cells, tissues, organs, and the whole organism). Thus, understanding how materials influence functions of multiple cells is essential for designing biomaterials that can regenerate bone with proper nerve and blood supply [125]. Representative examples of biomaterials used for innervated and vascularized bone regeneration are summarized in Table 2 [110, 114, 126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144].

Table 2 Representative examples of biomaterials used for innervated and vascularized bone regeneration

Material properties

Natural bone tissue is heterogeneous due to the varying density and distribution of bone tissues, blood vessels, nerves, and peripheral tissues, which makes it challenging to regenerate different tissues using the morphological characteristics of bioactive materials [145]. The physicochemical properties of materials, such as stiffness, surface topography, pore structures, conductivity, and piezoelectricity are crucial for tissue regeneration because these properties significantly influence cellular behaviors [146,147,148,149].

Stiffness

Stiffness is defined as the ration of stress to strain. It is considered a crucial biomechanical factor for bone tissue engineering [150]. Bone tissue comprises cortical and cancellous bone, with elastic moduli ranging from 10–25 GPa and 0.1–2.0 GPa, respectively [151]. An ideal scaffold material should match the stiffness of human bone to fulfill load-bearing function. Material stiffness could regulate cell behaviors such as cell adhesion, proliferation, migration, and differentiation, by making cells perceive the mechanical properties of ECM [150]. This process, known as mechanotransduction, involves cells converting extracellular physical sensations into intracellular biochemical signals [152]. Scientists found that osteogenic differentiation of MSCs was enhanced which were cultured on a stiff ECM (40 kPa) compared to a soft ECM (4.5 kPa) [153]. In another study, it was proved that the osteogenic differentiation of the MSCs occurred predominantly at 11–30 kPa, especially at 22 kPa in 3D hydrogel [154]. Mechanistically, the stiffer ECM regulates glutamine metabolism to contribute to osteogenesis, with Yes-associated protein playing an indispensable role in this process [152, 155]. Recent studies reported a significant role of the matrix stiffness in angiogenesis and neurotization. Exposure of human umbilical vein endothelial cells (HUVECs) to substrates of varying stiffness modulated the expression of major angiogenesis mediators and growth factors involved in bone repair and regeneration [156]. Collagen and hydroxyapatite (HA) mixtures in varying proportions were coated on decellularized cancellous bone to investigate the effect of stiffness on angiogenesis and bone regeneration. Compared to matrices with a stiffness of (13.00 ± 5.55) kPa, (13.87 ± 1.51) kPa and (37.70 ± 19.60) kPa showed higher expression of osteopontin, OCN, and increased aggregation of blood vessel-like ECs [157]. In addition, the cell spreading area and neurite length of differentiated neuron-like PC-12 cells significantly improved on 34.9 kPa gelatin methacryloyl (GelMA) substrates. This finding indicates that stiffer materials enhance cell adhesion and proliferation [158, 159]. It has been shown that osteogenic differentiation, vascular differentiation, and neural differentiation of stem cells differ when they are cultured on substrates with different stiffness [154, 160, 161]. When cultured on substrates with low, intermediate, or high stiffness, human MSCs differentiate into neurons, myoblasts, or osteoblasts, respectively. The stiffest substrate (25–40 kPa) demonstrated predominantly osteogenic differentiation [162]. Thus, determining the optimal stiffness of biomaterials is necessary for enhancing the growth and activity of different cell types to achieve neurovascularization in bone regeneration.

Surface topography

Careful assessment of surface topography is essential to enhance the regenerative capacity of bone. Similar to the stiffness, cells sense the material surface via adhesion receptors such as integrins. These molecules bind with intracellular functional proteins, triggering signal transduction and affecting further cellular responses to the implanted materials [150]. Surface topography plays a crucial role in regulating various cell behaviors including stem cell differentiation, myoblast migration, osteoblast maturation, and angiogenesis [163,164,165,166]. Both natural bone and blood vessel walls have some nanostructured surfaces composed of ECM proteins [124]. Surface modification is a well-known topographic feature for improved bioactivity, promotion of osteogenesis, enhancement of angiogenesis, moderation of pro-inflammatory responses, augmentation of anti-inflammatory responses, and reduction of osteoclast resorption activity [24, 167].

Methods for roughening smooth implants include sandblasting and acid etching, electrophoretic deposition, anodization, and hydrothermal treatment [168,169,170,171,172,173]. Micro-/nano-structured implants outperform smooth ones in both healthy and compromised animals [174,175,176]. A recent study investigated the effect of titanium nanotubes with different diameters (30 nm, 70 nm, or 110 nm) on osseointegration [177]. Titanium nanotubes with a 70 nm diameter significantly reduced early inflammation, promoted osteogenesis-angiogenesis coupling, and enhanced peri-implant osseointegration. These nanotubes also favored macrophage polarization towards an anti-inflammatory M2 phenotype through the FAK-phosphatidylinositide 3-kinase (PI3K)-mediated integrin α5/β3 pathway. Likewise, nano-structured surfaces of polymer substrates like polyurethane or poly(lactic-co-glycolic-acid) enhance the adhesion and growth of vascular-related cells such as ECs and vascular smooth muscle cells [178, 179]. Optimizing the topography and orientation of nerve tissue in biomaterials improves nerve tissue engineering outcomes by providing a suitable microenvironment for nerve cell growth and alignment [180].

In addition, a series of polyhedral bioceramic scaffolds fabricated through 3D printing and featured various spatial topologies [truncated tetrahedron, truncated octahedron, truncated cube, and cuboctahedron] were developed to enhance both nerves and blood vessels integration during bone regeneration. Notably, the truncated tetrahedron and truncated octahedron scaffolds exhibited the highest surface-to-volume ratio, providing increased surface area that supports greater cell adhesion and proliferation. This, in turn, facilitated both nerve regeneration and angiogenesis. Furthermore, these scaffolds demonstrated significantly higher compressive strength and featured tunable mechanical properties by design, ultimately promoting bone regeneration through activation of the PI3K/Akt signaling pathway [181]. These examples illustrate the importance of surface topography in designing biomaterials for effective tissue regeneration.

Pore structures

Pore structures with appropriate pore size and interconnected porosity are crucial for mimicking natural bone tissue [182]. Generally, large pore sizes favor blood vessel growth, while small pore sizes lead to poor contact between newly-formed tubes [183]. Connections between pores create anastomoses between capillaries, forming a functional vascular network. Polylactic acid scaffolds with a pore size less than 125 μm affect vessels penetration, while those with a diameter exceeding 250 μm promote vessel formation [184].

For Ti6Al4V scaffolds, the optimal pore size for angiogenesis is 550 μm [185]. A study using a 3D porous, biodegradable calcium phosphate scaffold found the optimal pore size for angiogenesis and bone formation is 540 μm. Scaffold porosity significantly influences arterial growth [186]. Scaffolds with a porosity of over 80% and a pore size of 400–600 µm are beneficial for bone regeneration [187,188,189]. Higher scaffold porosity increases the number of ECs, active osteoblasts, and bone mass. These findings indicate that angiogenesis and osteogenesis seem to depend on more porosity than pore size [190]. Although sufficient porosity is essential, inadequate connection and communication between pores may still result in poor vascularization [191]. The interconnection between pores plays a critical role in determining blood permeability and cell migration. These factors are vital for blood vessel formation within the scaffold [192]. 3D porous beta-tricalcium phosphate (β-TCP) scaffolds with identical pore sizes but varying interconnection sizes (100, 120, and 150 µm) were developed to investigate the effects of pore interconnection on angiogenesis. An in vivo study using a rabbit femoral condyle defect model demonstrated that a 150 µm interconnection size significantly improved vascularization by activating the PI3K/Akt signaling pathway [193].

The optimal microstructural parameters of conduit designs seem 50–70% porosity, 5–30 μm pore size, and approximately 600 μm wall thickness for nerve regeneration [194]. Interconnected macropores are crucial for nutrient transport, cellular infiltration, and bone ingrowth, while micropores around 10 μm promote protein adsorption, cellular adhesion, and ion exchange. More research is needed to determine the optimal pore sizes for bone repair materials to support neurovascularization effectively [195, 196].

Conductivity

Conductive biomaterials have been studied extensively since their introduction in 1994 [197]. They possess high electrical conductivity and electrochemical properties, enhancing cellular signaling in damaged tissues [198,199,200,201]. Research has shown that electrical stimulation aids in the repair and regeneration of damaged tissues including bone, blood vessel, nerve, and ligaments [202]. Unlike external electrical stimulation, conductive biomaterials do not require external devices. This offers a new solution for promoting neurovascularization in bone regeneration, with significant clinical potential [203].

Black phosphorus is an isomer and the most stable form of phosphorus at room temperature and pressure. Black phosphorus nanosheets have superior conductivity, making them ideal for electrical stimulation in bone and nerve regeneration [134]. A core–shell electroactive membrane, developed using coaxial electrospinning technology, was designed to mimic natural bone membranes. These nanosheets were introduced onto the biomimetic periosteum via electrostatic interaction to enhance conductivity. This electrically-active periosteum enhances axon growth, promotes neurotransmitter secretion, and induces neurogenic osteogenesis (Fig. 4a) [204]. However, black phosphorus degrades in oxygen and water, which limits its biomedical applications [205]. Modification strategies such as polymer coating, surface chemical modification, and cell membrane embedding have been used to address this issue [206, 207]. In addition, copper ion-modified germanium phosphide nanosheets have been used to prepare a conductive GelMA hydrogel. This hydrogel demonstrated excellent innervation and vascularization during bone regeneration, with the added benefit of bacterial clearance during wound healing (Fig. 4b) [127].

Fig. 4
figure 4

Conductive-based biomaterials promote innervated and vascularized bone regeneration. a Schematic of the (core)-polycaprolacton/(shell)–DNM biomimetic periosteum (PD)@black phosphorus promoting neurogenic bone regeneration [203]. b Schematic of the process used for the fabrication of the gelatin methacryloyl (GelMA)/GeP@Cu electroactive hydrogel and its multiple therapeutic actions supporting bone regeneration [127]. c Fabrication of Mo2Ti2C3 MXene hydrogel and its application in bone defects. d Immunofluorescence results showed MXene hydrogel increased the relative intensity of 5-hydroxytryptamine (5-HT) and bone morphogenetic protein 2 (BMP-2) expression at 8 weeks [130]. Scar bar = 100 μm. bFGF basic fibroblast growth factor, BMSCs bone marrow mesenchymal stem cells, CGRP calcitonin gene-related peptide, HF hydrofluoric acid, ROS reactive oxygen species, TPAOH tetrapropyl ammonium hydroxide, VEGF vascular endothelial growth factor

A biocompatible and electrically conductive Mo2Ti2C3 MXene hydrogel was prepared using Mo–Ti to replace the Ti–Ti structure (Fig. 4c). The hydrogel upregulated the expression of the sensory neuron marker 5-hydroxytryptamine and the osteogenic factor BMP-2 (Fig. 4d) [130]. These results confirmed that the hydrogel promotes neurogenesis and bone regeneration in vivo. Despite these achievements, there is a lack of systematic analysis and evaluation of the long-term biosafety of MXene nanomaterials and their degradation products. Future research should focus on understanding the immunogenicity, pharmacokinetics, biological distribution, and long-term toxicity of these materials in animal models [208]. The development of conductive biomaterials for tissue engineering is still in early stages, facing significant scientific and technological challenges. Future efforts should improve features such as biodegradability, biocompatibility, stability, and compatibility to bridge the gap between laboratory findings and clinical applications.

Piezoelectricity

Unlike conductive materials, piezoelectric materials can immediately generate electrical signals in response to mechanical stress. This helps eliminate the risk of tissue damage from external electrical stimulation [209,210,211]. Because natural bone itself exhibits piezoelectric properties, different piezoelectric polymers such as polyvinylidene fluoride, polyhydroxyalkanoates, and barium titanate have been used for bone repair. These materials are selected because of their flexibility, ease of synthesis, and ability to offer an electrical stimulus [212,213,214,215]. However, their limited degradability and lack of bioactive components are insufficient for promoting neurovascularized bone regeneration [216, 217]. Whitlockite (WH), an inorganic nanoparticle with a composition similar to HA, releases Ca2+, Mg2+, and PO43− ions during degradation, and the released Mg2+ ions are known to support nerve and blood vessel formation [218]. Moreover, WH exhibits good piezoelectric properties upon sintering [219]. For example, a 3D-printed polycaprolactone/piezoelectric WH (PWH) composite scaffold was developed to mimic natural bone functionality. This scaffold enhances angiogenesis, promotes neuronal differentiation, suppresses osteoclast activity, and ultimately improves osteogenesis. The synergistic effect of bioactive ion release and piezoelectricity in this biodegradable PWH scaffold supports innervated and vascularized bone regeneration [140].

Furthermore, PWH has been utilized to reestablish the piezoelectric properties of natural periosteum, an important factor in bone healing. A double-network hydrogel composed of chelated alginate, GelMA, and PWH was designed to emulate the viscoelasticity and piezoelectricity of natural periosteum. Combined with a bone-like substrate, this periosteum-inspired structure reproduces the heterogeneous architecture of native bone tissue. When subjected to low-intensity pulsed ultrasound stimulation, this bioinspired scaffold significantly enhances early vascularization and neurogenesis. Under dynamic physiological conditions, the double-layer scaffold can function as a self-powered electrical stimulator, accelerating bone regeneration. This provides a valuable reference for exploring the physical properties of materials optimized for neurovascularized bone regeneration [220].

Acellular biological factors

Biomaterials alone are ineffective in promoting osteogenesis due to the lack of bioactive components. They fail to mimic the body’s dynamic responsiveness to bone defects. Seed cells are the foundation of tissue engineering and regenerative medicine. An ideal seed cell should have strong proliferation ability, strong environmental adaptability, and good tissue compatibility [17]. Various seed cells have been employed, including BMSCs, SCs, and ECs. In addition to the cell delivery strategy (Table 3) [132, 221,222,223,224,225,226,227,228], some acellular biological factors have been loaded into biomaterials to provide additional stimuli. The following subsections will highlight representative examples of innervated and/or vascularized bone regeneration.

Table 3 Cell delivery strategy for innervated and vascularized bone regeneration

Neurotrophins

NGF was the first discovered neurotrophin [229]. It is widely expressed in osteoblasts, osteoclasts, osteocytes, and osteochondrocytes. The neurotrophin family includes NGF, brain-derived neurotrophic factor, glial cell line-derived neurotrophic factor, neurotrophin (NT)3, and NT4/5 [18]. Neurotrophins and their receptors regulate osteoblastogenesis, osteoclastogenesis, chondrogenesis, and angiogenesis during bone formation and injury repair [230,231,232]. There are two receptors, tyrosine kinase receptor A (TrkA; high-affinity) and p75 neurotrophic factor receptor (p75NTR; low-affinity) [233,234,235]. Neurotrophins support many kinds of neural activities, including axonal growth, synaptic plasticity, cell differentiation, and myelination by activating distinct TrkA [236]. Additionally, the p75NTR promotes osteogenesis by stimulating the BMP/ small mothers against decapentaplegic homolog 1 (SMAD1) signaling pathway, and inhibits bone resorption by down-regulating receptor activator of nuclear factor-κB ligand (RANKL) expression [230]. By binding to these receptors, NGF initiates signaling cascades like MAPK and PI3K/Akt, which sensitize neurons and stimulate axon and dendrite growth [163, 230, 237, 238].

Moreover, NGF exhibits angiogenic properties [239,240,241]. Meanwhile, β-NGF was locally applied with collagen bone fillers for critical-sized bone defect repair in rats. The β-NGF promotes nerve growth and stimulates VEGF synthesis via TrkA and ERK2 pathways. Although the use of NGF in neurovascularized bone regeneration is promising, its vulnerability and short half-life require suitable delivery vehicles for controlled, sustained delivery [164, 165].

Laminins (LMs) are important structural proteins in ECM, containing domains with affinity for growth factors. Two LM isoforms, LM332 and LM411, bind to BMP-2 and β-NGF, respectively. LM/polyethyleneglycol (PEG)-based hydrogels enhance BMP-2 and β-NGF bioactivity and stability, promoting bone and nerve regeneration (Fig. 5a, b) [166]. When NGF binds to porcine dermis-derived ECM nanofibrous scaffolds, it promotes bone defect healing by repairing damaged sensory nerves. NGF activates the TrkA receptor and stimulates CGRP secretion to promote angiogenesis (Fig. 5c) [110]. Laponite, a synthetic 2D silicate, fixes NGF in biomaterials by electrostatic adsorption [242]. Laponite loaded with NGF and BMSCs in a hybrid hydrogel regulates nervous system function, vascularization, and ossification to form functional bone tissue (Fig. 5d) [222]. Clinical trials suggest targeting NGF provides pain relief and improves physical function in osteoarthritic patients [180, 243]. This may be an effective target for pain treatment.

Fig. 5
figure 5

Strategies of delivering nerve growth factor (NGF) using biomaterials address burst release of NGF for innervated and vascularized bone regeneration. a Immunofluorescence results showed laminin (LM)332/polyethyleneglycol (PEG) effectively delivers bone morphogenetic protein 2 (BMP-2) and promotes the expression of late osteogenic markers osteopontin (OPN) and osteocalcin (OCN). b Immunofluorescence represented images of DRG cells in diverse culture conditions, which showed DRG cultured in LM411/PEG + GCRDVPMSMRGGDRCG peptide (VPM) hydrogels with 1 µg/ml of β-NGF showed the longest neurite outgrowth [165]. Scale bar = 500 µm. c The adsorption capacity of the acellular scaffold was leveraged to construct a sustained release system of NGF, which promoted sensory nerves reinnervation and bone repair [110]. d The schematic diagram showed the preparation of bioprinted constructs, which promote bone regeneration through sensory nerves and blood vessels regeneration [221]. AlgMA alginate methacrylate, BMP-2 bone morphogenetic protein 2, BMSCs bone marrow mesenchymal stem cells, CGRP calcitonin gene-related peptide, DAPI 4',6-Diamidino-2-phenylindole, DRG dorsal root ganglia, FN fibronectin, GelMA gelatin methacryloyl, MSCs mesenchymal stem cells, OM osteogenic media, TrkA tyrosine kinase receptor A, UV ultraviolet

Brain-derived neurotrophic factor binds to TrkB and p75NTR, and is crucial for neuron survival and differentiation [230]. This neurotrophin also promotes angiogenesis and bone formation during human fracture healing via the TrkB [244]. Injecting NT3 promotes BMP-2 and VEGF mRNA expression to enhance osteogenesis and angiogenesis [64]. Future efforts should explore the roles of other neurotrophins in bone and evaluate their clinical biosafety.

Peptides

Accelerating angiogenesis and innervation through peptides is a significant development in bone regeneration. Integrins, major cell-surface adhesion receptors, play key roles in cell spreading and proliferation [245]. Several bioactive peptides bind integrins to improve wound healing. Peptides such as Cys-Ala-Gly, Arg-Gly-Asp-Val (REDV), and Ser-Val-Val-Tyr-Gly-Leu-Arg have high affinity with ECs [246,247,248]. The REDV peptide, known for selective adsorption and proliferation of ECs, is used in surface modification of biomaterials for bone regeneration [249, 250]. A scaffold coated with REDV promotes early intrabony vascularization by binding to α4β1 integrin to attract ECs [251].

Bioactive epitopes in LM include Tyr-Lle-Gly-Ser-Arg (YIGSR) and Ile-Lys-Val-Ala-Val (IKVAV) [252, 253]. The YIGSR peptide guides EC migration, whereas IKVAV enhances EC mobilization, and capillary branching [254, 255]. Grafted into elastin-like polypeptides and branched PEG, these peptides form a composite matrix that, when implanted, lead to osteoid tissues with bone cells, vascular networks, and neuronal structures. However, the limited microporosity of PEG-based hydrogels prevents adequate cell infiltration after in vivo implantation. To address this issue, an all-natural polymer matrix was developed that incorporates ELPM80-alkene, biomimetic peptides containing IKVAV and YIGSR adhesion sequences, the lipophilic Suppocire® NA 15 (SNA15), and HA particles. This formulation provides the components essential for promoting cell colonization, bone mineralization, and the integration of multiple tissues. The formulation was found to enhance innervated and vascularized bone regeneration [256].

The degradation of hydrogels is critical to coordinate tissue infiltration for successful bone healing. Specific degradation peptide sequences can alter hydrogel degradation, promoting target cell migration to injury sites [257]. MMP-cleavable peptides, recognized and degraded by MMPs, serve as triggers in degradable biomaterials. GPQGIWGQ, a commonly-used sequence, is susceptible to MMP-2, -9, and -14 [258]. MMP-2 and -9 degrade ECM, enabling axonal outgrowth and neural cell pathfinding. Conversely, MMP-14 allows ECs to cleave ECM for lumen formation [259,260,261]. A hydrogel formed by polyethylene glycol and GPQGIWGQ was developed as a biomimetic periosteum for repairing bone defects [131]. This biomaterial supported EC migration in vitro, increased neurovascularization and enhanced bone regeneration in vivo. However, peptide degradation rates vary and are susceptible to MMP subtypes [262]. Given the complex and dynamic intercellular microenvironment, more than one MMP may be present at injury sites. Thus, MMP-cleavable peptide-based hydrogels require further investigation to be clinically applicable for bone regeneration.

RNAs

Gene therapy offers an alternative to protein therapy, and the first clinical trials began in the early 1990s [263]. RNAs-based therapeutics, such as mRNAs, microRNAs (miRNAs), small interfering RNA (siRNAs), and long noncoding RNAs, provide new approaches for treating bone diseases [264, 265]. These therapeutics can flexibly express proteins locally and intracellularly. Protein production may be maintained in-situ longer, reducing the need for higher therapeutic protein levels [266]. Delivery of RNAs shows potential to modulate neurobiological and angiogenesis processes, offering new opportunities for bone tissue engineering [267, 268].

miRNA

miRNAs are endogenous small noncoding RNAs (approximately 22 nucleotides) post-transcriptionally regulating gene expression [269]. Many miRNAs modulate neurobiological processes, including axonal outgrowth, synaptogenesis, and neural plasticity [270,271,272]. Among them, miR-222 has the potential to improve innervation in bone tissue engineering [273,274,275]. Co-delivery of miR-222 and aspirin promoted neurogenesis and bone formation in vivo [276]. Several miRNAs, such as miRNA-126, miRNA-210, miRNA-21, and miRNA-675, promote angiogenesis during bone regeneration [277,278,279,280]. They target HIF-1α/VEGF signaling, to improve microcirculation status in the bone injury area to facilitate bone regeneration.

siRNA

siRNAs effectively silence genes post-transcriptionally in eukaryotic cells [281]. RNA interference is crucial in bone regeneration by modulating osteocyte proliferation, differentiation, and function [282]. Inhibiting soluble Flt-1 (sFlt-1) and p75NTR may enhance vascular and neural differentiation, aiding repair of bone defects. However, siRNA therapies face delivery barriers like membrane impermeability, nuclease degradation, and lysosomal degradation [283, 284]. Dual siRNA copolymers, loaded into hybrid cell membranes derived from anti-inflammatory macrophages and osteogenic-induced MSCs, have been employed to address these delivery barriers [138]. This method creates a better bone defect microenvironment by improving angiogenesis, neurogenesis, and inflammatory regulation.

Despite progress, miRNAs and siRNAs applications in bone regeneration are limited to early preclinical trials. RNAs-based therapeutics face challenges in controlling entry routes, gene targeting, and determining optimal dosages.

Inorganic ions

Inorganic ions play an important role in bone repair and regeneration by regulating cellular behaviors and improving the bone microenvironment [285,286,287,288]. For instance, zinc ion has strong antibacterial activity, which can promote the healing of infected bone defects [289]. Metal ions also promote nerve and blood vessel regeneration in bone [134, 290]. Therefore, integrating bone scaffolds with inorganic ions is a promising therapeutic approach for bone defect repair.

Magnesium

Magnesium is essential for bone health, with approximately 60% stored in the bone matrix [291]. Magnesium ions increase the proliferation and function of stem cells, promoting peripheral nerve repair [292]. Implant-derived Mg2+ enters sensory neuron dorsal root ganglia and promotes CGRP-vesicle accumulation and exocytosis, which enhances osteogenic gene expression in periosteum-derived stem cells (Fig. 6a) [293].

Fig. 6
figure 6

Inorganic ions-based biomaterials promote innervated and vascularized bone regeneration. a Schematic showing diffusion of implant-derived Mg2+ promotes osteogenic differentiation toward the periosteum that is innervated by sensory neurons [292]. b Immunofluorescence staining of overexpressing semaphorin 3A (Sema3A) in sensory nerves showed a large number of Leptin receptor (LepR)+ cell, Calcitonin gene-related peptide (CGRP)+ nerve fibers, and CD31+ vessels. Scale bar = 100 μm. c 3D-reconstructed superficial and interior images of femoral condyle defects showed overexpressing Sema3A in sensory nerves could accelerate bone regeneration [304]. Scale bar = 100 μm. d Ce-eggshell membrane (ESM) enhanced gene expressions of vascular endothelial growth factor (VEGF), platelet-derived growth factor-BB (PDGF-BB) and immunofluorescence images demonstrated a significant upregulation of SLIT3 in macrophages after Ce-ESM simulation [136]. *P < 0.05; **P < 0.01; ***P < 0.001, ns non-significant. Scale bar = 50 μm. cAMP cyclic adenosine monophosphate, CALCRL calcitonin receptor-like receptor, CGRP calcitonin gene-related peptide, CREB1 cAMP-responsive element binding protein 1, DRG dorsal root ganglion, MAGT1 magnesium induces magnesium transporter 1, PDSC periosteum-derived stem cell, RAMP1 receptor activity-modifying protein 1, SLIT slit guidance ligand, TRPM7 transient receptor potential cation channel subfamily M member 7, VEGF vascular endothelial growth factor

However, adding Mg2+ to biomaterials presents some challenges. Magnesium ions show a stage-dependent therapeutic effect. In the late stage of osteogenesis, Mg2+ upregulates Matrix Gla protein, a mineral-binding  ECM protein, which, in turn, inhibits HA crystal formation. Moreover, Mg2+ can replace Ca2+ in HA, inhibiting mineralization and osteogenesis. Controlling Mg2+ release is crucial in bone tissue engineering [294,295,296]. A 3D-printed dual-ion chronological release scaffold was designed. The early complete release of Mg2+ can effectively enhance neuro-vascularization without the potential inhibition on late osteogenesis, while long-term release of Zn2+ is responsible for promoting new bone formation [297,298,299,300].

Silicon

Silicon is the major trace element in the human body and is known for its positive effects on osteoblasts, osteoclasts, and ECs [301,302,303]. For instance, silicon-doped HA coatings on titanium implants promoted HUVEC migration and tube formation. These silicon-doped coatings also enhanced the expression of osteogenic markers in MC3T3-E1 cells, compared to HA-coated implants [304]. However, its impact on nerve regeneration is unclear. A silicified collagen scaffold has been shown to induce semaphorin 3A secreted by sensory nerves. In turn, semaphorin 3A stimulates neurovascularization in bone regeneration (Fig. 6b, c) [305].

Inorganic materials such as calcium silicate nanowires, nanoclays, and lithium-magnesium-silicon bioceramics have been used experimentally to assist bone regeneration. These biomaterials release biologically-active ions like Mg2+, Si4+, and Li+ to promote innervated and vascularized bone regeneration [222, 226]. The combined effect of multiple inorganic ions in biomaterials may exceed the role of a single ion.

Cerium

Cerium is recognized for its anti-inflammatory properties and its ability to enhance angiogenesis, neuroprotection, and bone repair [306]. Cerium can switch its oxidation state between cerium III and cerium IV. This property endows cerium oxide with attractive bio-catalytic and immunomodulatory properties for regulating the bone microenvironment [307]. Cerium (III, IV) oxide-mineralized ESMs (Ce-ESMs) was prepared through biomimetic mineralization to simulate natural  periosteum. The Ce-ESMs demonstrate superior mechanical properties and immunomodulatory capabilities. Under Ce-ESM stimulation, macrophages transform into tartrate-resistant acid phosphatase (TRAP)+ pre-osteoclasts. These active pre-osteoclasts secret VEGF, PDGF-BB, and SLIT3 to orchestrate bone regeneration and neurovascularization (Fig. 6d) [136].

Calcium

Calcium ions play an important structural role in bones, blood vessels, and nerves. Approximately 99% of the body’s calcium is found in the bones and teeth. The calcium is stored as carbonated apatite, which is the primary mineral phase of bone. The apatite crystallites provide strength to the skeletal system and serve as a metabolic reservoir for cellular fluids [308, 309]. In addition, calcium stimulates angiogenesis by promoting the proliferation of ECs, and upregulates the expression of VEGF and basic fibroblast growth factor [297]. In neurons, calcium is essential for signal transmission across synapses. Synaptic transmission occurs when an action potential reaches a nerve terminal, causing Ca2+ channels to open. This results in a highly localized and transient increase in intracellular Ca2+ at the active zone. The Ca2+ trigger exocytosis of synaptic vesicle, release of neurotransmitters, and initiate synaptic transmission [310]. Although there is a lack of relevant research on calcium ions in the field of innervated and vascularized bone regeneration, this novel type of ionic material offers novel approaches and insights for innervated and vascularized bone regeneration.

Copper

Copper ions enhance angiogenesis by stimulating proliferation and migration of ECs. These ions also activate pro-angiogenic factors such as VEGF, basic fibroblast growth factor, TNF-α, and IL-1 [311,312,313]. Copper-containing biomaterials, including Cu-doped HA, Cu-doped TCP, Cu-doped bioglass, copper sulfate, and copper sulfide, are used extensively in bone repair [314,315,316,317]. For instance, copper nanoparticles have been incorporated into calcium phosphate cement (CPC) to create Cu-doped CPC. This Cu-doped CPC promotes osteogenic differentiation, proliferation of HUVECs, and in vitro tube formation. Hence, it has the potential to facilitate the repair of cancerous bone defects [318]. Similarly, copper ions show potential for nerve regeneration. As previously mentioned, GelMA/GeP@Cu exhibits excellent electrical conductivity and antibacterial properties. This bioactive material has been reported to upregulate the expression of neuronal class III β-tubulin 1 and microtubule-associated protein 2 in neuroectodermal stem cells. Hence, GelMA/GeP@Cu has the potential to stimulate neurite growth and neural differentiation [127]. However, it is important to note that excessive Cu2+ can have adverse effects, such as cytotoxicity and the induction of apoptosis. This is due to the production of reactive oxygen species via Fenton-like and Haber–Weiss reactions [319,320,321,322]. In a study, Cu2+ and human-exfoliated deciduous teeth-derived exosomes were combined with hyaluronic acid hydrogel to promote periodontal bone regeneration. A concentration of 5.0 μg/ml Cu2+ was found to significantly upregulate the mRNA expression of OCN and runt-related transcription factor 2 (Runx2) in human periodontal ligament stem cells. This effect was more pronounced compared to a concentration of 7.0 μg/ml. These findings suggest that 5.0 μg/ml of Cu2+ exhibits both osteoinductive properties and favorable cytocompatibility [323].

Exosomes

Exosomes are considered as key mediators of cell-to-cell communication. They contain proteins, lipids, and nucleic acids (e.g., mRNAs, small RNAs), which recipient cells absorb to exert functions [324]. Exosomes from different cell origins, such as BMSCs, umbilical cord MSCs, and endothelial progenitor cells, have been widely used in vascularized bone regeneration [277, 325, 326]. In the context of nerve regeneration, exosomes derived from SCs (SC-exos) have shown potential for treating peripheral nerve and spinal cord injuries. They can also enhance angiogenesis during nerve functional recovery [327, 328]. To ensure a sustainable and stable release, SC-exos were icorporated into GelMA. This assembly facilitated innervated and vascularized bone regeneration, with the potential for immune regulation (Fig. 7a) [133]. BMSCs and SC-exos were used as bioinks in GelMA and silk methacrylate hybrid hydrogels to mediate the SC-mediated nerve-bone crosstalk to promote osteogenesis [115]. These bio-printed constructs enhance neurovascularized bone regeneration by stimulating the nervous system.

Fig. 7
figure 7

Natural and engineered exosome-based biomaterials promote innervated and vascularized bone regeneration. a At different stages of bone regeneration, Schwann cell-exosome (SC-exos)/hydrogel improves the osteogenic microenvironment and promotes neurovascularized bone regeneration [133]. b Schematic showing electrospun biomimetic periosteum loaded with aptamers engineered exosomes. These entities can target injured axons and regenerate blood vessels and bone. c Compared with the control group, aptamers engineered exosomes promoted dorsal root ganglion (DRG) axons growth and showed clear guidance [137]. d The N-exos-functionalized LDM-printed hierarchical porous scaffolds could promote the axonal growth and calcitonin gene-related peptide (CGRP) expression of sensory neurons and synergistically stimulate the osteogenic differentiation capacity of bone marrow mesenchymal stem cells (BMSCs) [142]. ARG arginase, BDNF brain-derived neurotrophic factor, Exos exosomes, FFT Fast Fourier Transform, GDNF glial-derived neurotrophic factor, iNOS inducible nitric oxide synthase, JUK3 c-Jun N-terminal kinase 3, LDM low temperature deposition modelling, MAPK mitogen-activated protein kinase, NF200 neurofilament 200, NGF nerve growth factor, NT3 neurotrophin-3, PPE PCL@PEI@exosome, PPEA PCL@PEI@exosome@aptamer, SMAD small mothers against decapentaplegic homolog, TGF-β transforming growth factor-β

However, nature exosomes still face challenges like large-scale production, stability, standard isolation, quality control, drug loading, and low targeting ability [133]. Engineered exosomes offer a solution by improving cell targeting and therapeutic effectiveness. Aptamer-engineered SC-exos were synthesized to phosphatidylserine on the cell surface to promote axonal fusion after axonal injury. The loaded SC-exos present on electrospun polycaprolactone membranes promoted axonal regeneration of dorsal root ganglia, tube formation by HUVECs, and healing of bone defects in vivo (Fig. 7b, c) [137]. In addition, NGF pre-stimulated MSC-containing porous scaffolds facilitated innervated bone regeneration in vivo (Fig. 7d) [142].

While these biomaterials have shown potential for promoting neurovascularized bone regeneration both in vitro and in vivo, further preclinical and clinical research is necessary to fully evaluate their osteogenic effects, application methods, and cost-effectiveness. Table 4 provides an overview of various material types and considerations for their selection [110, 114, 126, 127, 129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144]. Currently, research often lacks long-term biocompatibility assessments, focusing primarily on short-term effects on cell activity in vitro. Future studies should prioritize evaluating the metabolism of these materials in vivo, including their local and systemic toxicity to major organs. Another challenge lies in the choice of control materials in studies. Unmodified raw materials are frequently used as negative controls, while clinically established bone repair materials, such as Bio-Oss, BoneCeramic, and Puros, are not typically used as positive controls. Incorporating clinically used bone repair materials as positive controls will better validate the clinical potential of new materials.

Table 4 Overview of common material types for innervated and vascularized bone regeneration

Preclinical studies and clinical trials

Despite the recent surge in bench-top and ex vivo studies, there remains a notable scarcity of preclinical and clinical research on the utilization of neurovascular bone grafting materials. In 2013, Kamburoǧlu et al. [329] reported that the prefabricated neuro-osseous flap-maintained bone metabolic activity and promoted neovascularization; they demonstrated the beneficial effect of the prefabricated flap on bone repair for the first time. Subsequent animal studies showed that vascularized fibula flaps and sural nerve grafting are effective to reconstruct long bone defects with extensive soft-tissue damages [330,331,332]. In addition, rib composite flaps with intercostal nerves and internal thoracic vessels have been shown to be promising for mandibular defect reconstruction [333]. Recent research indicates that innervated and vascularized iliac bone flaps offer advantages such as preserving lower lip sensation and effectively reducing bone resorption [334,335,336]. Neurovascularized allogeneic bone may be a potential candidate for bone tissue regeneration.

Biomaterials used for innervated and vascularized bone regeneration include hydrogels, fiber spinning, and hard scaffolds. The bone regenerative capacity of these biomaterials in vivo is influenced by multiple factors, including the type, microstructure, and pore characteristics of the biomaterials [337]. Other important factors are the animal species used, the defect size, and the implantation period, all of which can affect the behavior of biomaterials in vivo. Most applications of these biomaterials have been tested in small animal models such as mice, rats, and rabbits. The rat cranial defect model with a 5 mm in diameter is the most commonly used, due to the small size of rats, their ease of handling, and low housing requirements. However, rats have small, long bones with thin, weak cortices and do not exhibit Haversian-type cortex remodeling, unlike larger animals [338]. Moreover, rats experience ongoing growth or modeling due to their open growth plates. The use of fractures, osteotomy, and defect sites, as well as methods of internal and external fixation in rat models, does not closely match those used in clinical settings [339]. Rabbits are also frequently used because their mid-diaphyseal bone mineral density is similar to that of humans [340]. Nonetheless, the higher bone turnover rate and faster skeletal changes in rabbits make them an undesirable choice as a model for autogenous bone and marrow harvesting, processing, or transplantation. Furthermore, rabbits are notoriously sensitive to glucocorticoid stimulation, resulting in very oily marrow with physical properties that are distinctly different from human marrow [341, 342]. In summary, clear differences in bone microstructure and remodeling between small animal models and humans have been documented. The selection of animal models should match their similarity to the intended clinical application and mimic the underlying bone biology seen in human clinical settings. Therefore, future in vivo studies should consider using large animal models, such as bama minipigs, which have bone characteristics more closely aligned with human bone [343]. Pigs are considered the preferred animal model compared to sheep, despite their denser trabecular network. They are described as a highly representative model of human bone regeneration processes in terms of anatomical and morphological features, healing capacity, and remodeling, as well as bone mineral density and concentration [344, 345]. However, pigs are often overlooked due to the complexity of handling and the relatively small size of their tibia and femora [346].

Moreover, the presented study outcomes indicate that specific issues remain to be addressed before clinical translation. Medical-grade polycaprolactone-TCP scaffolds, as a second-generation scaffold, are currently in the Food and Drug Administration (FDA) preapproval stage. In a long-term (12-month) preclinical study, the scaffold failed to induce defect consolidation in a segmental tibial animal model (sheep). However, scaffolds combined with MSCs or recombinant human BMP-7 (rhBMP-7) showed significantly greater bone formation and superior strength compared to the autograft [347]. Additionally, various growth factors, such as BMP-2 and BMP-7, have been widely used in clinics and are approved by the FDA [348, 349]. The use of BMP promotes bone integration and improves the success rate of surgeries. However, BMP-2 has been reported as a dual-function cytokine that promotes ectopic bone formation through osteoinductive action and induces neuroinflammation [350]. Therefore, many challenges remain in translating preclinical studies to clinical trials.

Despite the encouraging results, all experiments in materials for regenerating innervated and vascularized bone (section "Materials for regenerating innervated and vascularized bone") were conducted in animal models, and none of these findings have yet been translated into human clinical trials. The unavailability of human clinical studies on neurovascular biomaterials for human bone regeneration highlights several challenges hindering clinical translation. Current studies primarily use small animals, such as mice, rats, rabbits, and guinea pigs. However, large animal models offer significant advantages due to their closer anatomical similarity to humans. The limited use of neurovascularization materials in preclinical studies, especially in large animals, poses a significant challenge. Previous study has shown significant heterogeneity in current surgical methods for inducing bone defects in rats, which reduces the reproducibility and comparability of preclinical studies [351]. There is still a need to develop standardized approaches for creating bone defect animal models to reliably verify the osteogenic performance of biomaterials. Further challenges include demonstrating and assessing dynamic material properties. The long-term biocompatibility and biodegradation rates of hard scaffolds, hydrogels, and spinners in vivo may vary due to their differing material properties. Therefore, it is necessary to design specific types of materials to match the corresponding healing rate of bone, depending on the characteristics of the defect sites. Moreover, there is a lack of comparative studies between neurovascular biomaterials and commercially available bone repair materials used in clinical settings. The benefits of incorporating novel material design features that result in only minor improvements in bone regeneration should be carefully weighed against the challenges of obtaining regulatory approval.

In action, ethical considerations, regulatory hurdles, and the high cost of developing and testing new treatments further complicate the progression from animal models to human trials. Many new applications fail to gain approval, even with positive clinical results, due to concerns from regulatory agencies. These concerns often include insufficient justification for clinical comparator selection, inappropriate endpoint design, and inadequate clinical data analysis methodology. For example, Opaxio’s European approval was denied despite a 42-day improvement in overall survival compared to the comparator. The developer of Opaxio later withdrew the application after European officials raised concerns over the clinical trial regimen in 2009 [352]. This example underscores the importance for both pharmaceutical companies and organizations developing biological materials to engage with regulators early and throughout the biomaterial development process. By seeking advice and addressing potential concerns proactively, companies can mitigate the risk of non-approval or delays.

To overcome these regulatory hurdles, the authors propose three solutions:

  1. (1)

    Enhanced preclinical testing: improving animal models to better simulate human bone biology, particularly by using large animal models like bama minipigs, would provide more reliable data before moving to human trials. This would address concerns about the limited applicability of small animal models.

  2. (2)

    Adaptive clinical trials: implementing adaptive trial designs allows for modifications based on interim results, enhancing both the safety and efficacy assessments while maintaining scientific rigor. This approach enables trials to respond dynamically to evolving data, potentially reducing the time to approval.

  3. (3)

    Collaborative approaches with regulators: establishing close collaboration with regulatory bodies, such as the FDA, from the early stages of development is crucial. Engaging in pre-submission meetings and incorporating feedback on trial design and data collection can prevent delays and streamline the approval process.

By addressing these challenges and incorporating these solutions, the clinical application of neurovascular bone repair materials may be advanced more efficiently.

Specific solutions also include identifying the specific requirements of targeted patients during the initial design stages. Due to the various causes of human bone defects, the clinical application of biological materials must take into account and compensate for patient-specific factors, as bone tissue can also be affected by the disease-related microenvironment. For example, in patients with bone defects related to malignancy, a biomaterials-only approach to bone tissue engineering may be preferable, as proliferation-stimulating biomolecules, such as growth factors, should not be introduced into former tumor sites [348]. To further enhance the neurovascularization capability of bone tissue affected by tumors, it is crucial to design biomaterials that incorporate additional elements beyond NGF, such as inorganic ions and exosomes, for optional use. Therefore, these materials should be subclassified based on the patient’s specific conditions or complications.

Involving specialist physicians with experiential knowledge of patient needs, clinical realities, and potential safety concerns is also crucial. In-depth communication between with biomaterial designers and clinicians is of significance to prevent cognitive differences as a result of common-sense understanding errors, differences in professional opinions, etc. Furthermore, different surgical methods for the same surgery, such as open suture or arthroscopic injection for implantation of cartilage implants, have different requirements for biomaterial structure and properties (e.g., viscoelasticity, strength, viscosity, injectability). Therefore, biomaterial researchers should establish a close relationship with clinicians, observe the diagnosis and treatment procedures in clinical practice, get to know the key requirements of clinical translation, and further seek solutions together [353].

In addition, it is essential to ensure that the designed materials are compatible with the evaluation standards of governmental regulatory agencies. For example, medical products in the field of tissue-engineered cartilage repair usually contain a combination of scaffold materials with cytokines and/or cells; different countries have different definitions of combined products. In China, combined products will be regulated as a single entity, whether it is a drug or a medical device based on certain products. In addition, biologics or cell- and tissue-based components are not separated from the drug class, unlike in Japan or the United States. In this process, the developers first determine the primary mode of action of the product, which determines whether the product’s properties are drug-led, bio-led, or medical-device-led. Products with different attributes have different requirements in production quality system, risk assessment, clinical evaluation, etc. Therefore, attention to which product type of neurovascularized bone regeneration material belongs to is crucial for governmental regulation, which profoundly affects the progress of product development and marketing [354].

Conclusions and future perspectives

The mechanism and role of the neurovascular system in bone regeneration have garnered significant attention. Bones are covered by neural and vascular networks essential for the development, remodeling, and repair. Nerves promote blood vessel regeneration by secreting neurotransmitters and participate in various bone tissue activities. Blood vessels provide oxygen and nutrients to nerve fibers and bone tissue. Together, nerve fibers and blood vessels maintain the microenvironment for bone tissue regeneration, addressing clinical issues such as fractures and non-unions.

However, the crucial role of the neural networks in promoting bone regeneration is often overlooked in the design of biomaterials, which may result in delayed or compromised healing. The exact mechanism of how the neurovascular system regulates the bone defect microenvironment remains unclear. With the advent of neurovascularized bone regeneration materials, design strategies are diversifying, including neurotrophic factors, peptides, RNA, inorganic ions, and exosomes to promote neurovascularized bone tissue regeneration.

Despite advancements, most research has not yet reached the stage of clinical trials. Key factors include the lack of preclinical translational studies in large animals, the complexity of replicating the intricate neurovascular structures in engineered grafts, and the need for robust and reproducible methods to ensure safety and efficacy. Additionally, the impact of biomaterial properties on the intrabony neurovascular system is often overlooked. Effective neurovascularization in bone regeneration requires biomaterials that create a microenvironment meeting the functional needs of nerves, blood vessels, and bone tissue. However, regeneration of multiple tissues remains a bottleneck in bone tissue engineering, as bone regenerates better than nerves. Architectural design must address the contradictory requirements for regenerating the three tissues, such as stiffness, roughness, pore size and porosity, and conductivity.

Achieving clinical translation involves controlling the orderly growth of two tissues into the bone defect area to exert osteogenic effects. Future prospects for neurovascularized bone regeneration materials to facilitate clinical translation include:

  1. (1)

    As an emerging technology, organ-on-a-chip refers to a biomimetic micro-engineered system that mimics the microenvironment of native tissue and organs, based on a microfluidic chip that combines biology, materials science, and engineering. A microfluidic osteogenesis-on-a-chip device has been developed to simulate a 3D environment and fluid shear stresses [355]. Advances in microfluidic device fabrication techniques hold the potential to create more realistic and sophisticated pre-established peripheral vascular networks in bone grafts before implantation. However, challenges include the source of cells, scalability, standardization of manufacturing processes, and the limitations of chip size [356]. Moreover, no regulatory standards currently exist for organ-on-a-chip systems. It is essential for governmental regulatory agencies to develop ethical and regulatory guidelines to promote the advancement of this technology.

  2. (2)

    Osteo-organoids, which combine bioactive factors, scaffolds, and functional cells, have been used for various bone defect repairs [357]. An organoid-based strategy combined with 4D printing technology is expected to precisely fit the geometry of bone defects over time. Functional transformation during the post-printing stage may coordinate intrabony neurovascular system regeneration and facilitate dynamic bone remodeling. However, unlike single-material, non-cellular 3D-printed bone scaffolds that have been applied in clinical practice, bio-printed organoids are still in the early stages of development. Not only does the production technology need further refinement, but the storage and transport of organoids also present challenges. The ex vivo expansion of specific stem cells requires large-scale cell proliferation, which demands stringent expertise in both hardware and operational aspects [358]. Currently, organoid storage and transport rely primarily on long-term cryopreservation and short-term tissue preservation solutions, which are difficult to apply in clinical practice. Future in vivo studies should focus on developing more effective methods for organoid storage and transport.

  3. (3)

    Real-time monitoring of bone regeneration and resorption is currently achieved through techniques like fluorescence probes, prussian blue nanoparticles, superparamagnetic iron oxide nanoparticles, and sensors [359,360,361,362]. However, these technologies do not enable real-time monitoring of neurovascular regeneration in bone. Advanced imaging techniques, such as two-photon laser scanning microscopy combined with fluorescence technology, hold potential for real-time monitoring of blood vessels and nerves, though they require exposure of the monitoring site. Presently, sensing bandages and electroactive dressings can non-invasively monitor the healing process of skin defects by measuring indicators such as pH, resistance, temperature, and pressure [363,364,365]. Whether similar techniques can be applied to bone regeneration remains an area for future exploration. Advances in these technologies could facilitate more comprehensive efficacy evaluations of neurovascularized materials, leading to more accurate preclinical assessments of drugs and bone implant biomaterials.

  4. (4)

    Designing biomaterials that sequentially release bioactive factors is advantageous for simulating the physiological process of neurovascular regeneration during bone healing. Whether these bioactive factors could induce heterotopic ossification or cause biological complications remains an open question. Previous studies have reported a relationship between neurotrophins and pain during fracture healing [366,367,368]. Additionally, TrkA, the specific binding receptor for NGF, has been proven to be a potent carcinogenic driver when overexpressed [369]. Thus, it is crucial to carefully regulate the dosage and release rate of neurotrophins during the design of neurovascular biomaterials, particularly for patients with bone defects associated with malignancy [348]. Moreover, whether long-term accumulation of biomaterial degradation products could lead to complications also requires thorough investigation in preclinical studies.

  5. (5)

    Artificial intelligence (AI) and machine learning can automatically identify and extract key features from patients’ medical imaging data, such as the size, shape, and location of bone defects, providing precise guidance for scaffold design [370]. At the same time, AI systems can autonomously learn and optimize scaffold design parameters such as structural strength, biocompatibility, and biomechanical performance [371]. In addition, continuous ethical review, the establishment of data-sharing platforms, and the definition of data standards are crucial for advancing the role of AI in biomaterial research [372, 373]. To date, only a few AI models have been approved by the FDA in the field of orthopedic diseases [374,375,376]. Therefore, more robust datasets and AI models are needed to predict the formation of blood vessels and nerves, which would improve the design and application of biomaterials.

Advancement of these technologies has opened new vistas for the rational design of bioactive materials. In the future, scientists have to pay more attention to the crosstalk between blood vessels and nerves in bone, and their interaction with bone tissue. This is significant for guiding the design and clinical transformation of intraosseous biomaterials.

Availability of data and materials

Not applicable.

Abbreviations

AI:

Artificial intelligence

BMP-2:

Bone morphogenetic protein 2

BMSCs:

Bone marrow mesenchymal stem cells

Ce-ESMs:

Cerium (III, IV) oxide-mineralized ESMs

CGRP:

Calcitonin gene-related peptide

CNS:

Central nervous system

CPC:

Calcium phosphate cement

ECM:

Extracellular matrix

ECs:

Endothelial cells

FAK:

Focal adhesion kinase

FDA:

Food and Drug Administration

Flt-1:

Fms-related receptor tyrosine kinase 1

GelMA:

Gelatin methacryloyl

HA:

Hydroxyapatite

HIF-1α:

Hypoxia-inducible factor 1-alpha

HPAA:

High-molecular-weight-polyacrylic acid

HUVECs:

Human umbilical vein endothelial cells

IL:

Interleukin

LMs:

Laminins

miRNAs:

MicroRNAs

MMP:

Matrix metalloproteinase

MSCs:

Mesenchymal stem cells

mTOR:

Mammalian target of rapamycin

nano β-TCP:

Nano beta-tricalcium phosphate

NGF:

Nerve growth factor

NT3:

Neurotrophin-3

NT4/5:

Neurotrophin-4/5

OCN:

Osteocalcin

p75NTR:

P75 neurotrophic factor receptor

PDGF-BB:

Platelet-derived growth factor type BB

PEG:

Polyethyleneglycol

PI3K:

Phosphatidylinositide 3-kinase

PWH:

Piezoelectric WH

REDV:

Arg-Gly-Asp-Val

rhBMP-7:

Recombinant human BMP-7

RNAs:

Ribonucleic acids

Runx2:

Runt-related transcription factor 2

SCs:

Schwann cells

Sema:

Semaphorin

siRNA:

Small interfering RNA

SLIT3:

Slit guidance ligand 3

TrkA:

Tyrosine kinase receptor A

VEGF:

Vascular endothelial growth factor

WH:

Whitlockite

YIGSR:

Tyr-Lle-Gly-Ser-Arg

References

  1. Qin Q, Lee S, Patel N, Walden K, Gomez-Salazar M, Levi B, et al. Neurovascular coupling in bone regeneration. Exp Mol Med. 2022;54(11):1844–9.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  2. Armiento AR, Hatt LP, Sanchez Rosenberg G, Thompson K, Stoddart MJ. Functional biomaterials for bone regeneration: a lesson in complex biology. Adv Funct Mater. 2020;30(44):1909874.

    Article  CAS  Google Scholar 

  3. Jia Z, Xu X, Zhu D, Zheng Y. Design, printing, and engineering of regenerative biomaterials for personalized bone healthcare. Prog Mater Sci. 2023;134:101072.

    Article  CAS  Google Scholar 

  4. Wang B, Feng C, Liu Y, Mi F, Dong J. Recent advances in biofunctional guided bone regeneration materials for repairing defective alveolar and maxillofacial bone: a review. Jpn Dent Sci Rev. 2022;58:233–48.

    Article  PubMed  PubMed Central  Google Scholar 

  5. de Melo PD, Habibovic P. Biomineralization-inspired material design for bone regeneration. Adv Healthc Mater. 2018;7(22):e1800700.

    Article  Google Scholar 

  6. Andrew H, Schmidt M. Autologous bone graft: is it still the gold standard? Injury. 2021;52:18–22.

    Article  Google Scholar 

  7. Zhang Y, Liu X, Zeng L, Zhang J, Zuo J, Zou J, et al. Tissue engineering: polymer fiber scaffolds for bone and cartilage tissue engineering. Adv Funct Mater. 2019;29(36):1970246.

    Article  Google Scholar 

  8. Tang S, Dong Z, Ke X, Luo J, Li J. Advances in biomineralization-inspired materials for hard tissue repair. Int J Oral Sci. 2021;13(1):42.

    Article  PubMed  PubMed Central  Google Scholar 

  9. Qu H, Fu H, Han Z, Sun Y. Biomaterials for bone tissue engineering scaffolds: a review. RSC Adv. 2019;9(45):26252–62.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  10. Hara-Irie F, Amizuka N, Ozawa H. Immunohistochemical and ultrastructural localization of CGRP-positive nerve fibers at the epiphyseal trabecules facing the growth plate of rat femurs. Bone. 1996;18(1):29–39.

    Article  PubMed  CAS  Google Scholar 

  11. Mach DB, Rogers SD, Sabino MC, Luger NM, Schwei MJ, Pomonis JD, et al. Origins of skeletal pain: sensory and sympathetic innervation of the mouse femur. Neuroscience. 2002;113(1):155–66.

    Article  PubMed  CAS  Google Scholar 

  12. Deng AF, Wang FX, Wang SC, Zhang YZ, Bai L, Su JC. Bone-organ axes: bidirectional crosstalk. Mil Med Res. 2024;11(1):37.

    PubMed  PubMed Central  Google Scholar 

  13. Burger MG, Grosso A, Briquez PS, Born GME, Lunger A, Schrenk F, et al. Robust coupling of angiogenesis and osteogenesis by VEGF-decorated matrices for bone regeneration. Acta Biomater. 2022;149:111–25.

    Article  PubMed  CAS  Google Scholar 

  14. Tuckermann J, Adams RH. The endothelium-bone axis in development, homeostasis and bone and joint disease. Nat Rev Rheumatol. 2021;17(10):608–20.

    Article  PubMed  PubMed Central  Google Scholar 

  15. Behera J, Kumar A, Voor MJ, Tyagi N. Exosomal lncRNA-H19 promotes osteogenesis and angiogenesis through mediating Angpt1/Tie2-NO signaling in CBS-heterozygous mice. Theranostics. 2021;11(16):7715.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  16. Hann SY, Cui H, Esworthy T, Zhou X, Lee S, Plesniak MW, et al. Dual 3D printing for vascularized bone tissue regeneration. Acta Biomater. 2021;123:263–74.

    Article  PubMed  CAS  Google Scholar 

  17. Sun W, Ye B, Chen S, Zeng L, Lu H, Wan Y, et al. Neuro-bone tissue engineering: emerging mechanisms, potential strategies, and current challenges. Bone Res. 2023;11(1):65.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  18. Wan QQ, Qin WP, Ma YX, Shen MJ, Li J, Zhang ZB, et al. Crosstalk between bone and nerves within bone. Adv Sci (Weinh). 2021;8(7):2003390.

    Article  PubMed  CAS  Google Scholar 

  19. Lv X, Gao F, Li TP, Xue P, Wang X, Wan M, et al. Correction: skeleton interoception regulates bone and fat metabolism through hypothalamic neuroendocrine NPY. Elife. 2023;12:e85738.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Mei H, Wu Y, Feng Q, Li X, Zhou J, Jiang F, et al. The interplay between the nerves and skeleton: a 30-year bibliometric analysis. Ann Transl Med. 2023;11(1):9.

    Article  PubMed  PubMed Central  Google Scholar 

  21. Leroux A, Paiva Dos Santos B, Leng J, Oliveira H, Amedee J. Sensory neurons from dorsal root ganglia regulate endothelial cell function in extracellular matrix remodelling. Cell Commun Signal. 2020;18(1):162.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  22. Gaete PS, Lillo MA, Puebla M, Poblete I, Figueroa XF. CGRP signalling inhibits NO production through pannexin-1 channel activation in endothelial cells. Sci Rep. 2019;9(1):7932.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Taberner L, Bañón A, Alsina B. Sensory neuroblast quiescence depends on vascular cytoneme contacts and sensory neuronal differentiation requires initiation of blood flow. Cell Rep. 2020;32(2):107903.

    Article  PubMed  CAS  Google Scholar 

  24. Yu YL, Wu JJ, Lin CC, Qin X, Tay FR, Miao L, et al. Elimination of methicillin-resistant Staphylococcus aureus biofilms on titanium implants via photothermally-triggered nitric oxide and immunotherapy for enhanced osseointegration. Mil Med Res. 2023;10(1):21.

    PubMed  PubMed Central  CAS  Google Scholar 

  25. Zhang Y, Chen S, Qin X, Guo A, Li K, Chen L, et al. A versatile chitosan-based hydrogel accelerates infected wound healing via bacterial elimination, antioxidation, immunoregulation, and angiogenesis. Adv Healthc Mater. 2024;13(19):e2400318.

    Article  PubMed  Google Scholar 

  26. Wu J, Shen P, Qin X, Yang Y, Lin C, Li X, et al. Self-supply of H2O2 and O2 by a composite nanogenerator for chemodynamic therapy/hypoxia improvement and rapid therapy of biofilm-infected wounds. Chem Eng J. 2023;459:141507.

    Article  CAS  Google Scholar 

  27. Huelsboemer L, Knoedler L, Kochen A, Yu CT, Hosseini H, Hollmann KS, et al. Cellular therapeutics and immunotherapies in wound healing - on the pulse of time? Mil Med Res. 2024;11(1):23.

    PubMed  PubMed Central  Google Scholar 

  28. Li J, Zhang Z, Tang J, Hou Z, Li L, Li B. Emerging roles of nerve-bone axis in modulating skeletal system. Med Res Rev. 2024;44(4):1867–903.

    Article  PubMed  Google Scholar 

  29. Elefteriou F. Impact of the autonomic nervous system on the skeleton. Physiol Rev. 2018;98(3):1083–112.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  30. Wang XD, Li SY, Zhang SJ, Gupta A, Zhang CP, Wang L. The neural system regulates bone homeostasis via mesenchymal stem cells: a translational approach. Theranostics. 2020;10(11):4839–50.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  31. Liu S, Liu S, Li S, Liang B, Han X, Liang Y, et al. Nerves within bone and their application in tissue engineering of bone regeneration. Front Neurol. 2023;13:1085560.

    Article  PubMed  PubMed Central  Google Scholar 

  32. Damiati LA, El Soury M. Bone-nerve crosstalk: a new state for neuralizing bone tissue engineering-a mini review. Front Med (Lausanne). 2024;11:1386683.

    Article  PubMed  Google Scholar 

  33. Asghar A, Kumar A, Kant Narayan R, Naaz S. Is the cortical capillary renamed as the transcortical vessel in diaphyseal vascularity? Anat Rec (Hoboken). 2020;303(11):2774–84.

    Article  PubMed  Google Scholar 

  34. Thai J, Fuller-Jackson JP, Ivanusic JJ. Using tissue clearing and light sheet fluorescence microscopy for the three-dimensional analysis of sensory and sympathetic nerve endings that innervate bone and dental tissue of mice. J Comp Neurol. 2024;532(1):e25582.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  35. Xu Z, Kusumbe AP, Cai H, Wan Q, Chen J. Type H blood vessels in coupling angiogenesis-osteogenesis and its application in bone tissue engineering. J Biomed Mater Res B Appl Biomater. 2023;111(7):1434–46.

    Article  PubMed  CAS  Google Scholar 

  36. Zhang Z, Hao Z, Xian C, Fang Y, Cheng B, Wu J, et al. Neuro-bone tissue engineering: multiple potential translational strategies between nerve and bone. Acta Biomater. 2022;153:1–12.

    Article  PubMed  Google Scholar 

  37. Salhotra A, Shah HN, Levi B, Longaker MT. Mechanisms of bone development and repair. Nat Rev Mol Cell Biol. 2020;21(11):696–711.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  38. Watson EC, Adams RH. Biology of bone: the vasculature of the skeletal system. Cold Spring Harb Perspect Med. 2018;8(7):a031559.

    Article  PubMed  PubMed Central  Google Scholar 

  39. Berendsen AD, Olsen BR. Bone development. Bone. 2015;80:14–8.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  40. Kozhemyakina E, Lassar AB, Zelzer E. A pathway to bone: signaling molecules and transcription factors involved in chondrocyte development and maturation. Development. 2015;142(5):817–31.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  41. Lazarus A, Del-Moral PM, Ilovich O, Mishani E, Warburton D, Keshet E. A perfusion-independent role of blood vessels in determining branching stereotypy of lung airways. Development. 2011;138(11):2359–68.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  42. Cleaver O, Dor Y. Vascular instruction of pancreas development. Development. 2012;139(16):2833–43.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  43. Kumar A, Sood A, Singhmar R, Mishra YK, Thakur VK, Han SS. Manufacturing functional hydrogels for inducing angiogenic-osteogenic coupled progressions in hard tissue repairs: prospects and challenges. Biomater Sci. 2022;10(19):5472–97.

    Article  PubMed  CAS  Google Scholar 

  44. Gadomski S, Fielding C, García-García A, Korn C, Kapeni C, Ashraf S, et al. A cholinergic neuroskeletal interface promotes bone formation during postnatal growth and exercise. Cell Stem Cell. 2022;29(4):528-44.e9.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  45. Huber AB, Kolodkin AL, Ginty DD, Cloutier JF. Signaling at the growth cone: ligand-receptor complexes and the control of axon growth and guidance. Annu Rev Neurosci. 2003;26:509–63.

    Article  PubMed  CAS  Google Scholar 

  46. Dickson BJ. Molecular mechanisms of axon guidance. Science. 2002;298(5600):1959–64.

    Article  PubMed  CAS  Google Scholar 

  47. Tessier-Lavigne M, Goodman CS. The molecular biology of axon guidance. Science. 1996;274(5290):1123–33.

    Article  PubMed  CAS  Google Scholar 

  48. Honma Y, Araki T, Gianino S, Bruce A, Heuckeroth R, Johnson E, et al. Artemin is a vascular-derived neurotropic factor for developing sympathetic neurons. Neuron. 2002;35(2):267–82.

    Article  PubMed  CAS  Google Scholar 

  49. Mukouyama YS, Shin D, Britsch S, Taniguchi M, Anderson DJ. Sensory nerves determine the pattern of arterial differentiation and blood vessel branching in the skin. Cell. 2002;109(6):693–705.

    Article  PubMed  CAS  Google Scholar 

  50. Dolan CP, Yan M, Zimmel K, Yang TJ, Leininger E, Dawson LA, et al. Axonal regrowth is impaired during digit tip regeneration in mice. Dev Biol. 2019;445(2):237–44.

    Article  PubMed  CAS  Google Scholar 

  51. Guo Q, Chen N, Patel K, Wan M, Zheng J, Cao X. Unloading-induced skeletal interoception alters hypothalamic signaling to promote bone loss and fat metabolism. Adv Sci (Weinh). 2023;10(35):e2305042.

    Article  PubMed  Google Scholar 

  52. Jung WC, Levesque JP, Ruitenberg MJ. It takes nerve to fight back: the significance of neural innervation of the bone marrow and spleen for immune function. Semin Cell Dev Biol. 2017;61:60–70.

    Article  PubMed  CAS  Google Scholar 

  53. Fukuda T, Takeda S, Xu R, Ochi H, Sunamura S, Sato T, et al. Sema3A regulates bone-mass accrual through sensory innervations. Nature. 2013;497(7450):490–3.

    Article  PubMed  CAS  Google Scholar 

  54. Tanaka K, Hirai T, Kodama D, Kondo H, Hamamura K, Togari A. α1B-adrenoceptor signalling regulates bone formation through the up-regulation of CCAAT/enhancer-binding protein δ expression in osteoblasts. Br J Pharmacol. 2016;173(6):1058–69.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  55. Bajayo A, Bar A, Denes A, Bachar M, Kram V, Attar-Namdar M, et al. Skeletal parasympathetic innervation communicates central IL-1 signals regulating bone mass accrual. Proc Natl Acad Sci U S A. 2012;109(38):15455–60.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  56. Appelt J, Baranowsky A, Jahn D, Yorgan T, Kohli P, Otto E, et al. The neuropeptide calcitonin gene-related peptide alpha is essential for bone healing. EBioMedicine. 2020;59:102970.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  57. Mu C, Hu Y, Hou Y, Li M, He Y, Shen X, et al. Substance P-embedded multilayer on titanium substrates promotes local osseointegration via MSC recruitment. J Mater Chem B. 2020;8(6):1212–22.

    Article  PubMed  CAS  Google Scholar 

  58. Baldock PA, Lee NJ, Driessler F, Lin S, Allison S, Stehrer B, et al. Neuropeptide Y knockout mice reveal a central role of NPY in the coordination of bone mass to body weight. PLoS ONE. 2009;4(12):e8415.

    Article  PubMed  PubMed Central  Google Scholar 

  59. Kenan S, Onur ÖD, Solakoğlu S, Kotil T, Ramazanoğlu M, Çelik HH, et al. Investigation of the effects of semaphorin 3A on new bone formation in a rat calvarial defect model. J Craniomaxillofac Surg. 2019;47(3):473–83.

    Article  PubMed  Google Scholar 

  60. Negishi-Koga T, Shinohara M, Komatsu N, Bito H, Kodama T, Friedel RH, et al. Suppression of bone formation by osteoclastic expression of semaphorin 4D. Nat Med. 2011;17(11):1473–80.

    Article  PubMed  CAS  Google Scholar 

  61. Kim BJ, Lee YS, Lee SY, Baek WY, Choi YJ, Moon SA, et al. Osteoclast-secreted SLIT3 coordinates bone resorption and formation. J Clin Invest. 2018;128(4):1429–41.

    Article  PubMed  PubMed Central  Google Scholar 

  62. Rivera KO, Russo F, Boileau RM, Tomlinson RE, Miclau T, Marcucio RS, et al. Local injections of β-NGF accelerates endochondral fracture repair by promoting cartilage to bone conversion. Sci Rep. 2020;10(1):22241.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  63. Park EJ, Truong VL, Jeong WS, Min WK. Brain-derived neurotrophic factor (BDNF) enhances osteogenesis and may improve bone microarchitecture in an ovariectomized rat model. Cells. 2024;13(6):518.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  64. Su YW, Chung R, Ruan CS, Chim SM, Kuek V, Dwivedi PP, et al. Neurotrophin-3 induces BMP-2 and VEGF activities and promotes the bony repair of injured growth plate cartilage and bone in rats. J Bone Miner Res. 2016;31(6):1258–74.

    Article  PubMed  CAS  Google Scholar 

  65. Takeda S, Elefteriou F, Levasseur R, Liu X, Zhao L, Parker KL, et al. Leptin regulates bone formation via the sympathetic nervous system. Cell. 2002;111(3):305–17.

    Article  PubMed  CAS  Google Scholar 

  66. Zhu S, Zhu J, Zhen G, Hu Y, An S, Li Y, et al. Subchondral bone osteoclasts induce sensory innervation and osteoarthritis pain. J Clin Investig. 2019;129(3):1076–93.

    Article  PubMed  PubMed Central  Google Scholar 

  67. Kovács B, Vajda E, Nagy EE. Regulatory effects and interactions of the Wnt and OPG-RANKL-RANK signaling at the bone-cartilage interface in osteoarthritis. Int J Mol Sci. 2019;20(18):4653.

    Article  PubMed  PubMed Central  Google Scholar 

  68. Li J, Ding Z, Li Y, Wang W, Wang J, Yu H, et al. BMSCs-derived exosomes ameliorate pain via abrogation of aberrant nerve invasion in subchondral bone in lumbar facet joint osteoarthritis. J Orthop Res. 2020;38(3):670–9.

    Article  PubMed  CAS  Google Scholar 

  69. Romero-Morelos P, Ruvalcaba-Paredes E, Garciadiego-Cázares D, Pérez-Santos M, Reyes-Long S, Alfaro-Rodriguez A, et al. Neurophysiological mechanisms related to pain management in bone tumors. Curr Neuropharmacol. 2021;19(3):308–19.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  70. Shepherd AJ, Mickle AD, McIlvried LA, Gereau RWt, Mohapatra DP. Parathyroid hormone-related peptide activates and modulates TRPV1 channel in human DRG neurons. Eur J Pain. 2018;22(9):1685–90.

    Article  PubMed  CAS  Google Scholar 

  71. Kusumbe AP, Ramasamy SK, Itkin T, Mäe MA, Langen UH, Betsholtz C, et al. Age-dependent modulation of vascular niches for haematopoietic stem cells. Nature. 2016;532(7599):380–4.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  72. Peng Y, Wu S, Li Y, Crane JL. Type H blood vessels in bone modeling and remodeling. Theranostics. 2020;10(1):426–36.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  73. Xie H, Cui Z, Wang L, Xia Z, Hu Y, Xian L, et al. PDGF-BB secreted by preosteoclasts induces angiogenesis during coupling with osteogenesis. Nat Med. 2014;20(11):1270–8.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  74. Xu R, Yallowitz A, Qin A, Wu Z, Shin DY, Kim JM, et al. Targeting skeletal endothelium to ameliorate bone loss. Nat Med. 2018;24(6):823–33.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  75. Jones DT, Harris AL. Identification of novel small-molecule inhibitors of hypoxia-inducible factor-1 transactivation and DNA binding. Mol Cancer Ther. 2006;5(9):2193–202.

    Article  PubMed  CAS  Google Scholar 

  76. Chen W, Wu P, Yu F, Luo G, Qing L, Tang J. HIF-1α regulates bone homeostasis and angiogenesis, participating in the occurrence of bone metabolic diseases. Cells. 2022;11(22):3552.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  77. Song S, Zhang G, Chen X, Zheng J, Liu X, Wang Y, et al. HIF-1α increases the osteogenic capacity of ADSCs by coupling angiogenesis and osteogenesis via the HIF-1α/VEGF/Akt/mTOR signaling pathway. J Nanobiotechnol. 2023;21(1):257.

    Article  CAS  Google Scholar 

  78. Gao B, Lin X, Jing H, Fan J, Ji C, Jie Q, et al. Local delivery of tetramethylpyrazine eliminates the senescent phenotype of bone marrow mesenchymal stromal cells and creates an anti-inflammatory and angiogenic environment in aging mice. Aging Cell. 2018;17(3):e12741.

    Article  PubMed  PubMed Central  Google Scholar 

  79. Jiang L, Sheng K, Wang C, Xue D, Pan Z. The effect of MMP-2 inhibitor 1 on osteogenesis and angiogenesis during bone regeneration. Front Cell Dev Biol. 2021;8:596783.

    Article  PubMed  PubMed Central  Google Scholar 

  80. Zhang D, Du J, Yu M, Suo L. Urine-derived stem cells-extracellular vesicles ameliorate diabetic osteoporosis through HDAC4/HIF-1α/VEGFA axis by delivering microRNA-26a-5p. Cell Biol Toxicol. 2023;39(5):2243–57.

    Article  PubMed  CAS  Google Scholar 

  81. Park SJ, Lee JY, Lee SH, Koh JM, Kim BJ. SLIT2 inhibits osteoclastogenesis and bone resorption by suppression of Cdc42 activity. Biochem Biophys Res Commun. 2019;514(3):868–74.

    Article  PubMed  CAS  Google Scholar 

  82. Jiang L, Sun J, Huang D. Role of slit/robo signaling pathway in bone metabolism. Int J Biol Sci. 2022;18(3):1303.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  83. González-Hernández A, Marichal-Cancino BA, Lozano-Cuenca J, López-Canales JS, Muñoz-Islas E, Ramírez-Rosas MB, et al. Heteroreceptors modulating CGRP release at neurovascular junction: potential therapeutic implications on some vascular-related diseases. Biomed Res Int. 2016;2016(1):2056786.

    PubMed  PubMed Central  Google Scholar 

  84. Xu HK, Liu JX, Zheng CX, Liu L, Ma C, Tian JY, et al. Region-specific sympatho-adrenergic regulation of specialized vasculature in bone homeostasis and regeneration. iScience. 2023;26(9):107455.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  85. Zhou R, Zhou H, Rui L, Xu J. Bone loss and osteoporosis are associated with conversion from mild cognitive impairment to Alzheimer’s disease. Curr Alzheimer Res. 2014;11(7):706–13.

    Article  PubMed  CAS  Google Scholar 

  86. Liu D, Zhou H, Tao Y, Tan J, Chen L, Huang H, et al. Alzheimer’s disease is associated with increased risk of osteoporosis: the Chongqing aging study. Curr Alzheimer Res. 2016;13(10):1165–72.

    Article  PubMed  CAS  Google Scholar 

  87. Idelevich A, Baron R. Brain to bone: what is the contribution of the brain to skeletal homeostasis? Bone. 2018;115:31–42.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  88. Wadhwa R, Kumar M, Talegaonkar S, Vohora D. Serotonin reuptake inhibitors and bone health: a review of clinical studies and plausible mechanisms. Osteoporos Sarcopenia. 2017;3(2):75–81.

    Article  PubMed  PubMed Central  Google Scholar 

  89. Oury F, Yadav VK, Wang Y, Zhou B, Liu XS, Guo XE, et al. CREB mediates brain serotonin regulation of bone mass through its expression in ventromedial hypothalamic neurons. Genes Dev. 2010;24(20):2330–42.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  90. Dimitriou R, Jones E, McGonagle D, Giannoudis PV. Bone regeneration: current concepts and future directions. BMC Med. 2011;9:66.

    Article  PubMed  PubMed Central  Google Scholar 

  91. Claes L, Recknagel S, Ignatius A. Fracture healing under healthy and inflammatory conditions. Nat Rev Rheumatol. 2012;8(3):133–43.

    Article  PubMed  CAS  Google Scholar 

  92. Andrew JG, Andrew SM, Freemont AJ, Marsh DR. Inflammatory cells in normal human fracture healing. Acta Paediatr Scand. 1994;65(4):462–6.

    CAS  Google Scholar 

  93. Goerke SM, Obermeyer J, Plaha J, Stark GB, Finkenzeller G. Endothelial progenitor cells from peripheral blood support bone regeneration by provoking an angiogenic response. Microvasc Res. 2015;98:40–7.

    Article  PubMed  CAS  Google Scholar 

  94. Ramasamy SK, Kusumbe AP, Wang L, Adams RH. Endothelial Notch activity promotes angiogenesis and osteogenesis in bone. Nature. 2014;507(7492):376–80.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  95. Malda J, Klein TJ, Upton Z. The roles of hypoxia in the in vitro engineering of tissues. Tissue Eng. 2007;13(9):2153–62.

    Article  PubMed  CAS  Google Scholar 

  96. Faller DV. Endothelial cell responses to hypoxic stress. Clin Exp Pharmacol Physiol. 1999;26(1):74–84.

    Article  PubMed  CAS  Google Scholar 

  97. Schindeler A, McDonald MM, Bokko P, Little DG. Bone remodeling during fracture repair: the cellular picture. Semin Cell Dev Biol. 2008;19(5):459–66.

    Article  PubMed  CAS  Google Scholar 

  98. Street J, Bao M, DeGuzman L, Bunting S, Peale FV Jr, Ferrara N, et al. Vascular endothelial growth factor stimulates bone repair by promoting angiogenesis and bone turnover. Proc Natl Acad Sci U S A. 2002;99(15):9656–61.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  99. Sivaraj KK, Adams RH. Blood vessel formation and function in bone. Development. 2016;143(15):2706–15.

    Article  PubMed  CAS  Google Scholar 

  100. Schott NG, Friend NE, Stegemann JP. Coupling osteogenesis and vasculogenesis in engineered orthopedic tissues. Tissue Eng Part B Rev. 2021;27(3):199–214.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  101. Garland DE, Toder L. Fractures of the tibial diaphysis in adults with head injuries. Clin Orthop Relat Res. 1980;150:198–202.

    Article  Google Scholar 

  102. Wildburger R, Zarkovic N, Egger G, Petek W, Zarkovic K, Hofer H. Basic fibroblast growth factor (BFGF) immunoreactivity as a possible link between head injury and impaired bone fracture healing. Bone Miner. 1994;27(3):183–92.

    Article  PubMed  CAS  Google Scholar 

  103. Newman RJ, Stone MH, Mukherjee SK. Accelerated fracture union in association with severe head injury. Injury. 1987;18(4):241–6.

    Article  PubMed  CAS  Google Scholar 

  104. Khare G, Gautam V, Gupta L, Gupta A. A new hypothesis for faster healing of fractures in head injured patients. Indian J Med Sci. 1995;49(12):281–4.

    PubMed  CAS  Google Scholar 

  105. Wildburger R, Zarkovic N, Tonkovic G, Škoric T, Frech S, Hartleb M, et al. Post-traumatic hormonal disturbances: prolactin as a link between head injury and enhanced osteogenesis. J Endocrinol Investig. 1998;21(2):78–86.

    Article  CAS  Google Scholar 

  106. Li J, Ahmad T, Spetea M, Ahmed M, Kreicbergs A. Bone reinnervation after fracture: a study in the rat. J Bone Miner Res. 2001;16(8):1505–10.

    Article  PubMed  CAS  Google Scholar 

  107. Marrella A, Lee TY, Lee DH, Karuthedom S, Syla D, Chawla A, et al. Engineering vascularized and innervated bone biomaterials for improved skeletal tissue regeneration. Mater Today (Kidlington). 2018;21(4):362–76.

    Article  PubMed  CAS  Google Scholar 

  108. Lu YZ, Nayer B, Singh SK, Alshoubaki YK, Yuan E, Park AJ, et al. CGRP sensory neurons promote tissue healing via neutrophils and macrophages. Nature. 2024;628(8008):604–11.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  109. Xu J, Zhang Z, Zhao J, Meyers CA, Lee S, Qin Q, et al. Interaction between the nervous and skeletal systems. Front Cell Dev Biol. 2022;10:976736.

    Article  PubMed  PubMed Central  Google Scholar 

  110. Zhang Z, Wang F, Huang X, Sun H, Xu J, Qu H, et al. Engineered sensory nerve guides self-adaptive bone healing via NGF-TrkA signaling pathway. Adv Sci (Weinh). 2023;10(10):e2206155.

    Article  PubMed  Google Scholar 

  111. Ye L, Xu J, Mi J, He X, Pan Q, Zheng L, et al. Biodegradable magnesium combined with distraction osteogenesis synergistically stimulates bone tissue regeneration via CGRP-FAK-VEGF signaling axis. Biomaterials. 2021;275:120984.

    Article  Google Scholar 

  112. Tuo Y, Guo X, Zhang X, Wang Z, Zhou J, Xia L, et al. The biological effects and mechanisms of calcitonin gene-related peptide on human endothelial cell. J Recept Signal Transduct Res. 2013;33(2):114–23.

    Article  PubMed  CAS  Google Scholar 

  113. Paredes I, Himmels P, Ruiz de Almodóvar C. Neurovascular communication during CNS development. Dev Cell. 2018;45(1):10–32.

    Article  PubMed  CAS  Google Scholar 

  114. Li Q, Liu W, Hou W, Wu X, Wei W, Liu J, et al. Micropatterned photothermal double-layer periosteum with angiogenesis-neurogenesis coupling effect for bone regeneration. Mater Today Bio. 2022;18:100536.

    Article  PubMed  PubMed Central  Google Scholar 

  115. Wang T, Li W, Zhang Y, Xu X, Qiang L, Miao W, et al. Bioprinted constructs that simulate nerve-bone crosstalk to improve microenvironment for bone repair. Bioact Mater. 2023;27:377–93.

    PubMed  PubMed Central  CAS  Google Scholar 

  116. Deen M, Correnti E, Kamm K, Kelderman T, Papetti L, Rubio-Beltrán E, et al. Blocking CGRP in migraine patients - a review of pros and cons. J Headache Pain. 2017;18(1):96.

    Article  PubMed  PubMed Central  Google Scholar 

  117. Aubdool AA, Thakore P, Argunhan F, Smillie SJ, Schnelle M, Srivastava S, et al. A novel α-calcitonin gene-related peptide analogue protects against end-organ damage in experimental hypertension, cardiac hypertrophy, and heart failure. Circulation. 2017;136(4):367–83.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  118. Nilsson C, Hansen TK, Rosenquist C, Hartmann B, Kodra JT, Lau JF, et al. Long acting analogue of the calcitonin gene-related peptide induces positive metabolic effects and secretion of the glucagon-like peptide-1. Eur J Pharmacol. 2016;773:24–31.

    Article  PubMed  CAS  Google Scholar 

  119. Kelly RR, Sidles SJ, LaRue AC. Effects of neurological disorders on bone health. Front Psychol. 2020;11:612366.

    Article  PubMed  PubMed Central  Google Scholar 

  120. Berger JM, Singh P, Khrimian L, Morgan DA, Chowdhury S, Arteaga-Solis E, et al. Mediation of the acute stress response by the skeleton. Cell Metab. 2019;30(5):890-902.e8.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  121. Zhou R, Deng J, Zhang M, Zhou HD, Wang YJ. Association between bone mineral density and the risk of Alzheimer’s disease. J Alzheimers Dis. 2011;24(1):101–8.

    Article  PubMed  Google Scholar 

  122. Kang HG, Park HY, Ryu HU, Suk SH. Bone mineral loss and cognitive impairment: the PRESENT project. Medicine (Baltimore). 2018;97(41):e12755.

    Article  PubMed  Google Scholar 

  123. Li Y, Xiao Y, Liu C. The horizon of materiobiology: a perspective on material-guided cell behaviors and tissue engineering. Chem Rev. 2017;117(5):4376–421.

    Article  PubMed  CAS  Google Scholar 

  124. Foolen J, van Donkelaar CC, Soekhradj-Soechit S, Ito K. European society of biomechanics S.M. Perren award 2010: an adaptation mechanism for fibrous tissue to sustained shortening. J Biomech. 2010;43(16):3168–76.

    Article  PubMed  Google Scholar 

  125. Liu Z, Zhang J, Fu C, Ding J. Osteoimmunity-regulating biomaterials promote bone regeneration. Asian J Pharm Sci. 2023;18(1):100774.

    PubMed  PubMed Central  Google Scholar 

  126. Dos Santos BP, Garbay B, Fenelon M, Rosselin M, Garanger E, Lecommandoux S, et al. Development of a cell-free and growth factor-free hydrogel capable of inducing angiogenesis and innervation after subcutaneous implantation. Acta Biomater. 2019;99:154–67.

    Article  PubMed  Google Scholar 

  127. Xu Y, Xu C, Yang K, Ma L, Li G, Shi Y, et al. Copper ion-modified germanium phosphorus nanosheets integrated with an electroactive and biodegradable hydrogel for neuro-vascularized bone regeneration. Adv Healthc Mater. 2023;12(27):e2301151.

    Article  PubMed  Google Scholar 

  128. Li Q, Zhang H, Zeng Z, Yan S, Hei Y, Zhang Y, et al. Functionalized hydrogel-microsphere composites stimulating neurite outgrowth for vascularized bone regeneration. Biomater Sci. 2023;11(15):5274–86.

    Article  PubMed  CAS  Google Scholar 

  129. Barik D, Shyamal S, Das K, Jena S, Dash M. Glycoprotein injectable hydrogels promote accelerated bone regeneration through angiogenesis and innervation. Adv Healthc Mater. 2023;12(32):e2301959.

    Article  PubMed  Google Scholar 

  130. Wang H, Hsu YC, Wang C, Xiao X, Yuan Z, Zhu Y, et al. Conductive and enhanced mechanical strength of Mo2Ti2C3 MXene-based hydrogel promotes neurogenesis and bone regeneration in bone defect repair. ACS Appl Mater Interfaces. 2024;16(14):17208–18.

    Article  PubMed  CAS  Google Scholar 

  131. Li Y, Hoffman MD, Benoit DSW. Matrix metalloproteinase (MMP)-degradable tissue engineered periosteum coordinates allograft healing via early stage recruitment and support of host neurovasculature. Biomaterials. 2021;268:120535.

    Article  PubMed  CAS  Google Scholar 

  132. Zhang X, Zhang H, Zhang Y, Huangfu H, Yang Y, Qin Q, et al. 3D printed reduced graphene oxide-GelMA hybrid hydrogel scaffolds for potential neuralized bone regeneration. J Mater Chem B. 2023;11(6):1288–301.

    Article  PubMed  CAS  Google Scholar 

  133. Hao Z, Ren L, Zhang Z, Yang Z, Wu S, Liu G, et al. A multifunctional neuromodulation platform utilizing Schwann cell-derived exosomes orchestrates bone microenvironment via immunomodulation, angiogenesis and osteogenesis. Bioact Mater. 2022;23:206–22.

    PubMed  PubMed Central  Google Scholar 

  134. Xu Y, Xu C, He L, Zhou J, Chen T, Ouyang L, et al. Stratified-structural hydrogel incorporated with magnesium-ion-modified black phosphorus nanosheets for promoting neuro-vascularized bone regeneration. Bioact Mater. 2022;16:271–84.

    PubMed  PubMed Central  CAS  Google Scholar 

  135. Gu K, Tan Y, Li S, Chen S, Lin K, Tang Y, et al. Sensory nerve regulation via H3K27 demethylation revealed in akermanite composite microspheres repairing maxillofacial bone defect. Adv Sci (Weinh). 2024;11(30):e2400242.

    Article  PubMed  Google Scholar 

  136. Wan QQ, Jiao K, Ma YX, Gao B, Mu Z, Wang YR, et al. Smart, biomimetic periosteum created from the cerium(III, IV) oxide-mineralized eggshell membrane. ACS Appl Mater Interfaces. 2022;14(12):14103–19.

    Article  PubMed  CAS  Google Scholar 

  137. Su Y, Gao Q, Deng R, Zeng L, Guo J, Ye B, et al. Aptamer engineering exosomes loaded on biomimetic periosteum to promote angiogenesis and bone regeneration by targeting injured nerves via JNK3 MAPK pathway. Mater Today Bio. 2022;16:100434.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  138. Qiao F, Zou Y, Bie B, Lv Y. Dual siRNA-loaded cell membrane functionalized matrix facilitates bone regeneration with angiogenesis and neurogenesis. Small. 2024;20(8):e2307062.

    Article  PubMed  Google Scholar 

  139. Fitzpatrick V, Martín-Moldes Z, Deck A, Torres-Sanchez R, Valat A, Cairns D, et al. Functionalized 3D-printed silk-hydroxyapatite scaffolds for enhanced bone regeneration with innervation and vascularization. Biomaterials. 2021;276:120995.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  140. Wang L, Pang Y, Tang Y, Wang X, Zhang D, Zhang X, et al. A biomimetic piezoelectric scaffold with sustained Mg2+ release promotes neurogenic and angiogenic differentiation for enhanced bone regeneration. Bioact Mater. 2022;25:399–414.

    PubMed  PubMed Central  Google Scholar 

  141. Xia YJX, Wu X, Zhuang P, Guo X, Dai H. 3D-printed dual-ion chronological release functional platform reconstructs neuro-vascularization network for critical-sized bone defect regeneration. Chem Eng J. 2023;465:143015.

    Article  CAS  Google Scholar 

  142. Lian M, Qiao Z, Qiao S, Zhang X, Lin J, Xu R, et al. Nerve growth factor-preconditioned mesenchymal stem cell-derived exosome-functionalized 3D-printed hierarchical porous scaffolds with neuro-promotive properties for enhancing innervated bone regeneration. ACS Nano. 2024;18(10):7504–20.

    Article  PubMed  CAS  Google Scholar 

  143. Wang X, Ma Y, Chen J, Liu Y, Liu G, Wang P, et al. A novel decellularized matrix of Wnt signaling-activated osteocytes accelerates the repair of critical-sized parietal bone defects with osteoclastogenesis, angiogenesis, and neurogenesis. Bioact Mater. 2022;21:110–28.

    PubMed  PubMed Central  Google Scholar 

  144. Su Z, Guo C, Gui X, Wu L, Zhang B, Qin Y, et al. 3D printing of customized bioceramics for promoting bone tissue regeneration by regulating sympathetic nerve behavior. J Mater Chem B. 2024;12(17):4217–31.

    Article  PubMed  CAS  Google Scholar 

  145. Lv Z, Ji Y, Wen G, Liang X, Zhang K, Zhang W. Structure-optimized and microenvironment-inspired nanocomposite biomaterials in bone tissue engineering. Burns Trauma. 2024;12:tkae036.

    Article  PubMed  PubMed Central  Google Scholar 

  146. Yi B, Xu Q, Liu W. An overview of substrate stiffness guided cellular response and its applications in tissue regeneration. Bioact Mater. 2021;15:82–102.

    PubMed  PubMed Central  Google Scholar 

  147. D’Angelo M, Benedetti E, Tupone MG, Catanesi M, Castelli V, Antonosante A, et al. The role of stiffness in cell reprogramming: a potential role for biomaterials in inducing tissue regeneration. Cells. 2019;8(9):1036.

    Article  PubMed  PubMed Central  Google Scholar 

  148. Li J, Jiang X, Li H, Gelinsky M, Gu Z. Tailoring materials for modulation of macrophage fate. Adv Mater. 2021;33(12):2004172.

    Article  CAS  Google Scholar 

  149. Yang CY, Huang WY, Chen LH, Liang NW, Wang HC, Lu J, et al. Neural tissue engineering: the influence of scaffold surface topography and extracellular matrix microenvironment. J Mater Chem B. 2021;9(3):567–84.

    Article  PubMed  CAS  Google Scholar 

  150. Lee SS, Du X, Kim I, Ferguson SJ. Scaffolds for bone-tissue engineering. Matter. 2022;5(9):2722–59.

    Article  CAS  Google Scholar 

  151. Wang X, Xu S, Zhou S, Xu W, Leary M, Choong P, et al. Topological design and additive manufacturing of porous metals for bone scaffolds and orthopaedic implants: a review. Biomaterials. 2016;83:127–41.

    Article  PubMed  CAS  Google Scholar 

  152. Lee JH, Park JH, Lee JH, Lee HH, Knowles JC, Kim HW. Matrix-enabled mechanobiological modulation of osteoimmunology. Matter. 2022;5(10):3194–224.

    Article  CAS  Google Scholar 

  153. Kim E, Riehl BD, Bouzid T, Yang R, Duan B, Donahue HJ, et al. YAP mechanotransduction under cyclic mechanical stretch loading for mesenchymal stem cell osteogenesis is regulated by ROCK. Front Bioeng Biotechnol. 2024;11:1306002.

    Article  PubMed  PubMed Central  Google Scholar 

  154. Huebsch N, Arany PR, Mao AS, Shvartsman D, Ali OA, Bencherif SA, et al. Harnessing traction-mediated manipulation of the cell/matrix interface to control stem-cell fate. Nat Mater. 2010;9(6):518–26.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  155. Na J, Yang Z, Shi Q, Li C, Liu Y, Song Y, et al. Extracellular matrix stiffness as an energy metabolism regulator drives osteogenic differentiation in mesenchymal stem cells. Bioact Mater. 2024;35:549–63.

    PubMed  PubMed Central  CAS  Google Scholar 

  156. Santos L, Fuhrmann G, Juenet M, Amdursky N, Horejs CM, Campagnolo P, et al. Extracellular stiffness modulates the expression of functional proteins and growth factors in endothelial cells. Adv Healthc Mater. 2015;4(14):2056–63.

    Article  PubMed  CAS  Google Scholar 

  157. Chen G, Dong C, Yang L, Lv Y. 3D scaffolds with different stiffness but the same microstructure for bone tissue engineering. ACS Appl Mater Interfaces. 2015;7(29):15790–802.

    Article  PubMed  CAS  Google Scholar 

  158. Zhang T, Lin S, Shao X, Shi S, Zhang Q, Xue C, et al. Regulating osteogenesis and adipogenesis in adipose-derived stem cells by controlling underlying substrate stiffness. J Cell Physiol. 2018;233(4):3418–28.

    Article  PubMed  CAS  Google Scholar 

  159. Lee SS, Santschi M, Ferguson SJ. A biomimetic macroporous hybrid scaffold with sustained drug delivery for enhanced bone regeneration. Biomacromol. 2021;22(6):2460–71.

    Article  CAS  Google Scholar 

  160. Rosso G, Wehner D, Schweitzer C, Mollmert S, Sock E, Guck J, et al. Matrix stiffness mechanosensing modulates the expression and distribution of transcription factors in Schwann cells. Bioeng Transl Med. 2021;7(1):e10257.

    Article  PubMed  PubMed Central  Google Scholar 

  161. Liu Z, Tong H, Li J, Wang L, Fan X, Song H, et al. Low-stiffness hydrogels promote peripheral nerve regeneration through the rapid release of exosomes. Front Bioeng Biotechnol. 2022;10:922570.

    Article  PubMed  PubMed Central  Google Scholar 

  162. Engler AJ, Sweeney HL, Discher DE, Schwarzbauer JE. Extracellular matrix elasticity directs stem cell differentiation. J Musculoskelet Neuronal Interact. 2007;7(4):335.

    PubMed  CAS  Google Scholar 

  163. Huang J, Zhao L, Chen D. Growth factor signalling in osteoarthritis. Growth Factors. 2018;36(5–6):187–95.

    Article  PubMed  CAS  Google Scholar 

  164. Liu C, Li X, Zhao Q, Xie Y, Yao X, Wang M, et al. Nanofibrous bicomponent scaffolds for the dual delivery of NGF and GDNF: controlled release of growth factors and their biological effects. J Mater Sci Mater Med. 2021;32(1):9.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  165. Cross LM, Carrow JK, Ding X, Singh KA, Gaharwar AK. Sustained and prolonged delivery of protein therapeutics from two-dimensional nanosilicates. ACS Appl Mater Interfaces. 2019;11(7):6741–50.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  166. Dobre O, Oliva MA, Ciccone G, Trujillo S, Rodrigo-Navarro A, Venters DC, et al. A hydrogel platform that incorporates laminin isoforms for efficient presentation of growth factors-neural growth and osteogenesis. Adv Funct Mater. 2021;31(21):2010225.

    Article  CAS  Google Scholar 

  167. Deng J, Van Duyn C, Cohen DJ, Schwartz Z, Boyan BD. Strategies for improving impaired osseointegration in compromised animal models. J Dent Res. 2024;103(5):467–76.

    Article  PubMed  CAS  Google Scholar 

  168. Han J, Ma Q, An Y, Wu F, Zhao Y, Wu G, et al. The current status of stimuli-responsive nanotechnologies on orthopedic titanium implant surfaces. J Nanobiotechnol. 2023;21(1):277.

    Article  Google Scholar 

  169. Bauer S, Park J, Faltenbacher J, Berger S, von der Mark K, Schmuki P. Size selective behavior of mesenchymal stem cells on ZrO2 and TiO2 nanotube arrays. Integr Biol (Camb). 2009;1(8–9):525–32.

    Article  PubMed  CAS  Google Scholar 

  170. Zhao Q, Wu J, Li Y, Xu R, Zhu X, Jiao Y, et al. Promotion of bone formation and antibacterial properties of titanium coated with porous Si/Ag-doped titanium dioxide. Front Bioeng Biotechnol. 2022;10:1001514.

    Article  PubMed  PubMed Central  Google Scholar 

  171. Luo R, Jiao Y, Zhang S, Wu J, Wu X, Lu K, et al. Fabrication, properties and biological activity of a titanium surface modified with zinc via plasma electrolytic oxidation. Front Mater. 2023;10:1202110.

    Article  Google Scholar 

  172. Wu X, Jiao Y, Wu J, Zhang S, Xu R, Zhao Q, et al. Preparation, characterization, and bioactivities of cobalt, strontium and fluorine co-doped oxide films on titanium surface for clinical application. J Biomed Nanotechnol. 2024;20(4):678–86.

    Article  CAS  Google Scholar 

  173. Jiao Y, Liu Q, Chen J. Construction of N-halamine biocompatible multilayers onto BMP2 loaded titanium nanotubes for bacterial infection inhibition and osteogenic effect improvement. Mater Lett. 2020;267:127526.

    Article  CAS  Google Scholar 

  174. Deng J, Cohen DJ, Berger MB, Sabalewski EL, McClure MJ, Boyan BD, et al. Osseointegration of titanium implants in a botox-induced muscle paralysis rat model is sensitive to surface topography and semaphorin 3A treatment. Biomimetics (Basel). 2023;8(1):93.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  175. Deng J, Cohen DJ, Redden J, McClure MJ, Boyan BD, Schwartz Z. Differential effects of neurectomy and botox-induced muscle paralysis on bone phenotype and titanium implant osseointegration. Bone. 2021;153:116145.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  176. Deng J, Cohen DJ, Sabalewski EL, Van Duyn C, Wilson DS, Schwartz Z, et al. Semaphorin 3A delivered by a rapidly polymerizing click hydrogel overcomes impaired implant osseointegration in a rat type 2 diabetes model. Acta Biomater. 2023;157:236–51.

    Article  PubMed  CAS  Google Scholar 

  177. Tao B, Lan H, Zhou X, Lin C, Qin X, Wu M, et al. Regulation of TiO2 nanotubes on titanium implants to orchestrate osteo/angio-genesis and osteo-immunomodulation for boosted osseointegration. Mater Des. 2023;233:112268.

    Article  CAS  Google Scholar 

  178. Chung TW, Liu DZ, Wang SY, Wang SS. Enhancement of the growth of human endothelial cells by surface roughness at nanometer scale. Biomaterials. 2003;24(25):4655–61.

    Article  PubMed  CAS  Google Scholar 

  179. Miller DC, Thapa A, Haberstroh KM, Webster TJ. Endothelial and vascular smooth muscle cell function on poly(lactic-co-glycolic acid) with nano-structured surface features. Biomaterials. 2004;25(1):53–61.

    Article  PubMed  CAS  Google Scholar 

  180. Hochberg MC, Carrino JA, Schnitzer TJ, Guermazi A, Walsh DA, White A, et al. Long-term safety and efficacy of subcutaneous tanezumab versus nonsteroidal antiinflammatory drugs for hip or knee osteoarthritis: a randomized trial. Arthritis Rheumatol. 2021;73(7):1167–77.

    Article  PubMed  CAS  Google Scholar 

  181. Zhang H, Zhang M, Zhai D, Qin C, Wang Y, Ma J, et al. Polyhedron-like biomaterials for innervated and vascularized bone regeneration. Adv Mater. 2023;35(42):e2302716.

    Article  PubMed  Google Scholar 

  182. Karageorgiou V, Kaplan D. Porosity of 3D biomaterial scaffolds and osteogenesis. Biomaterials. 2005;26(27):5474–91.

    Article  PubMed  CAS  Google Scholar 

  183. Boire TC, Himmel LE, Yu F, Guth CM, Dollinger BR, Werfel TA, et al. Effect of pore size and spacing on neovascularization of a biodegradble shape memory polymer perivascular wrap. J Biomed Mater Res A. 2021;109(3):272–88.

    Article  PubMed  CAS  Google Scholar 

  184. Swanson WB, Omi M, Zhang Z, Nam HK, Jung Y, Wang G, et al. Macropore design of tissue engineering scaffolds regulates mesenchymal stem cell differentiation fate. Biomaterials. 2021;272:120769.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  185. Wang C, Xu D, Lin L, Li S, Hou W, He Y, et al. Large-pore-size Ti6Al4V scaffolds with different pore structures for vascularized bone regeneration. Mater Sci Eng C Mater Biol Appl. 2021;131:112499.

    Article  PubMed  CAS  Google Scholar 

  186. Liu H, Chen H, Han Q, Sun B, Liu Y, Zhang A, et al. Recent advancement in vascularized tissue-engineered bone based on materials design and modification. Mater Today Bio. 2023;23:100858.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  187. Pecci R, Baiguera S, Ioppolo P, Bedini R, Del Gaudio C. 3D printed scaffolds with random microarchitecture for bone tissue engineering applications: manufacturing and characterization. J Mech Behav Biomed Mater. 2020;103:103583.

    Article  PubMed  CAS  Google Scholar 

  188. Chen X, Fan H, Deng X, Wu L, Yi T, Gu L, et al. Scaffold structural microenvironmental cues to guide tissue regeneration in bone tissue applications. Nanomaterials (Basel). 2018;8(11):960.

    Article  PubMed  Google Scholar 

  189. Ye X, Leeflang S, Wu C, Chang J, Zhou J, Huan Z. Mesoporous bioactive glass functionalized 3D Ti-6Al-4V scaffolds with improved surface bioactivity. Materials (Basel). 2017;10(11):1244.

    Article  PubMed  Google Scholar 

  190. Sun X, Kang Y, Bao J, Zhang Y, Yang Y, Zhou X. Modeling vascularized bone regeneration within a porous biodegradable CaP scaffold loaded with growth factors. Biomaterials. 2013;34(21):4971–81.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  191. Boccaccini AR, Kneser U, Arkudas A. Scaffolds for vascularized bone regeneration: advances and challenges. Expert Rev Med Devices. 2012;9(5):457–60.

    Article  PubMed  CAS  Google Scholar 

  192. Lee JH, Parthiban P, Jin GZ, Knowles JC, Kim HW. Materials roles for promoting angiogenesis in tissue regeneration. Prog Mater Sci. 2021;117:100732.

    Article  CAS  Google Scholar 

  193. Xiao X, Wang W, Liu D, Zhang H, Gao P, Geng L, et al. The promotion of angiogenesis induced by three-dimensional porous beta-tricalcium phosphate scaffold with different interconnection sizes via activation of PI3K/Akt pathways. Sci Rep. 2015;5:9409.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  194. Kokai LE, Lin YC, Oyster NM, Marra KG. Diffusion of soluble factors through degradable polymer nerve guides: controlling manufacturing parameters. Acta Biomater. 2009;5(7):2540–50.

    Article  PubMed  CAS  Google Scholar 

  195. Rustom LE, Boudou T, Lou S, Pignot-Paintrand I, Nemke BW, Lu Y, et al. Micropore-induced capillarity enhances bone distribution in vivo in biphasic calcium phosphate scaffolds. Acta Biomater. 2016;44:144–54.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  196. Wang X, Lou T, Zhao W, Song G, Li C, Cui G. The effect of fiber size and pore size on cell proliferation and infiltration in PLLA scaffolds on bone tissue engineering. J Biomater Appl. 2016;30(10):1545–51.

    Article  PubMed  CAS  Google Scholar 

  197. Lu H, Zhang N, Ma M. Electroconductive hydrogels for biomedical applications. Wiley Interdiscip Rev Nanomed Nanobiotechnol. 2019;11(6):e1568.

    Article  PubMed  Google Scholar 

  198. Qin W, Li L, Niu W, Wang WR, Wu DW, Song CG, et al. Effects of electric field-modulated conductive hydrogel on osseoperception and osseointegration of dental implants. Adv Funct Mater. 2024;34:2400256.

    Article  CAS  Google Scholar 

  199. Wang Q, Wang H, Ma Y, Cao X, Gao H. Effects of electroactive materials on nerve cell behaviors and applications in peripheral nerve repair. Biomater Sci. 2022;10(21):6061–76.

    Article  PubMed  CAS  Google Scholar 

  200. DeVet T, Jhirad A, Pravato L, Wohl GR. Bone bioelectricity and bone-cell response to electrical stimulation: a review. Crit Rev Biomed Eng. 2021;49(1):1–19.

    Article  PubMed  Google Scholar 

  201. Shim G, Breinyn IB, Martínez-Calvo A, Rao S, Cohen DJ. Bioelectric stimulation controls tissue shape and size. Nat Commun. 2024;15(1):2938.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  202. Ferrigno B, Bordett R, Duraisamy N, Moskow J, Arul MR, Rudraiah S, et al. Bioactive polymeric materials and electrical stimulation strategies for musculoskeletal tissue repair and regeneration. Bioact Mater. 2020;5(3):468–85.

    PubMed  PubMed Central  Google Scholar 

  203. Liu S, Yu JM, Gan YC, Qiu XZ, Gao ZC, Wang H, et al. Biomimetic natural biomaterials for tissue engineering and regenerative medicine: new biosynthesis methods, recent advances, and emerging applications. Mil Med Res. 2023;10(1):16.

    PubMed  PubMed Central  CAS  Google Scholar 

  204. Su Y, Zeng L, Deng R, Ye B, Tang S, Xiong Z, et al. Endogenous electric field-coupled PD@BP biomimetic periosteum promotes bone regeneration through sensory nerve via fanconi anemia signaling pathway. Adv Healthc Mater. 2023;12(12):e2203027.

    Article  PubMed  Google Scholar 

  205. Zhao Y, Wang H, Huang H, Xiao Q, Xu Y, Guo Z, et al. Surface coordination of black phosphorus for robust air and water stability. Angew Chem Int Ed Engl. 2016;55(16):5003–7.

    Article  PubMed  CAS  Google Scholar 

  206. Guo Z, Chen S, Wang Z, Yang Z, Liu F, Xu Y, et al. Metal-ion-modified black phosphorus with enhanced stability and transistor performance. Adv Mater. 2017. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/adma.201703811.

    Article  PubMed  PubMed Central  Google Scholar 

  207. Zeng G, Chen Y. Surface modification of black phosphorus-based nanomaterials in biomedical applications: strategies and recent advances. Acta Biomater. 2020;118:1–17.

    Article  PubMed  CAS  Google Scholar 

  208. Zhong Y, Huang S, Feng Z, Fu Y, Mo A. Recent advances and trends in the applications of MXene nanomaterials for tissue engineering and regeneration. J Biomed Mater Res A. 2022;110(11):1840–59.

    Article  PubMed  CAS  Google Scholar 

  209. Wang Y, Zhu L, Du C. Progress in piezoelectric nanogenerators based on PVDF composite films. Micromachines (Basel). 2021;12(11):1278.

    Article  PubMed  Google Scholar 

  210. Yang F, Li J, Long Y, Zhang Z, Wang L, Sui J, et al. Wafer-scale heterostructured piezoelectric bio-organic thin films. Science. 2021;373(6552):337–42.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  211. Liu Y, Dzidotor G, Le TT, Vinikoor T, Morgan K, Curry EJ, et al. Exercise-induced piezoelectric stimulation for cartilage regeneration in rabbits. Sci Transl Med. 2022;14(627):eabi7282.

    Article  PubMed  CAS  Google Scholar 

  212. Ahmadi N, Kharaziha M, Labbaf S. Core-shell fibrous membranes of PVDF-Ba0.9Ca0.1TiO3/PVA with osteogenic and piezoelectric properties for bone regeneration. Biomed Mater. 2019;15(1):015007.

    Article  PubMed  Google Scholar 

  213. Petrovova E, Giretova M, Kvasilova A, Benada O, Danko J, Medvecky L, et al. Preclinical alternative model for analysis of porous scaffold biocompatibility in bone tissue engineering. Altex. 2019;36(1):121–30.

    Article  PubMed  Google Scholar 

  214. Zheng T, Yu Y, Pang Y, Zhang D, Wang Y, Zhao H, et al. Improving bone regeneration with composites consisting of piezoelectric poly(L-lactide) and piezoelectric calcium/manganese co-doped barium titanate nanofibers. Compos B Eng. 2022;234:109734.

    Article  CAS  Google Scholar 

  215. Cai K, Jiao Y, Quan Q, Hao Y, Liu J, Wu L. Improved activity of MC3T3-E1 cells by the exciting piezoelectric BaTiO3/TC4 using low-intensity pulsed ultrasound. Bioact Mater. 2021;6(11):4073–82.

    PubMed  PubMed Central  CAS  Google Scholar 

  216. Timin AS, Muslimov AR, Zyuzin MV, Peltek OO, Karpov TE, Sergeev IS, et al. Multifunctional scaffolds with improved antimicrobial properties and osteogenicity based on piezoelectric electrospun fibers decorated with bioactive composite microcapsules. ACS Appl Mater Interfaces. 2018;10(41):34849–68.

    Article  PubMed  CAS  Google Scholar 

  217. Zheng T, Zhao H, Huang Y, Gao C, Zhang X, Cai Q, et al. Piezoelectric calcium/manganese-doped barium titanate nanofibers with improved osteogenic activity. Ceram Int. 2021;47:28778–89.

    Article  CAS  Google Scholar 

  218. Zhou P, Liu T, Liu W, Sun L, Kang H, Liu K, et al. An antibacterial bionic periosteum with angiogenesis-neurogenesis coupling effect for bone regeneration. ACS Appl Mater Interfaces. 2024. https://doiorg.publicaciones.saludcastillayleon.es/10.1021/acsami.4c01206.

    Article  PubMed  PubMed Central  Google Scholar 

  219. Jang HL, Jin K, Lee J, Kim Y, Nahm SH, Hong KS, et al. Revisiting whitlockite, the second most abundant biomineral in bone: nanocrystal synthesis in physiologically relevant conditions and biocompatibility evaluation. ACS Nano. 2014;8(1):634–41.

    Article  PubMed  CAS  Google Scholar 

  220. Zhao Y, Cai Y, Wang W, Bai Y, Liu M, Wang Y, et al. Periosteum-bone inspired hierarchical scaffold with endogenous piezoelectricity for neuro-vascularized bone regeneration. Bioact Mater. 2024;44:339–53.

    PubMed  PubMed Central  Google Scholar 

  221. Cidonio G, Glinka M, Kim YH, Kanczler JM, Lanham SA, Ahlfeld T, et al. Nanoclay-based 3D printed scaffolds promote vascular ingrowth ex vivo and generate bone mineral tissue in vitro and in vivo. Biofabrication. 2020;12(3):035010.

    Article  PubMed  CAS  Google Scholar 

  222. Li W, Miao W, Liu Y, Wang T, Zhang Y, Wang W, et al. Bioprinted constructs that mimic the ossification center microenvironment for targeted innervation in bone regeneration. Adv Funct Mater. 2022;32(9):2109871.

    Article  CAS  Google Scholar 

  223. Shen M, Wang L, Gao Y, Feng L, Xu C, Li S, et al. 3D bioprinting of in situ vascularized tissue engineered bone for repairing large segmental bone defects. Mater Today Bio. 2022;16:100382.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  224. Li L, Li J, Zou Q, Zuo Y, Cai B, Li Y. Enhanced bone tissue regeneration of a biomimetic cellular scaffold with co-cultured MSCs-derived osteogenic and angiogenic cells. Cell Prolif. 2019;52(5):e12658.

    Article  PubMed  PubMed Central  Google Scholar 

  225. Zhang J, Suttapreyasri S, Leethanakul C, Samruajbenjakun B. Fabrication of vascularized tissue-engineered bone models using triaxial bioprinting. J Biomed Mater Res A. 2024;112(7):1093–106.

    Article  PubMed  CAS  Google Scholar 

  226. Zhang H, Qin C, Zhang M, Han Y, Ma J, Wu J, et al. Calcium silicate nanowires-containing multicellular bioinks for 3D bioprinting of neural-bone constructs. Nano Today. 2022;46:101584.

    Article  CAS  Google Scholar 

  227. Qin C, Zhang H, Chen L, Zhang M, Ma J, Zhuang H, et al. Cell-laden scaffolds for vascular-innervated bone regeneration. Adv Healthc Mater. 2023;12(13):e2201923.

    Article  PubMed  Google Scholar 

  228. Zhang M, Lin R, Wang X, Xue J, Deng C, Feng C, et al. 3D printing of Haversian bone-mimicking scaffolds for multicellular delivery in bone regeneration. Sci Adv. 2020;6(12):eaaz6725.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  229. Aloe L. Rita Levi-Montalcini: the discovery of nerve growth factor and modern neurobiology. Trends Cell Biol. 2004;14(7):395–9.

    Article  PubMed  CAS  Google Scholar 

  230. Liu Q, Lei L, Yu T, Jiang T, Kang Y. Effect of brain-derived neurotrophic factor on the neurogenesis and osteogenesis in bone engineering. Tissue Eng Part A. 2018;24(15–16):1283–92.

    Article  PubMed  CAS  Google Scholar 

  231. Zhao L, Lai Y, Jiao H, Huang J. Nerve growth factor receptor limits inflammation to promote remodeling and repair of osteoarthritic joints. Nat Commun. 2024;15(1):3225.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  232. Takeda K, Shiba H, Mizuno N, Hasegawa N, Mouri Y, Hirachi A, et al. Brain-derived neurotrophic factor enhances periodontal tissue regeneration. Tissue Eng. 2005;11(9–10):1618–29.

    Article  PubMed  CAS  Google Scholar 

  233. Shamloo A, Heibatollahi M, Mofrad MRK. Directional migration and differentiation of neural stem cells within three-dimensional microenvironments. Integr Biol (Camb). 2015;7(3):335–44.

    Article  PubMed  CAS  Google Scholar 

  234. Zhang F, Liu CL, Tong MM, Zhao Z, Chen SQ. Both Wnt/β-catenin and ERK5 signaling pathways are involved in BDNF-induced differentiation of pluripotent stem cells into neural stem cells. Neurosci Lett. 2019;708:134345.

    Article  PubMed  CAS  Google Scholar 

  235. Shu XQ, Mendell LM. Neurotrophins and hyperalgesia. Proc Natl Acad Sci U S A. 1999;96(14):7693–6.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  236. Abeynayake N, Arthur A, Gronthos S. Crosstalk between skeletal and neural tissues is critical for skeletal health. Bone. 2021;142:115645.

    Article  PubMed  CAS  Google Scholar 

  237. Kaplan DR, Hempstead BL, Martin-Zanca D, Chao MV, Parada LF. The trk proto-oncogene product: a signal transducing receptor for nerve growth factor. Science. 1991;252(5005):554–8.

    Article  PubMed  CAS  Google Scholar 

  238. Verdi JM, Birren SJ, Ibáñez CF, Persson H, Kaplan DR, Benedetti M, et al. p75LNGFR regulates Trk signal transduction and NGF-induced neuronal differentiation in MAH cells. Neuron. 1994;12(4):733–45.

    Article  PubMed  CAS  Google Scholar 

  239. Lazarovici P, Lahiani A, Gincberg G, Haham D, Fluksman A, Benny O, et al. Nerve growth factor-induced angiogenesis: 1. Endothelial cell tube formation assay. Methods Mol Biol. 2018;1727:239–50.

    Article  PubMed  CAS  Google Scholar 

  240. Li X, Li F, Ling L, Li C, Zhong Y. Intranasal administration of nerve growth factor promotes angiogenesis via activation of PI3K/Akt signaling following cerebral infarction in rats. Am J Transl Res. 2018;10(11):3481.

    PubMed  PubMed Central  CAS  Google Scholar 

  241. Troullinaki M, Alexaki VI, Mitroulis I, Witt A, Klotzsche-von Ameln A, Chung KJ, et al. Nerve growth factor regulates endothelial cell survival and pathological retinal angiogenesis. J Cell Mol Med. 2019;23(4):2362–71.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  242. Gaharwar AK, Cross LM, Peak CW, Gold K, Carrow JK, Brokesh A, et al. 2D nanoclay for biomedical applications: regenerative medicine, therapeutic delivery, and additive manufacturing. Adv Mater. 2019;31(23):e1900332.

    Article  PubMed  PubMed Central  Google Scholar 

  243. Wise BL, Seidel MF, Lane NE. The evolution of nerve growth factor inhibition in clinical medicine. Nat Rev Rheumatol. 2021;17(1):34–46.

    Article  PubMed  Google Scholar 

  244. Kilian O, Hartmann S, Dongowski N, Karnati S, Baumgart-Vogt E, Hartel FV, et al. BDNF and its TrkB receptor in human fracture healing. Ann Anat. 2014;196(5):286–95.

    Article  PubMed  Google Scholar 

  245. Kanchanawong P, Calderwood DA. Organization, dynamics and mechanoregulation of integrin-mediated cell-ECM adhesions. Nat Rev Mol Cell Biol. 2023;24(2):142–61.

    Article  PubMed  CAS  Google Scholar 

  246. Kanie K, Narita Y, Zhao Y, Kuwabara F, Satake M, Honda S, et al. Collagen type IV-specific tripeptides for selective adhesion of endothelial and smooth muscle cells. Biotechnol Bioeng. 2012;109(7):1808–16.

    Article  PubMed  CAS  Google Scholar 

  247. Wei Y, Ji Y, Xiao L, Lin Q, Ji J. Different complex surfaces of polyethyleneglycol (PEG) and REDV ligand to enhance the endothelial cells selectivity over smooth muscle cells. Colloids Surf B Biointerfaces. 2011;84(2):369–78.

    Article  PubMed  CAS  Google Scholar 

  248. Hamada Y, Nokihara K, Okazaki M, Fujitani W, Matsumoto T, Matsuo M, et al. Angiogenic activity of osteopontin-derived peptide SVVYGLR. Biochem Biophys Res Commun. 2003;310(1):153–7.

    Article  PubMed  CAS  Google Scholar 

  249. Hubbell JA, Massia SP, Desai NP, Drumheller PD. Endothelial cell-selective materials for tissue engineering in the vascular graft via a new receptor. Biotechnology (N Y). 1991;9(6):568–72.

    PubMed  CAS  Google Scholar 

  250. Plouffe BD, Radisic M, Murthy SK. Microfluidic depletion of endothelial cells, smooth muscle cells, and fibroblasts from heterogeneous suspensions. Lab Chip. 2008;8(3):462–72.

    Article  PubMed  CAS  Google Scholar 

  251. Alagoz AS, Rodriguez-Cabello JC, Hasirci V. PHBV wet-spun scaffold coated with ELR-REDV improves vascularization for bone tissue engineering. Biomed Mater. 2018;13(5):055010.

    Article  PubMed  Google Scholar 

  252. Liesi P, Närvänen A, Soos J, Sariola H, Snounou G. Identification of a neurite outgrowth-promoting domain of laminin using synthetic peptides. FEBS Lett. 1989;244(1):141–8.

    Article  PubMed  CAS  Google Scholar 

  253. Tashiro K, Sephel GC, Weeks B, Sasaki M, Martin GR, Kleinman HK, et al. A synthetic peptide containing the IKVAV sequence from the A chain of laminin mediates cell attachment, migration, and neurite outgrowth. J Biol Chem. 1989;264(27):16174–82.

    Article  PubMed  CAS  Google Scholar 

  254. Ren T, Yu S, Mao Z, Moya SE, Han L, Gao C. Complementary density gradient of poly(hydroxyethyl methacrylate) and YIGSR selectively guides migration of endotheliocytes. Biomacromol. 2014;15(6):2256–64.

    Article  CAS  Google Scholar 

  255. Nakamura M, Mie M, Mihara H, Nakamura M, Kobatake E. Construction of multi-functional extracellular matrix proteins that promote tube formation of endothelial cells. Biomaterials. 2008;29(20):2977–86.

    Article  PubMed  CAS  Google Scholar 

  256. Mahmoudi N, Roque M, Paiva Dos Santos B, Oliveira H, Siadous R, Rey S, et al. An elastin-derived composite matrix for enhanced vascularized and innervated bone tissue reconstruction: from material development to preclinical evaluation. Adv Healthc Mater. 2024;13(18):e2303765.

    Article  PubMed  Google Scholar 

  257. Aziz AH, Bryant SJ. A comparison of human mesenchymal stem cell osteogenesis in poly(ethylene glycol) hydrogels as a function of MMP-sensitive crosslinker and crosslink density in chemically defined medium. Biotechnol Bioeng. 2019;116(6):1523–36.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  258. Bracher M, Bezuidenhout D, Lutolf MP, Franz T, Sun M, Zilla P, et al. Cell specific ingrowth hydrogels. Biomaterials. 2013;34(28):6797–803.

    Article  PubMed  CAS  Google Scholar 

  259. Davis GE, Stratman AN, Sacharidou A, Koh W. Molecular basis for endothelial lumen formation and tubulogenesis during vasculogenesis and angiogenic sprouting. Int Rev Cell Mol Biol. 2011;288:101–65.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  260. Chan ZC, Oentaryo MJ, Lee CW. MMP-mediated modulation of ECM environment during axonal growth and NMJ development. Neurosci Lett. 2020;724:134822.

    Article  PubMed  CAS  Google Scholar 

  261. Wang L, Zhang ZG, Zhang RL, Gregg SR, Hozeska-Solgot A, LeTourneau Y, et al. Matrix metalloproteinase 2 (MMP2) and MMP9 secreted by erythropoietin-activated endothelial cells promote neural progenitor cell migration. J Neurosci. 2006;26(22):5996–6003.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  262. Chen W, Zhou Z, Chen D, Li Y, Zhang Q, Su J. Bone regeneration using MMP-cleavable peptides-based hydrogels. Gels. 2021;7(4):199.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  263. Gómez-Aguado I, Rodríguez-Castejón J, Vicente-Pascual M, Rodríguez-Gascón A, Solinís MÁ, Del Pozo-Rodríguez A. Nanomedicines to deliver mRNA: state of the art and future perspectives. Nanomaterials (Basel). 2020;10(2):364.

    Article  PubMed  Google Scholar 

  264. DeJulius CR, Walton BL, Colazo JM, d’Arcy R, Francini N, Brunger JM, et al. Engineering approaches for RNA-based and cell-based osteoarthritis therapies. Nat Rev Rheumatol. 2024;20(2):81–100.

    Article  PubMed  PubMed Central  Google Scholar 

  265. Raftery RM, Walsh DP, Castaño IM, Heise A, Duffy GP, Cryan SA, et al. Delivering nucleic-acid based nanomedicines on biomaterial scaffolds for orthopedic tissue repair: challenges, progress and future perspectives. Adv Mater. 2016;28(27):5447–69.

    Article  PubMed  CAS  Google Scholar 

  266. Tsekoura EK, KC RB, Uludag H. Biomaterials to facilitate delivery of RNA agents in bone regeneration and repair. ACS Biomater Sci Eng. 2017;3(7):1195–206.

    Article  PubMed  CAS  Google Scholar 

  267. Sharawy I. Neuroimmune crosstalk and its impact on cancer therapy and research. Discov Oncol. 2022;13(1):80.

    Article  PubMed  PubMed Central  Google Scholar 

  268. Zhang D, Wang Y, Zhou Z, Wang L, Liu C, Jiang Y. Role of miRNA-regulated type H vessel formation in osteoporosis. Front Endocrinol (Lausanne). 2024;15:1394785.

    Article  PubMed  Google Scholar 

  269. Xia M. Great potential of microRNA in cancer stem cell. J Cancer. 2008;4(3):79–89.

    CAS  Google Scholar 

  270. Akram R, Anwar H, Javed MS, Rasul A, Imran A, Malik SA, et al. Axonal regeneration: underlying molecular mechanisms and potential therapeutic targets. Biomedicines. 2022;10(12):3186.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  271. Silvestro S, Mazzon E. MiRNAs as promising translational strategies for neuronal repair and regeneration in spinal cord injury. Cells. 2022;11(14):2177.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  272. Xu X, Liu R, Li Y, Zhang C, Guo C, Zhu J, et al. Spinal cord injury: from microRNAs to exosomal microRNAs. Mol Neurobiol. 2024;61(8):5974–91.

    Article  PubMed  CAS  Google Scholar 

  273. Zhang N, Lin J, Lin VPH, Milbreta U, Chin JS, Chew EGY, et al. A 3D fiber-hydrogel based non-viral gene delivery platform reveals that microRNAs promote axon regeneration and enhance functional recovery following spinal cord injury. Adv Sci (Weinh). 2021;8(15):e2100805.

    Article  PubMed  Google Scholar 

  274. Zhou S, Zhang S, Wang Y, Yi S, Zhao L, Tang X, et al. miR-21 and miR-222 inhibit apoptosis of adult dorsal root ganglion neurons by repressing TIMP3 following sciatic nerve injury. Neurosci Lett. 2015;586:43–9.

    Article  PubMed  CAS  Google Scholar 

  275. Wang J, Yang Q, Saiding Q, Chen L, Liu M, Wang Z, et al. Geometric angles and gene expression in cells for structural bone regeneration. Adv Sci (Weinh). 2023;10(32):e2304111.

    Article  PubMed  Google Scholar 

  276. Lei L, Liu Z, Yuan P, Jin R, Wang X, Jiang T, et al. Injectable colloidal hydrogel with mesoporous silica nanoparticles for sustained co-release of microRNA-222 and aspirin to achieve innervated bone regeneration in rat mandibular defects. J Mater Chem B. 2019;7(16):2722–35.

    Article  PubMed  CAS  Google Scholar 

  277. Jia Y, Zhu Y, Qiu S, Xu J, Chai Y. Exosomes secreted by endothelial progenitor cells accelerate bone regeneration during distraction osteogenesis by stimulating angiogenesis. Stem Cell Res Ther. 2019;10(1):12.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  278. Liu XD, Cai F, Liu L, Zhang Y, Yang AL. microRNA-210 is involved in the regulation of postmenopausal osteoporosis through promotion of VEGF expression and osteoblast differentiation. Biol Chem. 2015;396(4):339–47.

    Article  PubMed  CAS  Google Scholar 

  279. Yang C, Liu X, Zhao K, Zhu Y, Hu B, Zhou Y, et al. miRNA-21 promotes osteogenesis via the PTEN/PI3K/Akt/HIF-1α pathway and enhances bone regeneration in critical size defects. Stem Cell Res Ther. 2019;10(1):65.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  280. Costa V, Raimondi L, Conigliaro A, Salamanna F, Carina V, De Luca A, et al. Hypoxia-inducible factor 1A may regulate the commitment of mesenchymal stromal cells toward angio-osteogenesis by mirna-675-5P. Cytotherapy. 2017;19(12):1412–25.

    Article  PubMed  CAS  Google Scholar 

  281. Leng Q, Chen L, Lv Y. RNA-based scaffolds for bone regeneration: application and mechanisms of mRNA, miRNA and siRNA. Theranostics. 2020;10(7):3190–205.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  282. Nguyen MK, Jeon O, Dang PN, Huynh CT, Varghai D, Riazi H, et al. RNA interfering molecule delivery from in situ forming biodegradable hydrogels for enhancement of bone formation in rat calvarial bone defects. Acta Biomater. 2018;75:105–14.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  283. Whitehead KA, Langer R, Anderson DG. Knocking down barriers: advances in siRNA delivery. Nat Rev Drug Discov. 2009;8(2):129–38.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  284. Wittrup A, Lieberman J. Knocking down disease: a progress report on siRNA therapeutics. Nat Rev Genet. 2015;16(9):543–52.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  285. Qian RC, Zhou ZR, Guo W, Wu Y, Yang Z, Lu Y. Cell surface engineering using DNAzymes: metal ion mediated control of cell-cell interactions. J Am Chem Soc. 2021;143(15):5737–44.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  286. Tsvetkov P, Coy S, Petrova B, Dreishpoon M, Verma A, Abdusamad M, et al. Copper induces cell death by targeting lipoylated TCA cycle proteins. Science. 2022;375(6586):1254–61.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  287. O’Neill E, Awale G, Daneshmandi L, Umerah O, Lo KW. The roles of ions on bone regeneration. Drug Discov Today. 2018;23(4):879–90.

    Article  PubMed  CAS  Google Scholar 

  288. Sansone V, Pagani D, Melato M. The effects on bone cells of metal ions released from orthopaedic implants. A review. Clin Cases Miner Bone Metab. 2013;10(1):34–40.

    PubMed  PubMed Central  Google Scholar 

  289. Zhao Y, Li J, Liu L, Wang Y, Ju Y, Zeng C, et al. Zinc-based tannin-modified composite microparticulate scaffolds with balanced antimicrobial activity and osteogenesis for infected bone defect repair. Adv Healthc Mater. 2023;12(20):e2300303.

    Article  PubMed  Google Scholar 

  290. Jiao Y, Wang X, Chen JH. Biofabrication of AuNPs using coriandrum sativum leaf extract and their antioxidant, analgesic activity. Sci Total Environ. 2021;767:144914.

    Article  PubMed  CAS  Google Scholar 

  291. Staiger MP, Pietak AM, Huadmai J, Dias G. Magnesium and its alloys as orthopedic biomaterials: a review. Biomaterials. 2006;27(9):1728–34.

    Article  PubMed  CAS  Google Scholar 

  292. Sun L, Wang M, Chen S, Sun B, Guo Y, He C, et al. Molecularly engineered metal-based bioactive soft materials—neuroactive magnesium ion/polymer hybrids. Acta Biomater. 2019;85:310–9.

    Article  PubMed  CAS  Google Scholar 

  293. Zhang Y, Xu J, Ruan YC, Yu MK, O’Laughlin M, Wise H, et al. Implant-derived magnesium induces local neuronal production of CGRP to improve bone-fracture healing in rats. Nat Med. 2016;22(10):1160–9.

    Article  PubMed  PubMed Central  Google Scholar 

  294. Yuan Z, Wan Z, Gao C, Wang Y, Huang J, Cai Q. Controlled magnesium ion delivery system for in situ bone tissue engineering. J Control Release. 2022;350:360–76.

    Article  PubMed  CAS  Google Scholar 

  295. Tsao YT, Shih YY, Liu YA, Liu YS, Lee OK. Knockdown of SLC41A1 magnesium transporter promotes mineralization and attenuates magnesium inhibition during osteogenesis of mesenchymal stromal cells. Stem Cell Res Ther. 2017;8(1):39.

    Article  PubMed  PubMed Central  Google Scholar 

  296. Maradze D, Musson D, Zheng Y, Cornish J, Lewis M, Liu Y. High magnesium corrosion rate has an effect on osteoclast and mesenchymal stem cell role during bone remodelling. Sci Rep. 2018;8(1):10003.

    Article  PubMed  PubMed Central  Google Scholar 

  297. Kargozar S, Baino F, Hamzehlou S, Hill RG, Mozafari M. Bioactive glasses: sprouting angiogenesis in tissue engineering. Trends Biotechnol. 2018;36(4):430–44.

    Article  PubMed  CAS  Google Scholar 

  298. Hoppe A, Mouriño V, Boccaccini AR. Therapeutic inorganic ions in bioactive glasses to enhance bone formation and beyond. Biomater Sci. 2013;1(3):254–6.

    Article  PubMed  CAS  Google Scholar 

  299. Kargozar S, Baino F, Hamzehlou S, Hill RG, Mozafari M. Bioactive glasses entering the mainstream. Drug Discovery Today. 2018;23(10):1700–4.

    Article  PubMed  CAS  Google Scholar 

  300. Hoppe A, Güldal NS, Boccaccini AR. A review of the biological response to ionic dissolution products from bioactive glasses and glass-ceramics. Biomaterials. 2011;32(11):2757–74.

    Article  PubMed  CAS  Google Scholar 

  301. Zhou X, Zhang N, Mankoci S, Sahai N. Silicates in orthopedics and bone tissue engineering materials. J Biomed Mater Res A. 2017;105(7):2090–102.

    Article  PubMed  CAS  Google Scholar 

  302. Shi M, Zhou Y, Shao J, Chen Z, Song B, Chang J, et al. Stimulation of osteogenesis and angiogenesis of hBMSCs by delivering Si ions and functional drug from mesoporous silica nanospheres. Acta Biomater. 2015;21:178–89.

    Article  PubMed  CAS  Google Scholar 

  303. Ding Z, Qiao Y, Peng F, Xia C, Qian S, Wang T, et al. Si-doped porous TiO2 coatings enhanced in vitro angiogenic behavior of human umbilical vein endothelial cells. Colloids Surf B Biointerfaces. 2017;159:493–500.

    Article  PubMed  CAS  Google Scholar 

  304. Fu X, Liu P, Zhao D, Yuan B, Xiao Z, Zhou Y, et al. Effects of nanotopography regulation and silicon doping on angiogenic and osteogenic activities of hydroxyapatite coating on titanium implant. Int J Nanomed. 2020;15:4171–89.

    Article  CAS  Google Scholar 

  305. Ma YX, Jiao K, Wan QQ, Li J, Liu MY, Zhang ZB, et al. Silicified collagen scaffold induces semaphorin 3A secretion by sensory nerves to improve in-situ bone regeneration. Bioact Mater. 2021;9:475–90.

    PubMed  PubMed Central  Google Scholar 

  306. Yang B, Chen Y, Shi J. Reactive oxygen species (ROS)-based nanomedicine. Chem Rev. 2019;119(8):4881–985.

    Article  PubMed  CAS  Google Scholar 

  307. Casals E, Zeng M, Parra-Robert M, Fernandez-Varo G, Morales-Ruiz M, Jimenez W, et al. Cerium oxide nanoparticles: advances in biodistribution, toxicity, and preclinical exploration. Small. 2020;16(20):e1907322.

    Article  PubMed  Google Scholar 

  308. Ciosek Ż, Kot K, Kosik-Bogacka D, Lanocha-Arendarczyk N, Rotter I. The effects of calcium, magnesium, phosphorus, fluoride, and lead on bone tissue. Biomolecules. 2021;11(4):506.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  309. Shearer A, Molinaro M, Montazerian M, Sly JJ, Miola M, Baino F, et al. The unexplored role of alkali and alkaline earth elements (ALAEs) on the structure, processing, and biological effects of bioactive glasses. Biomater Sci. 2024;12(10):2521–60.

    Article  PubMed  CAS  Google Scholar 

  310. Südhof TC. Calcium control of neurotransmitter release. Cold Spring Harb Perspect Biol. 2012;4(1):a011353.

    Article  PubMed  PubMed Central  Google Scholar 

  311. Stähli C, Muja N, Nazhat SN. Controlled copper ion release from phosphate-based glasses improves human umbilical vein endothelial cell survival in a reduced nutrient environment. Tissue Eng Part A. 2013;19(3–4):548–57.

    Article  PubMed  Google Scholar 

  312. Rigiracciolo DC, Scarpelli A, Lappano R, Pisano A, Santolla MF, De Marco P, et al. Copper activates HIF-1α/GPER/VEGF signalling in cancer cells. Oncotarget. 2015;6(33):34158.

    Article  PubMed  PubMed Central  Google Scholar 

  313. Wu C, Zhou Y, Xu M, Han P, Chen L, Chang J, et al. Copper-containing mesoporous bioactive glass scaffolds with multifunctional properties of angiogenesis capacity, osteostimulation and antibacterial activity. Biomaterials. 2013;34(2):422–33.

    Article  PubMed  CAS  Google Scholar 

  314. Yu W, Sun TW, Ding Z, Qi C, Zhao H, Chen F, et al. Copper-doped mesoporous hydroxyapatite microspheres synthesized by a microwave-hydrothermal method using creatine phosphate as an organic phosphorus source: application in drug delivery and enhanced bone regeneration. J Mater Chem B. 2017;5(5):1039–52.

    Article  PubMed  CAS  Google Scholar 

  315. Zhang J, Wu H, He F, Wu T, Zhou L, Ye J. Concentration-dependent osteogenic and angiogenic biological performances of calcium phosphate cement modified with copper ions. Mater Sci Eng C. 2019;99:1199–212.

    Article  CAS  Google Scholar 

  316. Barralet J, Gbureck U, Habibovic P, Vorndran E, Gerard C, Doillon CJ. Angiogenesis in calcium phosphate scaffolds by inorganic copper ion release. Tissue Eng Part A. 2009;15(7):1601–9.

    Article  PubMed  CAS  Google Scholar 

  317. Lu Y, Li L, Lin Z, Wang L, Lin L, Li M, et al. A new treatment modality for rheumatoid arthritis: combined photothermal and photodynamic therapy using Cu7.2S4 nanoparticles. Adv Healthc Mater. 2018;7(14):e1800013.

    Article  PubMed  Google Scholar 

  318. Lin Z, Cao Y, Zou J, Zhu F, Gao Y, Zheng X, et al. Improved osteogenesis and angiogenesis of a novel copper ions doped calcium phosphate cement. Mater Sci Eng C Mater Biol Appl. 2020;114:111032.

    Article  PubMed  CAS  Google Scholar 

  319. Das BK, Verma SK, Das T, Panda PK, Parashar K, Suar M, et al. Altered electrical properties with controlled copper doping in ZnO nanoparticles infers their cytotoxicity in macrophages by ROS induction and apoptosis. Chem Biol Interact. 2019;297:141–54.

    Article  PubMed  CAS  Google Scholar 

  320. Si Y, Liu H, Yu H, Jiang X, Sun D. MOF-derived CuO@ZnO modified titanium implant for synergistic antibacterial ability, osteogenesis and angiogenesis. Colloids Surf B Biointerfaces. 2022;219:112840.

    Article  PubMed  CAS  Google Scholar 

  321. Angelé-Martínez C, Nguyen KVT, Ameer FS, Anker JN, Brumaghim JL. Reactive oxygen species generation by copper (II) oxide nanoparticles determined by DNA damage assays and EPR spectroscopy. Nanotoxicology. 2017;11(2):278–88.

    Article  PubMed  PubMed Central  Google Scholar 

  322. Bosch-Rué E, Díez-Tercero L, Rodríguez-González R, Bosch-Canals BM, Perez RA. Assessing the potential role of copper and cobalt in stimulating angiogenesis for tissue regeneration. PLoS ONE. 2021;16(10):e0259125.

    Article  PubMed  PubMed Central  Google Scholar 

  323. Yu Y, Li X, Ying Q, Zhang Z, Liu W, Su J. Synergistic effects of shed-derived exosomes, Cu2+, and an injectable hyaluronic acid hydrogel on antibacterial, anti-inflammatory, and osteogenic activity for periodontal bone regeneration. ACS Appl Mater Interfaces. 2024;16(26):33053–69.

    Article  PubMed  CAS  Google Scholar 

  324. Pegtel DM, Gould SJ. Exosomes. Annu Rev Biochem. 2019;88:487–514.

    Article  PubMed  CAS  Google Scholar 

  325. Zhang Y, Xie Y, Hao Z, Zhou P, Wang P, Fang S, et al. Correction to “Umbilical mesenchymal stem cell-derived exosome-encapsulated hydrogels accelerate bone repair by enhancing angiogenesis.” ACS Appl Mater Interfaces. 2022;14(12):14834–5.

    Article  PubMed  CAS  Google Scholar 

  326. Zhang L, Jiao G, Ren S, Zhang X, Li C, Wu W, et al. Exosomes from bone marrow mesenchymal stem cells enhance fracture healing through the promotion of osteogenesis and angiogenesis in a rat model of nonunion. Stem Cell Res Ther. 2020;11(1):38.

    Article  PubMed  PubMed Central  Google Scholar 

  327. Huang JH, Chen YN, He H, Fu CH, Xu ZY, Lin FY. Schwann cells-derived exosomes promote functional recovery after spinal cord injury by promoting angiogenesis. Front Cell Neurosci. 2023;16:1077071.

    Article  PubMed  PubMed Central  Google Scholar 

  328. Sun J, Zeng Q, Wu Z, Li Z, Gao Q, Liao Z, et al. Enhancing intraneural revascularization following peripheral nerve injury through hypoxic Schwann-cell-derived exosomes: an insight into endothelial glycolysis. J Nanobiotechnol. 2024;22(1):283.

    Article  CAS  Google Scholar 

  329. Kamburoğlu HO, Safak T, Ersoy US, Ocal E, Evrenos MK, Sonmez E, et al. A new flap design: prefabricated neuro-osseous flap. Ann Plast Surg. 2013;70(3):317–23.

    Article  PubMed  Google Scholar 

  330. Wang CY, Chai YM, Wen G, Han P. One-stage reconstruction of composite extremity defects with a sural neurocutaneous flap and a vascularized fibular graft: a novel chimeric flap based on the peroneal artery. Plast Reconstr Surg. 2013;132(3):428e-e437.

    Article  PubMed  CAS  Google Scholar 

  331. Wang C, Xu J, Wen G, Chai Y. Reconstruction of complex tissue defect of forearm with a chimeric flap composed of a sural neurocutaneous flap and a vascularized fibular graft: a case report. Microsurgery. 2018;38(7):790–4.

    Article  PubMed  Google Scholar 

  332. Tanaka K, Okazaki M, Homma T, Yano T, Mori H. Bilateral inferior alveolar nerve reconstruction with a vascularized sural nerve graft included in a free fibular osteocutaneous flap after segmental mandibulectomy. Head Neck. 2016;38(5):E111–4.

    Article  PubMed  Google Scholar 

  333. Zhang B, Li KY, Jiang LC, Meng Z, Wang XM, Cui FZ, et al. Rib composite flap with intercostal nerve and internal thoracic vessels for mandibular reconstruction. J Craniofac Surg. 2016;27(7):1815–8.

    Article  PubMed  PubMed Central  Google Scholar 

  334. Wang L, Wei JH, Yang X, Yang ZH, Sun MY, Cheng XB, et al. Preventing early-stage graft bone resorption by simultaneous innervation: innervated iliac bone flap for mandibular reconstruction. Plast Reconstr Surg. 2017;139(5):1152e–61e.

    Article  PubMed  CAS  Google Scholar 

  335. Abdelrehem A, Shi J, Wang X, Wu Z, Mashrah MA, Zhang C, et al. Novel loop neurorrhaphy technique to preserve lower lip sensate in mandibular reconstruction using an innervated vascularized iliac bone flap. Head Neck. 2022;44(1):46–58.

    Article  PubMed  Google Scholar 

  336. Shi J, Zhang Y, Zhang B, Wu Z, Gupta A, Wang J, et al. Loop-neurorrhaphy technique for preventing bone resorption and preserving lower lip sensation in mandibular reconstruction using vascularized iliac bone flap: a single-center randomized clinical trial. Plast Reconstr Surg. 2024;154(5):1004e-e1014.

    Article  PubMed  CAS  Google Scholar 

  337. El-Rashidy AA, Roether JA, Harhaus L, Kneser U, Boccaccini AR. Regenerating bone with bioactive glass scaffolds: a review of in vivo studies in bone defect models. Acta Biomater. 2017;62:1–28.

    Article  PubMed  CAS  Google Scholar 

  338. Pearce AI, Richards RG, Milz S, Schneider E, Pearce SG. Animal models for implant biomaterial research in bone: a review. Eur Cell Mater. 2007;13:1–10.

    Article  PubMed  CAS  Google Scholar 

  339. Muschler GF, Raut VP, Patterson TE, Wenke JC, Hollinger JO. The design and use of animal models for translational research in bone tissue engineering and regenerative medicine. Tissue Eng Part B Rev. 2010;16(1):123–45.

    Article  PubMed  Google Scholar 

  340. Wang X, Mabrey JD, Agrawal CM. An interspecies comparison of bone fracture properties. Biomed Mater Eng. 1998;8(1):1–9.

    PubMed  CAS  Google Scholar 

  341. Kabata T, Kubo T, Matsumoto T, Hirata T, Fujioka M, Takahashi KA, et al. Onset of steroid-induced osteonecrosis in rabbits and its relationship to hyperlipaemia and increased free fatty acids. Rheumatology (Oxford). 2005;44(10):1233–7.

    Article  PubMed  CAS  Google Scholar 

  342. Miyanishi K, Yamamoto T, Irisa T, Yamashita A, Jingushi S, Noguchi Y, et al. Bone marrow fat cell enlargement and a rise in intraosseous pressure in steroid-treated rabbits with osteonecrosis. Bone. 2002;30(1):185–90.

    Article  PubMed  CAS  Google Scholar 

  343. Fu J, Xiang Y, Ni M, Qu X, Zhou Y, Hao L, et al. In vivo reconstruction of the acetabular bone defect by the individualized three-dimensional printed porous augment in a swine model. Biomed Res Int. 2020;2020:4542302.

    Article  PubMed  PubMed Central  Google Scholar 

  344. Aerssens J, Boonen S, Lowet G, Dequeker J. Interspecies differences in bone composition, density, and quality: potential implications for in vivo bone research. Endocrinology. 1998;139(2):663–70.

    Article  PubMed  CAS  Google Scholar 

  345. Thorwarth M, Schultze-Mosgau S, Kessler P, Wiltfang J, Schlegel KA. Bone regeneration in osseous defects using a resorbable nanoparticular hydroxyapatite. J Oral Maxillofac Surg. 2005;63(11):1626–33.

    Article  PubMed  Google Scholar 

  346. Newman E, Turner AS, Wark JD. The potential of sheep for the study of osteopenia: current status and comparison with other animal models. Bone. 1995;16(4 Suppl):277S-S284.

    Article  PubMed  CAS  Google Scholar 

  347. Reichert JC, Cipitria A, Epari DR, Saifzadeh S, Krishnakanth P, Berner A, et al. A tissue engineering solution for segmental defect regeneration in load-bearing long bones. Sci Transl Med. 2012;4(141):141ra93.

    Article  PubMed  Google Scholar 

  348. James AW, LaChaud G, Shen J, Asatrian G, Nguyen V, Zhang X, et al. A review of the clinical side effects of bone morphogenetic protein-2. Tissue Eng Part B Rev. 2016;22(4):284–97.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  349. Gillman CE, Jayasuriya AC. FDA-approved bone grafts and bone graft substitute devices in bone regeneration. Mater Sci Eng C Mater Biol Appl. 2021;130:112466.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  350. Nguyen V, Meyers CA, Yan N, Agarwal S, Levi B, James AW. BMP-2-induced bone formation and neural inflammation. J Orthop. 2017;14(2):252–6.

    Article  PubMed  PubMed Central  Google Scholar 

  351. Sun Y, Helmholz H, Willumeit-Römer R. Surgical classification for preclinical rat femoral bone defect model: standardization based on systematic review, anatomical analysis and virtual surgery. Bioengineering (Basel). 2022;9(9):476.

    Article  PubMed  Google Scholar 

  352. He H, Liu L, Morin EE, Liu M, Schwendeman A. Survey of clinical translation of cancer nanomedicines-lessons learned from successes and failures. Acc Chem Res. 2019;52(9):2445–61.

    Article  PubMed  CAS  Google Scholar 

  353. Wang L, Guo X, Chen J, Zhen Z, Cao B, Wan W, et al. Key considerations on the development of biodegradable biomaterials for clinical translation of medical devices: with cartilage repair products as an example. Bioact Mater. 2022;9:332–42.

    PubMed  CAS  Google Scholar 

  354. Guo X, Ma Y, Min Y, Sun J, Shi X, Gao G, et al. Progress and prospect of technical and regulatory challenges on tissue-engineered cartilage as therapeutic combination product. Bioact Mater. 2022;20:501–18.

    PubMed  PubMed Central  Google Scholar 

  355. Bahmaee H, Owen R, Boyle L, Perrault CM, Reilly GC, García Granada AA, et al. Design and evaluation of an osteogenesis-on-a-chip microfluidic device incorporating 3D cell culture. Front Bioeng Biotechnol. 2020;8:557111.

    Article  PubMed  PubMed Central  Google Scholar 

  356. Thakar RG, Fenton KN. Bioethical implications of organ-on-a-chip on modernizing drug development. Artif Organs. 2023;47(10):1553–8.

    Article  PubMed  PubMed Central  Google Scholar 

  357. Zhao X, Li N, Zhang Z, Hong J, Zhang X, Hao Y, et al. Beyond hype: unveiling the real challenges in clinical translation of 3D printed bone scaffolds and the fresh prospects of bioprinted organoids. J Nanobiotechnol. 2024;22(1):500.

    Article  Google Scholar 

  358. Dai K, Wang J, Liu C. Biomaterial-assisted therapeutic cell production and modification in vivo. Exp Hematol. 2024;133:104192.

    Article  PubMed  CAS  Google Scholar 

  359. Park CS, Ha TH, Kim M, Raja N, Yun HS, Sung MJ, et al. Fast and sensitive near-infrared fluorescent probes for ALP detection and 3d printed calcium phosphate scaffold imaging in vivo. Biosens Bioelectron. 2018;105:151–8.

    Article  PubMed  Google Scholar 

  360. Chen L, Tang S, Zhang J, Zhong C, Xu X, Yan J, et al. Prussian blue nanohybridized multicellular spheroids as composite engraftment for antioxidant bone regeneration and photoacoustic tomography. ACS Nano. 2024;18(36):24770–83.

    Article  PubMed  CAS  Google Scholar 

  361. Feng Q, Fatima K, Yang A, Li C, Chen S, Yang G, et al. Multi-modal imaging for dynamic visualization of osteogenesis and implant degradation in 3D bioprinted scaffolds. Bioact Mater. 2024;37:119–31.

    PubMed  PubMed Central  CAS  Google Scholar 

  362. Blázquez-Carmona P, Sanchez-Raya M, Mora-Macías J, Gómez-Galán JA, Domínguez J, Reina-Romo E. Real-time wireless platform for in vivo monitoring of bone regeneration. Sensors (Basel). 2020;20(16):4591.

    Article  PubMed  Google Scholar 

  363. Kalasin S, Sangnuang P, Surareungchai W. Intelligent wearable sensors interconnected with advanced wound dressing bandages for contactless chronic skin monitoring: artificial intelligence for predicting tissue regeneration. Anal Chem. 2022;94(18):6842–52.

    Article  PubMed  CAS  Google Scholar 

  364. Nguyen N, Lin ZH, Barman SR, Korupalli C, Cheng JY, Song NX, et al. Engineering an integrated electroactive dressing to accelerate wound healing and monitor noninvasively progress of healing. Nano Energy. 2022;99:107393.

    Article  CAS  Google Scholar 

  365. Zhang Y, Hu Y, Montelongo Y, Hsu M, Blyth J, Jiang N, et al. A conformable holographic sensing bandage for wound monitoring. Adv Funct Mater. 2024;34(16):2308490.

    Article  CAS  Google Scholar 

  366. Parker RS, Nazzal MK, Morris AJ, Fehrenbacher JC, White FA, Kacena MA, et al. Role of the neurologic system in fracture healing: an extensive review. Curr Osteoporos Rep. 2024;22(1):205–16.

    Article  PubMed  PubMed Central  Google Scholar 

  367. Roemer FW, Hochberg MC, Carrino JA, Kompel AJ, Diaz L, Hayashi D, et al. Role of imaging for eligibility and safety of a-NGF clinical trials. Ther Adv Musculoskelet Dis. 2023;15:1759720X231171768.

    Article  PubMed  PubMed Central  Google Scholar 

  368. Wang W, Gong Z, Wang K, Tian M, Zhang Y, Li X, et al. Activation of the BMP2-SMAD1-CGRP pathway in dorsal root ganglia contributes to bone cancer pain in a rat model. Heliyon. 2024;10(6):e27350.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  369. Fan Y, Zhang B, Du X, Wang B, Yan Q, Guo L, et al. Regulating tumorigenicity and cancer metastasis through TRKA signaling. Curr Cancer Drug Targets. 2024;24(3):271–87.

    Article  PubMed  CAS  Google Scholar 

  370. Lee J, Chung SW. Deep learning for orthopedic disease based on medical image analysis: present and future. Appl Sci. 2022;12(2):681.

    Article  CAS  Google Scholar 

  371. Choi AH. The finite element approach/bone remodeling and osseointegration of implants. Singapore: Springer; 2023. p. 7–21.

    Google Scholar 

  372. Bai L, Xia Z, Triffitt JT, Su J. Generation artificial intelligence (GenAI) and biomaterials translational: steering innovation without misdirection. Biomater Transl. 2024;5(1):1–2.

    PubMed  PubMed Central  Google Scholar 

  373. Benjamens S, Dhunnoo P, Meskó B. The state of artificial intelligence-based FDA-approved medical devices and algorithms: an online database. NPJ Digit Med. 2020;3:118.

    Article  PubMed  PubMed Central  Google Scholar 

  374. Li J, Chen J, Bai H, Wang H, Hao S, Ding Y, et al. An overview of organs-on-chips based on deep learning. Research (Wash D C). 2022;2022:9869518.

    PubMed  Google Scholar 

  375. Qi R, Zou Q. Trends and potential of machine learning and deep learning in drug study at single-cell level. Research (Wash D C). 2023;6:0050.

    PubMed  CAS  Google Scholar 

  376. Lenskjold A, Brejnebøl MW, Rose MH, Gudbergsen H, Chaudhari A, Troelsen A, et al. Artificial intelligence tools trained on human-labeled data reflect human biases: a case study in a large clinical consecutive knee osteoarthritis cohort. Sci Rep. 2024;14(1):26782.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This work was supported by the Foundation of National Clinical Research Center for Oral Diseases (LCA202204), the Key Research and Development Program of Shaanxi (2024GH-YBXM-19), the Clinical New Technology Program of Air Force Medical University (LX2023-306), the China Postdoctoral Science Foundation (2019M653969), the Thousand Talents Plan of Shaanxi Province (to Jing Wang), the National Natural Science Foundation of China (82101069, 22205257), the Logistics Independent Research Project of PLA (to Yang Jiao), the Beijing Natural Science Foundation (7242279), the Beijing Nova Program (20230484283), the Beijing Municipal Science & Technology Commission (Z221100007422130), and the Open Project of State Key Laboratory of Trauma and Chemical poisoning (SKLO202401).

Author information

Authors and Affiliations

Authors

Contributions

XLL, YQZ, and LM drafted the main part of the manuscript and reviewed the manuscript. YXA, FW, JYH, and JYH contributed design of tables and figures. FRT modifies and refines the language. JW, YJ, and ZM provided oversight of the drafting of the manuscript and provided substantive improvements. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Zhao Mu, Yang Jiao or Jing Wang.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

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

Li, XL., Zhao, YQ., Miao, L. et al. Strategies for promoting neurovascularization in bone regeneration. Military Med Res 12, 9 (2025). https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s40779-025-00596-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s40779-025-00596-1

Keywords