Abstract
This perspective articulates the transformative role of biomineralization in materials science and biomedical engineering. Establishing fundamental principles through systematic analysis of mineralization categories, dynamic processes, and crystal nucleation/growth mechanisms, the review progresses to contemporary bioinspired applicationsfrom biotemplated nanocarriers for targeted drug delivery to precision tooth remineralization and engineered bone scaffolds. Critical examination of persistent challenges (morphological precision, scalable production, biological template design) precedes discussion of emerging technological vectors: superhydrophilic/hydrophobic interfacial engineering and hybrid composite systems. The discourse extends to diagnostic biosensing platforms and AI-optimized mineralization architectures as frontier applications. Conclusively, the work frames an interdisciplinary convergence of biological, chemical, and engineering paradigms essential for realizing biomineralization’s potential while mapping strategic research directions.
Keywords: biomineralization, mechanism, dynamic crystallization, bioinspired material, biomedical engineering


1. From Micro to Nanoscale: The Biomineralization Paradigm
Biomineralization, the process by which living organisms produce minerals with precise structural and functional properties, is fundamentally important due to its ability to orchestrate material synthesis. This biologically driven process enables the formation of hierarchical structures with exceptional mechanical strength, biocompatibility, and functionality, as seen in natural systems like bone, teeth, and mollusk shells. The unique advantage of biomineralization lies in its capacity to achieve nanoscale control over crystal nucleation, growth, and organization, guided by organic molecules such as proteins and polysaccharides. This precision allows for the tailoring of material properties at molecular and atomic levels, a feat unattainable by conventional synthetic methods. In biomedical applications, biomineralization promises transformative benefits, including the development of bioinspired scaffolds for tissue regeneration that mimic the extracellular matrix, targeted drug delivery systems with controlled release profiles, and advanced implants with enhanced osseointegration. By controlling mineralization process, it becomes possible to engineer materials with optimized porosity, surface chemistry, and mechanical properties, ensuring compatibility with biological systems. Looking forward, biomineralization will drive innovations in personalized medicine, enabling the design of nanostructured biomaterials that respond dynamically to physiological cues, improve therapeutic outcomes, and reduce adverse effects. Additionally, its potential extends to regenerative therapies, where nanoscale mineral constructs could guide cellular differentiation and tissue repair with unprecedented precision, revolutionizing treatments for bone disorders, dental defects, and beyond. This perspective underscores the necessity of harnessing biomineralization’s control to unlock a new era of biomedical engineering solutions with enhanced functionality and clinical efficacy.
1.1. Definition and History of Biomineralization
Biomineralization, a term coined by Erben for a journal in the 1970 (Figure ), refers to the process by which living organisms produce minerals. These biominerals, formed under controlled biological conditions, often display unique properties such as distinct shapes, sizes, crystallinity, and isotopic or trace element compositions, differing from their inorganically formed counterparts. The term “biomineralization” encompasses this complexity.
1.
Developmental history of biomineralization. The milestone events in the history of biomineralization are marked in red. Abbreviations: CaCO3, Calcium Carbonate. ACP, Amorphous Calcium Phosphate. DLP, Dense Liquid Phase. LLPS, Liquid–Liquid Phase Separation.
The progress in biomineralization research has been inseparable from advancements in microscopy techniques and molecular biology (Reviewed by Dauphin). From the initial use of optical microscopes to the more sophisticated electron microscopes, and further to atomic force microscopes capable of manipulating molecules and atoms, these technologies have progressively deepened our understanding of biomineralization from microscale to nanoscale. Concurrently, developments in molecular biology have elucidated the regulatory role of genes in the biomineralization process, paving the way for greater innovation and application in materials science, medicine, and other related fields. Early researchers laid the groundwork for this field with their pioneering observations and theories. For instance, Harting proposed that organic components control bone structure, establishing a foundational concept for biomineralization studies. Frémy delved into the composition of H, C, N, and O in bone tissue, further elucidating the role of organic constituents within bones. Bøggild systematically classified different types of shell layers, while Schmidt focused on the mineralogical aspects of shells, providing essential mineralogical context for understanding biomineralization processes. By mid-20th century, Anderson et al. discovered that the earliest mineral deposits were associated with membrane-bound vesicles, highlighting the critical role of cell membranes in biomineralization. Kitano demonstrated that organic components could control the polymorphic transformation of CaCO3, thereby influencing mineral morphology. Greenawalt et al. revealed the presence of insoluble calcium phosphate particles within mitochondria, underscoring the complexity of intracellular biomineralization. Miles summarized the structural and chemical organization of teeth, offering comprehensive insights into the role of biomineralization in hard tissue formation.
Erben introduced the term “biomineralization”, formally naming the field. This represented a paradigm shift in biomineralization research, transitioning from microscale to nanoscale frameworks. Posner et al. confirmed the presence of ACP during mineralization, emphasizing its importance as a precursor material. Lowenstam showed that organic matrices could induce biomineralization, fostering deeper exploration into the mechanisms of organic–inorganic interactions. Stephen Mann’s research focuses on the nanoscale mechanisms of biomimetic materials and biomineralization processes. He proposed a theory that organic molecules serve as templates for the formation of minerals, and systematically clarified the regulatory role of the inorganic–organic interface in controlling the morphology and structure of nanoscale mineral crystals from the 1990s to the early 21st century. Boonrungsiman et al. found that ACP precursors are transferred to intracellular calcium apatite matrix vesicles, clarifying the details of intracellular mineralization. From 2010 to 2020, the discovery of amorphous nanoscale precursors marked significant progress in the field of biomineralization research. Gilbert et al. emphasized the universal role of amorphous nanoparticles (such as amorphous calcium carbonate (ACC) and amorphous phosphate calcium (ACP)) in processes like bone and tooth formation. He proposed the “particle attachment” mechanism in the early animal biomineralization process, revealing the complexity and flexibility of the biomineralization process at the nanoscale. Albéric et al. explores the interplay between calcite, amorphous calcium carbonate (ACC), and intracrystalline organics within sea urchin skeletal elements, shedding light on their mutual interactions and structural roles. In the 2020s, the multidisciplinary integration of nanobiomineralization with artificial intelligence, computational simulation, and synthetic biology has advanced the precise control and biomimetic applications of biomineralization mechanisms, with Fiona Meldrum from the University of Leedsa renowned hub for crystal mineralization research in the U.K.leading efforts in biomimetic mineralization and nanoscale crystal growth control to develop novel biomimetic materials. , Jin et al. uncovered the mechanism of ACP formation from DLP during LLPS, providing new molecular-level insights into biomineralization processes. The history of biomineralization research traces a remarkable journey of discovery (Figure ), where progressive investigations have systematically unraveled key steps and fundamental mechanisms, thereby establishing a robust scientific foundation for the development of innovative biomaterials and therapeutic strategies.
1.2. Scope and Architecture of This Perspective
This perspective comprehensively explores biomineralizationthe process by which living organisms form minerals with unique hierarchical structures and properties. It first establishes the fundamental mechanisms, including mineralization categories, dynamic processes, and the principles governing mineral nucleation and crystal growth, laying the groundwork for harnessing these natural phenomena. The discussion then focuses on the translational applications of bioinspired mineralization in functional materials, such as biotemplated nanoparticle synthesis, tooth repair/remineralization, and bone tissue engineering, demonstrating its innovative potential for addressing biomedical and engineering challenges. Finally, the perspective critically evaluates current challengesincluding precise control over mineral morphology/functionality, scalable production, and novel biological template developmentwhile exploring emerging technological pathways (e.g., superhydrophilic/hydrophobic surfaces, high-strength hybrid composites) and opportunities in new fields like healthcare diagnostics and biosensing. Integrating biology, chemistry, and engineering, this work provides an interdisciplinary roadmap from fundamental principles to future technological advancements.
2. The Mechanism of Biomineralization
Biomineralization is an extremely unique process in which living organisms can precisely control the formation of highly ordered mineral structures. Exploring the core mechanisms driving biomineralization and revealing how biological systems regulate the formation, structure and function of biominerals are topics of great concern.
2.1. Mineralization Categories
Biomineralization can be classified into two main types: biologically controlled mineralization and biologically induced mineralization. Biological controlled mineralization is the organism-directed formation of hierarchically structured minerals through regulated interactions between biomolecules/cells and ingested metal ions/anions. The process involves synergistic regulation by soluble and insoluble organic phases. Insoluble matrices provide a confined microenvironment for crystal nucleation and growth, acting as a structural template, whereas soluble factors dynamically modulate mineral composition and morphological development. Microbially controlled mineralization involves the direct role of macromolecular organic substrates or living cell structures in the nucleation, growth, and elemental composition of biominerals, with the reactions classified into extracellular, intercellular, and intracellular processes. Living organisms precisely harness the synergistic interplay between soluble and insoluble organic components to synthesize biominerals with tailored properties. The extent of biological control varies across species, yet most mineralization occurs in confined microenvironments. This spatial isolation contributes to the structural complexity and species-specificity of biominerals, ultimately enabling specialized biological functions. Biomineralization of magnetosomes in magnetotactic bacteria represents one of the few well-documented examples of microbially controlled mineralization.
Most microbial mineralization is of the type induced by organisms. The process of microbially induced carbonate precipitation is intricately linked to the metabolic activities of the associated microorganisms, encompassing intricate ion exchange and mass transfer dynamics. Living organisms cause changes in the local microenvironment through metabolic activities, creating physical and chemical conditions conducive to mineral precipitation. This leads to the process of biological mineral precipitation. The products of this process usually do not have specific biological functions. Although organisms have no control over the types and characteristics of minerals formed, their metabolic activities can regulate the physical and chemical conditions of the solution environment, such as pH, CO2 and the composition of metabolites, thereby influencing the formation of specific minerals. For instance, when studying the ability of SheG wanella oneidensis MRG1 to mineralize and form struvite in a bacterial culture system, it was found that different culture media containing the inoculated strain MRG1 all produced struvite, while the control group without inoculation did not form any precipitate and the solution remained clear throughout, indicating that strain MRG1 can promote the formation of struvite. The results of further experiments on changing the culture medium proved that the strain MRG1 produces ammonium and phosphate through its metabolic activities, and increases the pH of the system, creating physical and chemical conditions conducive to the precipitation of struvite. Additionally, the surface of the biological cells may also play an important role during the mineral nucleation induction period, because nucleation often occurs directly on the cell surface, and the formed minerals will also be closely bound to the cell surface. The minerals formed by biological induction exhibit significant heterogeneity. This heterogeneity is manifested in various aspects such as variable external morphology, moisture content, composition of trace elements, crystal structure and particle size.
2.2. Dynamic Processes of Biomineralization
Biomineralization is a regulated process leading to precise mineral deposition (Figure ). In the early stages of biomineralization, Amorphous Calcium Phosphate (ACP) appears featureless under TEM and shows a broad pattern in X-ray diffraction, with a Ca/P ratio of 1.5 or less. Posner et al. identified ACP (Ca9(PO4)6) in the mineralization process supported by further research. The early physiological formation of ACP, explaining that the aggregation of Ca2+ and HPO4 2– forms prenucleation complexes, which then aggregate into 3D polymeric structures preceding the nucleation of spherical ACP. Mineralized tissue formation begins with ACP deposition on fiber surfaces and progresses through continuous calcium uptake, transforming ACP into elongated apatite and ultimately HA, although the existence of an intermediate octacalcium phosphate (OCP) phase is debated. Spherical ACP evolves into step-like homogeneous crystals along a single orientation through a dynamic process, observed by Onuma et al. via AFM studies, showing initial elongation and six distinct growth steps within a 2D layer, which stack to form 3D structures. These CaP particles transition from initial dimensions resembling Posner’s clusters into stable ACP plates before evolving into the final crystalline HA phase.
2.
Diagrammatic models of the biomineralization process in intracellular and extracellular. (A) Intracellular biomineralization process occurs from the endoplasmic reticulum to the mitochondria. (B) Biomineralization process progresses from the mitochondria to intracellular membrane vesicles. (C) The biomineralization process extends from intracellular membrane vesicles to the extracellular space. (D) The formation process of mineralized collagen fibrils takes place. Abbreviations: ER, Endoplasmic Reticulum. SERCA, Sarco/Endoplasmic Reticulum Ca2+-ATPase. IP3Rs, Inositol 1,4,5-trisphosphate Receptors. RyRs, Ryanodine Receptors. OMM, Outer Mitochondrial Membrane. VDACs, Voltage-Dependent Anion Channels. MCU, Mitochondrial Calcium Uniporter. Pi, Inorganic Phosphate. G6 Pase, Glucose-6-Phosphatase. T2, Transport System 2. IMM, Inner Mitochondrial Membrane. PIC, Phosphate Carrier. ACP, Amorphous Calcium Phosphate. MVs, Matrix Vesicles. ENPP1, Ectonucleotide Pyrophosphatase/Phosphodiesterase 1. PPi, Pyrophosphate. TNAP, Tissue-Nonspecific Alkaline Phosphatase. ANK, Progressive Ankylosis protein. Pit1/Pit2, Sodium-dependent Phosphate Transporter 1 and 2. PHOSPHO1, Phosphatase, Orphan 1. ECM, Extracellular Matrix. NCPs, Non-Collagenous Proteins. GFs, Growth Factors. Adapted from reference. Available under a CC BY-NC-ND 4.0 license.
The transfer of Ca2+ from the ER to mitochondria involves Ca2+ influx into the ER via Ca2+ ATPases (Figure A). Ca2+ is released from the ER through IP3Rs or RyRs, inducing accumulation in endoplasmic reticulum-mitochondria encounter structures (ERMES). Increased cytoplasmic Ca2+ stabilizes ERMES contacts regulated by mitochondrial Rho GTPase. Ca2+ crosses the outer mitochondrial membrane (OMM) through VDACs, enhancing IP3R-induced signals and facilitating Ca2+ entry into the mitochondrial intermembrane space and matrix. The inner mitochondrial membrane (IMM), being highly selective and low affinity, presents the next barrier, with Ca2+ passing primarily through the MCU complex, influenced by mitochondrial membrane potential and local Ca2+ concentration. Unlike Ca2+, Pi can enhance mitochondrial mineralization granule formation, implying its transport from the ER. Pi transport via the ER occurs through the glucose-6-phosphatase (G6 Pase) system, where G6 Pase hydrolyzes glucose-6-phosphate to glucose and Pi within the ER lumen. Pi is then transported across the ER membrane by the T2 system. Due to the OMM’s high permeability to molecules under 5 kDa, Pi easily crosses it and is then taken up by the IMM via the PIC, which facilitates Pi influx alongside H+ flow.
Further research showed that mitochondria can absorb Ca2+ and Pi, facilitating the formation of calcium phosphate electron-dense granules. , These granules act as ACP precursors and can be transferred to intracellular calcium phosphate-containing MVs. ACP precursors accumulate in mitochondria, causing swelling and loss of membrane potential, leading to PINK1 accumulation on the outer mitochondrial membrane. PINK1 recruits PARKIN from the cytosol, which ubiquitinates mitochondrial proteins, prompting ACP-overloaded mitochondria to be engulfed by autophagosomes via P62 and LC3. These mitochondria are then isolated by a double membrane to form autophagosomes, which fuse with lysosomes to become autolysosomes where ACP precursors coalesce into larger granules, involving the BMP/Smad signaling pathway. The transfer of ACP precursors from mitochondria to MVs is critical, as Ca2+ influx can induce ROS generation, mitochondrial depolarization, and mtDNA damage.
The mineral density of ACP precursors increases in MVs, where nucleation and maturation lead to the initial formation of ACP particles. MVs contain membrane transporters and enzymes that create a suitable microenvironment for calcium phosphate nucleation and growth. Ca2+ crystallization is easier than Pi due to its strong binding to the negatively charged inner leaflet of the plasma membrane. Phosphatidylcholine and phosphatidylserine in plasma membranes have a high capacity for Ca2+ binding, facilitating its accumulation and crystallization in MVs. Pi enrichment and crystallization depend on Pi concentration and the PPi/Pi ratio. Pi binds with Ca2+ to form calcium phosphate, enhancing mineralization, while PPi, when hydrolyzed into Pi, promotes mineralization but also inhibits it by binding to nascent HA and inhibiting alkaline phosphatase. Phosphatases such as TNAP and ENPP1 regulate the PPi/Pi ratio, with TNAP converting PPi to Pi and ENPP1 generating PPi from ATP. Transporters like ANK and Pit1/Pit2 supply Pi to MVs, and PHOSPHO1 generates Pi from phosphocholine and phosphoethanolamine within MVs. (Figure B). The presence of Ca2+, Pi, NCPs, GFs, and other components in the ECM and bodily fluids, along with artificially added agents such as polycations and polyanions, creates an environment conducive to the further mineralization of MVs containing ACP or ACP particles (Figure C). Local environment promotes the deposition of calcium phosphate on collagen fibrils, facilitating the mineralization process in the ECM and forming of mineralized tropocollagen and collagen fibrils (Figure D).
Jin et al.’s newly study revealed that highly hydrated bicarbonate DLP forms through liquid–liquid phase separation. Solvated Ca2+·(HCO3 –) 2 complexes condense and react within confined DLP droplets, and acidic proteins and polymers extend the DLP’s lifetime without altering its pathway or chemical properties. The DLP then converts into ACPs, which serve as precursors for CaCO3 crystal formation, directly influencing the final mineral’s morphology and structure. Initially, there are difficulties in linking the behavior of calcium phosphate (CaP) particles within cellular organelles during various stages of mineral formation with their eventual deposition outside the cell. The precise regulatory processes within different organelles are not fully understood, nor is the mechanism that selectively enhances or restrains intracellular mineralization under varying conditions completely clarified need to be further explored.
2.3. Mineral Crystal Nucleation and Growth
Biomineralization involves the controlled formation of mineral phases through dynamic nucleation and growth processes, as illustrated in Figure . In the initial stage (Figure A), a supersaturated solution forms due to elevated local ion concentrations, establishing a critical gradient for mineral precipitation. This high ion concentration drives the aggregation of ions into clusters, setting the stage for nucleation. The nucleation phase (Figure B) proceeds through two primary pathways: classical nucleation theory (CNT) and nonclassical crystallization pathways (NCCPs). In CNT, stable nuclei form when the system overcomes the Gibbs free energy barrier (ΔG), which comprises surface free energy (ΔG s) and volumetric free energy (ΔG v). The energy barrier peaks at the critical radius (r c), where ΔG s increases due to interfacial tension, while ΔG v decreases as the bulk phase stabilizes. Beyond r c, ΔG declines, enabling spontaneous nucleus growth once the critical Gibbs free energy threshold (Gcrit) is surpassed. The entire process involves the loss of energy. Alternatively, NCCPs involve the formation of amorphous or metastable phases, such as nanoblocks, which are influenced by environmental factors including pH, temperature, ionic strength, and mechanical stress. These nanoblocks are further stabilized and directed by organic molecules, such as proteins, peptides, and polysaccharides, which regulate their assembly and transformation into crystalline structures. During crystal growth (Figure C), nanoblocks attach to surface templates, initially forming amorphous particles that transition into stable crystalline phases under specific conditions. The presence of organic molecules modulates crystal morphology and ensures structural integrity. Environmental factors and molecular regulation dictate the final crystal forms, resulting in ordered mineral structures characteristic of biomineralization. This interplay of physicochemical and biological factors underscores the complexity and precision of biomineralization processes.
3.
Biomineralization nucleation and growth. (A) Supersaturated solution with high local ion concentrations initiates mineral formation. (B) Nucleation stage. Stable nuclei form via classical nucleation theory (CNT) or nonclassical crystallization pathways (NCCPs) involving amorphous phases overcomes the Gibbs free energy barrier (ΔG) at critical radius (r c). (C) Crystal growth. Nanoblocks rearrange on surface templates under regulation by organic molecules, forming stable crystals. Abbreviations: ΔG, Gibbs free energy change; ΔG s, surface free energy change; ΔG v, volumetric free energy change; r c, critical radius; G crit, critical Gibbs free energy threshold.
The nucleation of calcium phosphate (CaP) compounds in vivo is regulated by intrinsic factors, particularly the concentrations of calcium and orthophosphate ions. Homeostatic mechanisms maintain these ion concentrations within a narrow range, with nearly 99% of calcium stored in the skeleton and small amounts circulating in the body, including calcium bound to plasma proteins and free calcium ions, controlled by interactions involving the bone, intestines, and kidneys, with fluctuations typically not exceeding 5%. Intracellular calcium concentrations are significantly lower compared to extracellular levels, maintained by calcium channels, pumps, mitochondrial uptake, and sequestration in the ER. For phosphorus, around 85% is incorporated into the mineral phase, with over 99% of the body’s phosphorus content located intracellularly, and only a small portion present in the extracellular space, primarily as phosphate ions in the plasma. The ionic product for HA can be calculated using the concentrations of calcium and orthophosphate ions, accounting for the high ionic strength of the biological medium. Comparing this value with HA’s solubility product (K_sp) indicates a metastable supersaturated state, where no spontaneous precipitation occurs. This metastability is also observed with other CaP phases such as OCP, Dicalcium Phosphate Dihydrate (DCPD), and ACP, whose K_sp values are documented. This supersaturation is essential for maintaining the coexistence of these mediums with mineralized HA tissues, preventing it from dissolving into the serum.
Extrinsic factors such as physicochemical conditionsincluding pH, temperature, ionic strength, peptide and proteinsignificantly influence ion activities and thus the solubility of CaP solids. In biological systems, these conditions are maintained in a physiological balance by homeostatic mechanisms. The presence of specific molecules that can also promote or inhibit the nucleation and growth of CaP include inorganic molecules like pyrophosphate (PP_i) and organic compounds, primarily noncollagenous proteins (NCPs). , A significant group of NCPs in mineralized tissues is the small integrin-binding ligand N-linked glycoprotein (SIBLING) family, known for its crucial role in regulating mineralization processes. These proteins share common structural and functional features; they are intrinsically disordered, contain domains that bind to collagen and HA, and possess an RGD motif that binds to cell membranes. Additionally, they undergo post-translational phosphorylation. The inhibitive role of these proteins is crucial in explaining the metastability of biological media concerning HA precipitation, as detailed earlier. For proteins like fetuin, this inhibition is attributed to the formation of complexes with HA precursors such as calcium ions or the sequestration of ACP nanoclusters by phosphopeptides. This binding can be strong enough to kinetically block mineralization or direct growth along specific pathways. However, the precise mechanisms and roles of these mineralization factors can vary significantly depending on the localization of the processes and the surrounding cofactor environment. Understanding CaP nucleation requires addressing the complex environmental conditions influenced by various physiological factors and biochemical processes.
The regulation mechanisms of mineral crystal nucleation and growth can be categorized into five key elements: local environment, matrix composition and arrangement, spatial control of ion transport, structural guidance by organic templates. morphological Shaping. This approach involves the precise control of ionic composition, solubility, supersaturation, nucleation, and crystal growth by manipulating the ionic environment. Polysaccharides and proteins play a crucial role in guiding these processes, as they can interact with ions and influence the formation and stability of mineral phases. By adjusting the concentrations and types of ions present, it is possible to fine-tune the conditions necessary for the desired mineralization outcomes. The global composition of the entire matrix is essential for the overall structure and function of the biomineralized material. This involves the total arrangement and integration of individual, mineral-based building units. By carefully selecting and arranging these components, one can create a matrix that supports the desired mechanical and biological properties. The arrangement of these units can significantly affect the performance and stability of the final biomineralized product. This control mechanism focuses on the formation of minerals or composites within enclosed spaces. By controlling the transport of ions within these confined environments, it is possible to achieve precise spatial control over the mineralization process. This approach ensures that minerals form in specific locations and in the desired configurations, which is critical for applications requiring high levels of structural precision and functionality. Template-mediated mineral formation involves the use of an organic framework that contains defined functional groups. These functional groups can influence the growth process by interacting with mineral ions and guiding their deposition. The organic template acts as a scaffold, providing a surface for mineral nucleation and growth. This method allows for the creation of highly ordered and controlled mineral structures, which can be used for various applications, from biomedical implants to advanced materials. The shaping of minerals into complex geometric forms is achieved through the controlled precipitation of minerals in enclosed spaces. By confining the mineralization process within specific boundaries, it is possible to direct the growth of minerals into intricate shapes and structures. This approach is particularly useful for creating biominerals with specific morphologies that are required for functional or aesthetic purposes, such as in the fabrication of biomimetic materials or in tissue engineering.
3. Bioinspired Mineralization for Functional Materials
Bioinspired mineralization techniques hold immense potential across various fields. In nanoparticle synthesis, biological templates enable precise control over morphology and functionality, enhancing performance and expanding industrial applications. For bone tissue engineering, these strategies produced biomaterials that promote bone regeneration, improving clinical outcomes. Similarly, in dental repair, mimicking natural tooth mineralization can restore damaged tissues effectively.
3.1. Biotemplated Synthesis
Biological templates play a crucial role in the synthesis of biomineral materials by mimicking the complex biomineralization processes. There are three primary types of biological templatesCaCO3 templates, silica/silicates templates, and HA templates. These natural templates not only provide a physical framework that restricts the growth of mineral phases but also influence the nucleation and development of mineral crystals through their surface chemical properties.
CaCO3 serves as an ideal biomimetic template material due to its polymorphism and easily controllable crystallization process. By mimicking the biomineralization process, researchers can utilize CaCO3 to guide the formation of nanoparticles, achieving precise control over particle size, shape, and surface properties. For instance, using CaCO3 microspheres or nanoparticles as templates can facilitate the uniform deposition of metal oxides or other functional nanomaterials, thereby developing new materials with specific performance characteristics. Schloßmacher et al. synthesized elastic calcite spicules up to 300 μm in length and 10 μm in diameter using the ammonia carbonate diffusion method with a solution of low-concentrated calcium chloride and silicatein-α to improve mechanical properties. Zhao et al. used varying imidazolium-based ionic liquids to produce rod-like nanosized crystals of calcite and aragonite in different ratios by mixing equimolar aqueous solutions of CaCO3 and calcium chloride with the ionic liquid. Without additives, rhombohedral calcite formed, while the ionic liquids templated calcite and aragonite mixtures. Xiong et al. developed an efficient method to create biomimetic composites by self-assembling nanocellulose and mineralizing amorphous CaCO3, resulting in materials with excellent mechanical properties, tunable chiral color, and water-triggered switchable photonics, features rare in most artificial mineralized materials.
Silica and its derived silicate compounds are renowned for their excellent chemical stability and tunable porous structures, making them one of the ideal biological templates for constructing biomaterials. Naturally occurring silicon-based structures in biological systems, such as sponge spicules and diatom cell walls, offer inspiration for the artificial synthesis of nanoparticles with complex morphologies. Luan et al. described chitosan-mediated silica nanoparticle formation at pH 5.5 using low-concentrated sodium silicate solutions and chitosan concentrations from 0 to 0.2% w/v. Chitosan accelerated the formation of hybrid silica, resulting in composite particles with an average 100 nm diameter and up to 26% organic content. Mehrali et al. investigated nanoscaled calcium silicate hydrates using calcium nitrate and sodium silicate as starting materials and sodium dodecyl sulfate as an organic additive. Sonication for up to 15 min followed by washing and drying produced uniform needle or sheet-like morphologies at various additive concentrations, contrasting with nonuniform needle-like crystals without the additive. The process leveraged electrostatic attraction and stereochemical matching between sulfate end-groups and calcium ions, enhancing nucleation and reaction with SiO3 2– groups.
Hydroxyapatite (HA), as the primary inorganic component of bones and teeth, exhibits significant potential in biomaterial synthesis due to its high affinity for human tissues and bioactivity. As a calcium phosphate mineral, HA can mimic the body’s mineralization processes, providing a biomimetic environment that facilitates the orderly arrangement and functional modification of nanoparticles. Studies have shown that nanocomposites synthesized using HA as a template not only retain their original mechanical strength and biocompatibility but also acquire additional functional properties such as antibacterial activity and osteoinductivity. Moreover, HA templates can help regulate the release behavior of nanoparticles, making them more suitable as drug carriers or components of biosensors. Gopi et al. synthesized HA using malic acid from commercial and apple sources as an additive. The apple-derived malic acid produced HA with lower crystallinity and smaller crystal sizes, showing strong antimicrobial activity against bacteria like Escherichia coli and Klebsiella due to elevated magnesium, sodium, and zinc ions.
3.2. Tooth Repair and Remineralization
The formation of teeth is a multistage and complex process, primarily consisting of two major components: enamel and dentin. Their development and mineralization processes share similarities but also have distinct characteristics (Figure A–C).
4.
Mineralization of tooth and bone. (A–C) Normal development of tooth. Enamel is marked in blue; Dentin is marked in yellow. (A′–E′) Normal development of bone. Vascular is marked in red. Abbreviations: PTH, Parathyroid Hormone. PTHrP, Parathyroid Hormone-related Protein.
Dentin, composed of collagen and minerals, loses its mineral phase to acid erosion in caries, weakening its mechanical properties when collagen degrades. Minerals in dentin are categorized as intrafibrillar and extrafibrillar, with intrafibrillar minerals crucial for nanoscale mechanical properties. Bioinspired materials have advanced dentin intrafibrillar remineralization by mimicking NCPs involved in mineral regulation. Dentin formation involves odontoblasts secreting an unmineralized matrix composed of collagen, proteoglycans, and NCPs such as dentin sialophosphoprotein (DSPP) and dentin matrix protein 1 (DMP1). Although these NCPs constitute less than 10% of the organic matrix, they play a critical role in facilitating intrafibrillar mineralization (The deposition of minerals within the collagen fibers, which is crucial for the mechanical properties of bone and dentin). NCPs, rich in carboxylic acid and phosphate groups, serve as nucleation sites for minerals, contributing to dentin’s toughness and structure, similar to bone at the nanoscale. Intrafibrillar minerals significantly enhance dentin’s mechanical properties and protect collagen from external threats. DSPP, comprising over 50% of dentin’s noncollagenous matrix, is highly acidic due to abundant aspartic acid and serine/phosphoserine residues, enabling HA nucleation and growth through repetitive Asp-Ser-Ser (DSS) units. These regions are endowed with a high negative charge density, enabling them to effectively capture and bind calcium ions. This electrostatic interaction not only facilitates the local enrichment of calcium ions, but also may, by binding to collagen fibers, “present” calcium ions to the mineralization front, thereby promoting the nucleation of HA.
Wang et al. created an amphiphilic oligopeptide that self-assembles with calcium ions to induce amorphous precursor formation and intrafibrillar mineralization of collagen fibrils, resulting in a 30 mm-thick remineralized layer and occlusion of deep dentinal tubules. Dendrimers, well-defined branched macromolecules with a central core, branches, and terminal groups, mimic globular proteins and have been used to regulate CaCO3 and phosphate crystallization, with PAMAM being the earliest type; modified PAMAM dendrimers like PAMAM-NH2, PAMAM-COOH, and PAMAM–OH with the enamel remineralization values were 76.42 ± 3.32%, 60.07 ± 5.92%, and 54.52 ± 7.81%, respectively. Further studies are needed to understand their interactions with collagen and mineral nucleation. PAMAM dendrimers can also influence HA shape and size through various surface groups and self-assemble similarly to amelogenin, forming nano- and microstructures. Chen et al. showed that PAMAM–PO3H2 produced a thicker remineralized layer on acid-etched enamel compared to PAMAM-COOH, highlighting its potential use in enamel restoration.
Enamel biomineralization, or amelogenesis, involves consecutive stages: secretory and maturation. During the secretory stage, ameloblasts produce enamel matrix proteins and proteinases, controlling initial crystal formation. In the maturation stage, crystal size increases, protein content decreases, forming a highly mineralized tissue. For enamel remineralization, biomineralization-inspired materials must meet several criteria: good biocompatibility to ensure safety when in contact with the oral cavity and digestive system; tight bonding with original enamel to prevent fracture and repeated restoration; similar structure to natural enamel to maintain mechanical properties; and rapid remineralization under physiological conditions to avoid harsh settings and prolonged application times.
Studies show leucine-rich amelogenin peptide (LRAP) induces ACP transformation into aligned, needle-like crystals, and LRAP-mediated remineralization, such as the use of LRAP for acid-etched enamel remineralization, demonstrates faster crystal nucleation and growth compared to full-length amelogenin, likely due to LRAP’s higher hydrophilicity. Dogan et al. designed a 15-amino acid peptide, shADP5, by identifying high-similarity segments in amelogenin and refining these regions. shADP5 formed a dense, 10 μm thick layer of HA crystals on demineralized enamel within 1 h. Biomineralization often involves acidic proteins in β-sheet conformations. Inspired by this, peptides with predefined secondary structures have been used as templates for inorganic mineral crystallization. Kirkham et al. used a self-assembling peptide, P11–4, to treat caries-like lesions, finding that it increased mineral deposition by promoting remineralization and inhibiting demineralization. Yang et al. used ALN-PAMAM-COOH to bind tightly to HA, promoting uniform crystal formation on acid-etched enamel and restoring microhardness. Various forms of calcium phosphate, including ACP, HA, and calcium phosphate ion clusters (CPICs), have been designed for bioinspired enamel regeneration. Tang et al. used triethylamine to stabilize CPICs and created ultrasmall clusters that remained stable without aggregating. This method facilitates the creation of a continuous crystalline–amorphous interface, enhancing the mechanical strength and texture of the repaired enamel, and significantly reducing repair time compared to traditional methods. Inspired by enamel’s intergranular amorphous phase, Wei et al. developed amorphous zirconium dioxide (ZrO2) ceramics for enamel repair, with comparable mechanical properties to native enamel but requires a high-temperature treatment (80 °C, 12 h), limiting clinical application.
3.3. Bone Tissue Engineering
The components and biomineralization processes of dentin and bone are similar. The formation and mineralization of bone are two closely interrelated processes that build and support the skeletal system, encompassing multiple stages from the differentiation of MSCs into osteoblasts, secretion of bone matrix, initial mineralization, further mineralization to maturation and remodeling (Figure A′–E′). Initially, MSCs differentiate into preosteoblasts during skeletal development or repair processes and ultimately mature into osteoblasts. These cells are responsible for producing unmineralized bone matrix (osteoid). Subsequently, osteoblasts secrete large quantities of type I collagen along with noncollagenous proteins such as osteopontin and osteocalcin, forming a soft matrix that provides a scaffold for subsequent mineralization. In the initial mineralization phase, calcium ions and phosphate ions aggregate and form tiny HA crystals, marking the onset of mineralization. As more HA crystals are added, they grow larger and denser, significantly enhancing the hardness and strength of the bone. Concurrently, some osteoblasts transform into osteocytes, which play a role in maintaining bone tissue health. Finally, during the maturation and remodeling phase, the formed bone tissue is both rigid and elastic, capable of adapting to changes in mechanical stress. Osteoclasts resorb aged or damaged bone, while osteoblasts generate new bone at the same site, facilitating bone remodeling and preserving skeletal health. The entire process of bone mineralization is regulated by various growth factors (GFs), hormones such as vitamin D and parathyroid hormone (PTH) and PTH-related protein (PTHrP), enzymes, and other molecules. Adequate nutritional supply, particularly of calcium and phosphorus, is also crucial for supporting normal mineralization.
Synthetic bone grafts with bone-like structures are challenging to create, but biomineralization offers inspiration. Habibovic et al. systematically review bioinspired materials mimicking bone’s ECM, focusing on key properties such as mineral content and distribution homogeneity in scaffolds. Bone biomineralization involves the development of mineral crystals within the ECM of bone tissue (reviewed in ref ). Mineral crystals first nucleate in these gaps and grow into plate-shaped nanocrystals. Bone mineral crystals develop through the transformation of ACP into crystalline HAP, rather than through the classical ion-by-ion addition process. Cui et al. studied the properties and clinical effects of mineralized collagen scaffolds prepared via collagen/apatite self-assembly, finding similarities to natural bone matrix and improving mechanical properties or bioactivities by incorporating polymers, minerals, growth factors, or cells. The scaffold’s biocompatibility and osteogenic ability made it effective in treating various bone defects and diseases. The biomineralization strategies in bone tissue engineering not only mimic the complex structure and function of natural bone but also, through the development of innovative materials and technologies, offer new avenues and hope for achieving more efficient and safer bone repair and regeneration.
4. Challenges and Opportunities
A key biomineralization challenge is precisely controlling mineral formation to tailor material properties. This requires mimicking natural biomolecule-inorganic ion interactions artificially while improving process efficiency and sustainability. Future research will focus on biomolecular control of nucleation/growth and designing novel biomimetic materials with unique functionalities.
4.1. Precise Control Strategies
Deciphering how biologically regulated mineralization through native environments, molecular self-assembly, or engineered scaffolds a fundamental challenge.
In situ mineralization involves the spontaneous formation of minerals on living organisms through ion adsorption and deposition. Cells responsible for mineralization must transport raw materials to deposition sites, often distant from the environmental ion source. Ions are frequently stored temporarily in membrane-bound vesicles before being redissolved, which helps maintain equilibrium between the shell and the environment. Understanding the equilibrium in mineral phase formation rather than focusing on uptake, transport, and storage processes could clarify the reasons behind nonequilibrium phenomena. Positively charged metal ions and negatively charged biomolecules accumulate under electrostatic action to form mineral shells. For example, Wang placed the negatively charged JEV SA14–14–2 vaccine on a calcium-ion-rich medium to create an eggshell-like coating, providing nucleation sites for calcium phosphate (CaP) biomineralization, enhancing thermal stability while preserving immunogenicity. Another example, Ma et al. synthesized CaCO3 from Ca2+ and CO2 produced through respiration, fixing it inside cells with the help of proteins and polysaccharides, thereby creating functional cells that can endogenously produce CaCO3 scaffolds under normal physiological conditions.
Self-assembled mineralization, where nucleation sites are fixed, forms continuous ordered structures through intermolecular self-assembly, emphasizing spontaneity and controllability. This process has potential applications in biological modification and biomimetic material preparation. Inspired by this, Yu et al. used the volatility of triethylamine to create small-sized CaP ion oligomers that spontaneously polymerized into CaP nanofibers under the regulation of linear polymers. Unlike traditionally brittle minerals, these nanofibers were flexible and could be layered into block-like hybrid minerals with flexibility and plasticity. Despite the natural tendency of most organisms to mineralize, active mineralization sites on their surfaces are uncommon. Consequently, to develop biomimetic materials that replicate the structures of living organisms, it is necessary to introduce artificial mineralization induction layers. Inspired by natural pearl nacre, researchers use interface interactions to alternate organic and inorganic materials, creating biomimetic structures. Layer-by-layer self-assembly forms controllable coatings by depositing oppositely charged layers. Wang et al. enhanced yeast cell mineralization using polyelectrolytes with carboxyl or sulfonate groups, creating a functional mineralization shell. Zhang et al. utilized evaporation-induced nanofiber self-assembly in conjunction with resin to develop a composite material with improved performance.
Framework mineralization biomimetics uses a suitable framework to obtain composite materials through mineralization. Despite its similarity to natural biomineralization, current technology can only produce basic multilevel structures due to biological complexity. Scaffolded nucleic acid structures can serve as templates or scaffolds for biomineralization. For instance, DNA nanostructures can precisely locate and control the position and morphology of mineral deposition. By designing specific DNA sequences and structures, it is possible to guide the formation of minerals at the nanoscale into predetermined shapes and sizes. Combining scaffolded nucleic acids with biomineralization can lead to the development of composite materials that possess both organic and inorganic properties. These materials may exhibit enhanced mechanical strength, thermal stability, electrical conductivity, or other physicochemical characteristics. DNA nanostructures have been used to promote the mineralization of HA, resulting in biomimetic bone materials with superior mechanical performance. Scaffolded nucleic acid structures are typically biocompatible and can be modified to introduce specific functional groups or ligands, enhancing their interaction with biomolecules. Utilize the DNA framework to guide the deposition of specific minerals at the lesion site, or encapsulate drugs within the mineral shell to achieve targeted release. Lei et al. developed a biomimetic self-maturation mineralization system using RNA-stabilized amorphous calcium phosphate (RNA-ACP) and ribonuclease (RNase) to induce crystal epitaxial growth.
4.2. Scaling Up and Biocompatibility
As biomineralization technology transitions from laboratory-scale development to industrial-scale production, the biocompatibility of materials must be rigorously ensured to prevent any potential immune responses or toxic effects, thereby guaranteeing their safety and efficacy in vivo. To facilitate the broad application of biomineralized products, researchers are actively exploring innovative strategies to address biocompatibility issues that may arise during large-scale fabrication, ensuring these challenges are met without compromising the superior properties of the materials.
Researchers are relentlessly striving to increase production of biomineralization materials. Yao et al. studied the production scale and biomineralization processes, focusing on enhancing the production scale of biomineralized materials. They explored the use of nonclassical crystallization pathways, such as the involvement of amorphous precursors and liquid precursor phases, to improve the efficiency and scalability of biomineralization. By optimizing these pathways and using polyelectrolytes to enhance nucleation, the study aimed to develop more effective and scalable methods for producing biomineralized materials, which can be applied in various fields, including biological system improvement and biomimetic material design. Zhao et al. studied the production scale and efficiency of biomineralization processes using trained bacterial strains. They focused on enhancing the production scale by optimizing microbial domestication, mutation breeding, targeted screening, and biostimulation. Specifically, they proposed using appropriate enrichment media to promote the growth of native urea-hydrolytic bacteria, which have better environmental adaptability and reduce the costs and environmental risks associated with transporting and releasing non-native bacteria. These methods aim to improve the efficiency and scalability of biomineralization for applications in geotechnical and environmental engineering.
At the same time, biocompatibility cannot be ignored. Silk-based biomaterials, known for their large-scale production and excellent biocompatibility, stand out as promising candidates. Özcan et al. have integrated inorganic components like HA, SiO2, CaCO3, and metals into silk, enhancing their functionalities. Additionally, HA/silk sericin coatings can improve the biocompatibility of metal implants in vivo. Guo et al. studied the biocompatibility and biomineralization of a DCPD coating on additively manufactured NiTi alloy. As a precursor of HA, DCPD effectively promoted HA deposition, enhancing the biomineralization property of the NiTi alloy. Cao et al. developed cell-like microreactors by encapsulating a ruthenium catalyst and hydrophobic organo-photocatalyst within dipeptide coacervates, forming colloidally stable artificial organelles that undergo dimerization and cellular uptake to trigger internal abiotic structures, thereby demonstrating a versatile platform for biocompatible, easily fabricated, and tunable artificial organelles.
4.3. New Biomaterial Paths
Integrating multidisciplinary knowledge and technologies, the development of these advanced biomineralization templates and strategies not only broadens our understanding of biomineralization mechanisms but also paves new ways to address current challenges in materials science.
4.3.1. Superhydrophilic/Hydrophobic Surface Materials
Motivated by the structures of lotus leaves and pearl nacre, Dai et al. used a layer-by-layer self-assembly method to create a novel nacre-like hybrid mesh combining graphene oxide and calcium carbonate (GO-CaCO3). This material featured superhydrophilic properties and underwater superoleophobic behavior. The biomimetic material exhibited strong mechanical properties, boasting a Young’s modulus of 25.4 ± 2.6 GPa, and achieved a separation efficiency of up to 99% for oil–water mixtures, highlighting its potential for oily wastewater treatment. With growing interest in underwater superhydrophobic materials, the development of transparent, mechanically robust superhydrophobic materials for underwater use has become a key area of research. Accordingly, Chen et al. developed a method involving the spontaneous spreading and photo-cross-linking of droplets of chitosan modified with methacrylic anhydride (CSMA) solution at the interface of a superhydrophilic surface and oil. They then biomimetically mineralized negatively charged amorphous calcium carbonate (ACC) onto the surface of the resulting CSMA membrane, producing an underwater superhydrophobic NIM membrane with exceptional performance.
4.3.2. High Strength, High Toughness Hybrid Composite Materials
Spider silk is famous for its remarkable strength, toughness, thermal conductivity, and super shrinkage, surpassing the performance of many synthetic fibers like nylon and carbon fiber. Because of its superior mechanical properties and adaptability, researchers have been motivated to develop new high-strength and high-toughness hybrid composite materials through biomimetic approaches. The interior of spider silk features a network of nanofibers with a structure akin to a fishing net, forming a semicrystalline fiber with a core–shell architecture that contributes to its exceptional durability. With deeper investigations into the structure of spider silk, scientists now employ recombinant spider silk proteins, nonspider silk proteins, and various polymer materialsincluding hydrogels, polyurethanes, and celluloseto engineer artificial fibers that mimic spider silk. Yu et al. used inorganic ion polymerization to create rigid protein crystals and flexible amorphous proteins with HA and poly(vinyl alcohol) (PVA), forming an organic–inorganic integrated structure with hierarchical arrangements, and the biomimetic spider silk hybrid fibers, designed with repetitive protein sequences, achieved performance closely matching that of natural spider silk.
Silk protein is produced by mature sileworms. In mulberry silkworm cocoons, about 97% of the content is made up of silk fibroin and silk sericin. , Extensive research has explored the structure and diverse applications of silk proteins in areas such as tissue engineering, drug delivery, smart textiles, and sensing technologies. Both silk proteins and collagen, as structural proteins, are rich in polar amino acids and form organized fibrous structures through the hierarchical assembly of nanofibrils. The structural and compositional similarities between silk proteins and collagen, a key component in vertebrate bones, highlight the potential of silk for biomineralization studies. The presence of carboxyl, carbonyl, hydroxyl, and amino groups in silk proteins facilitates the binding of ions, promoting the nucleation and growth of inorganic substances. HA/silk sericin composites can be produced via simple precipitation, resulting in low crystallinity (about 60%) similar to natural HA. 100 Higher silk sericin concentrations increase particle size and aggregation but decrease crystallinity. This method is time-consuming, taking over 2 h, while a mesoscopic oscillatory flow reactor can produce the composites four times faster. Cheng et al. used recombinant silk fibroin (RSF) as a template to produce hollow, rice-like CaCO3/RSF hybrids, with the final morphology dependent on the initial pH of the RSF solution. Zhang et al. used silk fibroin microspheres for biomineralization, observing a transformation from dumbbell-shaped to spherical and eventually rhombohedral microparticles as mineralization time increased.
Seashells consist of approximately 95% by volume of inorganic CaCO3 and 5% protein, making them an ideal polymer nanocomposite with high mechanical properties. Seashells form a robust brick-and-mortar mesostructure by integrating hard inorganic components with soft organic components. The layered structure and micropores effectively distribute external impact forces throughout the material. Inspired by the ordered structure of mother of pearl, Rodrigues have used two-dimensional nanosheets and polymers to create organic–inorganic nanocomposites with superior mechanical properties. However, these synthesized materials often exhibit uneven stress distribution, crack propagation, and fracture failures under bending stress compared to natural mother of pearl. To tackle this challenge, Yu et al. integrated ultrasmall calcium phosphate oligomers into PAV and sodium alginate networks, developing a layered composite material with a highly integrated and ordered nanostructure. Due to the strong intermolecular interactions of the small inorganic building blocks within the layered structure, the resulting material demonstrated ultrahigh bending strain (>50%) and toughness, exceeding that of natural mother of pearl and most other synthetic layered materials.
4.4. Emerging Fields of Biomineralization
As interest in biomineralization increases, researchers are moving beyond replicating biological structures to developing functional systems that combine living and material properties, enabling precise control of biological processes. The future studies focus more on the synthesis and applications of the innovative functional systems, particularly in collagen mineralization, hard tissue regeneration, vaccine development, drug delivery, tumor treatment, cell surface engineering, biomimetic artificial organelles, and advancements in energy and environmental technology. Here are some examples of future fields of biomineralization (Figure ).
5.
Future applications of biomineralization.
4.4.1. Healthcare
4.4.1.1. Tissue Repair
Tooth repair through biomineralization is pivotal for restoring the complex structure and function of teeth. Tang’s group developed CaCO3 ion clusters that form a 2.5 μm repair layer within 48 h, achieving a hardness of 3.19 ± 0.05 GPa and an elastic modulus of 84.55 ± 12.38 GPa, closely mimicking natural enamel. Dentin, which supports enamel and protects the pulp, has limited regenerative capacity and is critical for sensing stimuli. To address severe dentin damage and alleviate pain, rapid repair is essential. Liu et al. introduced a biomimetic strategy using CaP oligomers that bind to organic molecules, infiltrating exposed dentin collagen within 5 min to form CaP-collagen complexes, enabling fast remineralization. While these methods show significant potential, developing high-strength repair layers for commercial application requires further research.
Bones can self-repair minor issues, significant defects from trauma, infections, accidents, or tumor removal require bone transplantation or artificial repair materials. Ceramic implants (e.g., hydroxyapatite and tricalcium phosphate) offer biological functionality, while shape-adaptable polymeric materials enable targeted factor delivery for accelerated bone healing. Marelli et al. used biomineralization to create a dense hybrid gel from silk fibroin peptides, achieving a 9-fold increase in compression modulus. The addition of CaCl2 and Na2HPO4 to silk sericin solutions induces the formation of HA nanoneedles, which enhance the viability and osteogenic differentiation of bone marrow MSCs. Advanced techniques, like antibody-functionalized AFM imaging, reveal nanoscale energy transfer, cell interactions, and material properties during mineralization. The sol–gel hybrid assembly method produced a superlight SF/SiO2 aerogel with a honeycomb microstructure, suitable for bone tissue repair. Bone defects can be treated not only by directly supplying calcium but also through a combination of conductive nanomaterials and electrical stimulation, utilizing the bioelectric properties of hard tissues. Yu et al. developed bionic nanoconductive hydrogels with electrical activity, biocompatibility, and bone-inductive capabilities. These hydrogels activate signaling pathways related to bone regeneration, promoting intracellular calcium accumulation, which enhances bone differentiation and mineralization, leading to effective bone healing and regeneration.
4.4.1.2. Vaccine
To address sudden infections and mitigate the threat of highly virulent viruses, the development of vaccines and adjuvants is crucial for public health. Vaccines are commonly used but require significant refrigeration resources to maintain activity due to temperature sensitivity, prompting researchers to use genetic engineering and biomineralization strategies, particularly coating vaccines, to enhance their stability. CaP is noted for its biocompatibility and is frequently used as a transfer agent and adjuvant in vaccines. When calcium ion levels on the viral surface are elevated, CaP forms a mineralized shell through biomineralization. Wang et al. utilized this process to coat an adenovirus vector AIDS vaccine with CaCO3, improving its stability and delaying immune recognition. Further, they enhanced the adhesion of nasal vaccines by coating the viral surface with CaP nanoshells, which significantly increased the antibody response and bioavailability. Viruses can also activate natural immune responses in organisms, which inspired Tang et al. develop a material-based therapeutic strategy. They immobilized wild virus strains within chitosan hydrogels through electrostatic adsorption and supplemented them with CaCO3 nanoparticles to recruit immune cells, creating a highly safe and effective vaccine that forms a targeted immune microenvironment in the body, continuously recruiting immune cells to activate the immune response while preventing virus leakage.
4.4.1.3. Drug Delivery
Biomineralization is currently used in drug delivery because the mineralized shell enhances drug stability and safety while improving therapeutic efficacy by slowing drug metabolism. Qiao et al. created hybrid mineralized nanovesicles (MM@LCaP) by modifying pH-responsive nanoparticles with dexamethasone sodium phosphate (DSP) and calcium, coated with an M2 macrophage membrane, to treat acute respiratory distress syndrome (ARDS). In acidic inflammatory conditions, MM@LCaP released drugs safely, promoting anti-inflammatory macrophage polarization. Wand et al. summarized biomineralization applications in nanoparticle drug delivery systems, the rational design of nanoparticle-based drug delivery systems can address the challenges hindering the clinical translation of nanomaterials.
4.4.1.4. Antibacterial
Through biomineralization, CuS, AgCl, and Ag3PO4 have been effectively combined with silk proteins, resulting in excellent antibacterial activity. , Studies have reported synthesizing AgCl nanocrystals by immersing silk fibers in alternating AgNO3 and NaCl solutions, followed by rinsing. The resulting AgCl/SF nanocomposites show promise as catalysts and antibacterial agents. Recent studies have showcased the versatility of biomineralized nanoflowers in biosensing, including a signal-quenching apt sensor for carcinoembryonic antigen detection, disposable electrodes for ds-DNA and daunorubicin sensing, antimicrobial agents effective against E. coli, a paper-based biosensor for rapid phenylalanine detection, and a smartphone-assisted portable biosensor for real-time epinephrine quantification, demonstrating enhanced sensitivity, selectivity, and practical applications in diagnostics and antimicrobial use.
4.4.1.5. Cancer Therapy
Bioinspired materials, offering enhanced sensitivity, selectivity, and functionality, is of significant potential for cancer therapy area. Wang et al. demonstrated that biomineralized gold nanofibers on silk fibroin were more effective at killing breast cancer cells than spherical gold nanoparticles when injected into the tumor site. Liu et al. created monodisperse vaterite microspheres from CaCO3 and silk fibroin, which encapsulate and release doxorubicin (DOX) in response to pH changes, enabling sustained drug release. Shuai et al. developed pH-responsive HA/silk sericin microspheres that release DOX more rapidly in acidic tumor environments, reducing side effects in normal tissues. Yang et al. synthesized SF@MnO2 nanoparticles to load DOX showed enhanced bioavailability and responded to tumor-specific stimuli. The significant concentration gradient and heterogeneous distribution of calcium ions can lead to abnormal soft tissue deposits like kidney stones and vascular calcification, contributing to cell death. Leveraging this, Zhao et al. developed a drug-free cancer treatment by targeting folate receptors with folate and locally concentrating calcium ions, inducing cancer cell calcification and death, effectively inhibiting tumor growth and metastasis while protecting normal cells. To address the risk of hypercalcemia from direct high-calcium injections, Tang et al. developed a macromolecular drug by cross-linking tumor-targeting folate molecules with carboxyl-rich poly polymers, which indirectly induces calcium accumulation and mineralization on tumor cells, forming a shell that disrupts their nutrient uptake and metabolism, achieving a therapeutic effect.
4.4.2. Biosensing
4.4.2.1. Silica Biosensor
The biosilica microstructures provide multiple benefits for biosensing applications, such as their biocompatibility, chemical stability, and facile functionalization. Moreover, the porous structure of diatom biosilica enables effective immobilization of biomolecules like enzymes, antibodies, and DNA probes, thereby improving the sensitivity and specificity of biosensor devices. Kim et al. presented a highly stretchable and self-healing conductive hydrogel, which was synthesized using marine biomaterials such as catechol, chitosan, and diatom biosilica. Buiculescu et al. develop a new amperometric biosensor, which acetylcholinesterase was covalently immobilized onto gold nanoparticles to enhance stability and electron transfer. Choi et al. developed a high-performance glucose biosensor by encapsulating glucose oxidase within a silica matrix using a self-assembling fusion protein.
4.4.2.2. CaCO3 Biosensor
Recent biomimetic innovations have advanced biosensor development by utilizing CaCO3’s unique properties, exemplified by a self-healing ionic skin that integrates amorphous CaCO3 nanoparticles with alginate and poly(acrylic acid), offering high sensitivity and adaptability. A biomimetic hydrogel created with stabilized ACC nanoparticles via dual-ion doping exhibits exceptional extensibility, toughness, and strain sensitivity, along with biocompatibility and flame retardancy, thus advancing wearable strain sensor technology. Advanced biosensing platforms with enhanced functionalities have been developed by mimicking natural calcification processes, overcoming the limitations of naturally occurring CaCO3 biominerals and paving the way for next-generation biosensors with superior performance.
4.4.2.3. Magnetic Biosensor
A novel magnetic biosensor for ultrasensitive detection of arsenic in water has been developed by genetically engineering magnetotactic bacteria, expanding the field of magnetosome-based biosensors. Magnetosome-nanobody complexes, with high binding affinity for environmental pollutants, have been used to develop a sensitive and specific ELISA for detecting tetrabromobisphenol A. This reusable biosensor is cost-effective, environmentally friendly, and promising for environmental monitoring, food safety, and medical diagnostics. Wu et al. developed a cost-effective and environmentally friendly immunoassay for detecting fipronil in water using Nb-magnetosomes from genetically engineered magnetotactic bacteria, which showed superior binding affinity. Sannigrahi et al. developed a biosensor using a magnetosome-anti-Salmonella antibody complex derived from magnetosomes of Magnetospirillum sp. An electrochemical immunosensor using biomagnetosomes, polyaniline nanogold, and an ionic liquid demonstrated enhanced sensitivity, stability, and specificity for detecting staphylococcal enterotoxin B in milk, and a similar magnetosome-based impedimetric biosensor was developed for detecting white spot syndrome virus in seafood.
4.4.2.4. Hydroxyapatite Biosensor
HA’s superior biocompatibility, adsorption capacity, stability, and sensitivity, along with the enhanced surface area of nanoscale HA, make it an excellent choice for biosensor fabrication, amplifying its potential in various applications. Recent studies have shown that HA-based biosensors, exhibit enhanced sensitivity and versatility through dual-signal amplification, with a detection limit of 50 fg/mL for PDGF-BB, highlighting their potential for various biomarker sensing applications. A label-free HA-based electrochemical biosensor achieved high sensitivity and specificity for detecting microRNA let-7a by leveraging HA nanoparticles’ selectivity for double-stranded nucleic acids, offering a promising and straightforward method for clinical diagnostics. A HA-gold nanocomposite-modified immunosensor on a screen-printed carbon electrode demonstrated high sensitivity and specificity for detecting SARS-CoV-2 antibodies, making it a promising tool for COVID-19 serological diagnosis and vaccine efficacy evaluation. Researchers developed a novel MRI contrast agent (Apt-TDNs-GdHA) by combining gadolinium-doped HA with tetrahedral DNA nanostructures and a high-affinity aptamer (AS1411) targeting nucleolin, resulting in significantly enhanced T1-weighted imaging and improved stability, making it a promising tool for tumor cell imaging and cancer diagnosis.
4.4.3. Newly Areas
4.4.3.1. Environmental Protection
Human activities, especially the mining industry, are the primary sources of arsenic pollution in the environment, resulting in harmful element in solid, liquid, and gaseous wastes. For varying concentrations of arsenic in water, existing physical and chemical removal methods have their limitations, while biomineralization, which uses microorganisms to treat arsenic contamination, is gaining attention for its potential in effectiveness, stability, and application prospects. Sulfate-reducing bacteria reduce sulfate under anaerobic conditions to form sulfides. These sulfides can react with As(V) to form sulfur–arsenic minerals (such as FeAsS), thereby reducing the mobility of arsenic. For instance, bacteria in the Desulfovibrio genus reduce SO4 2– to H2S under anaerobic conditions, and H2S reacts with A(V) to form AsS precipitates. Vera-Espíndola et al. provides a systematic summary focusing on the application of using microorganisms to remove toxic elements like arsenic from effluents.
4.4.3.2. Energy Utilization
Biomineralization for energy production is also a new direction. Xiong et al. to design a core–shell structure by silicifying green algae, allowing continuous hydrogen production under natural oxygen-containing conditions despite hydrogenase’s oxygen sensitivity, by maintaining internal hypoxia while enabling external photosynthesis, thus balancing photosynthetic electron production and hydrogenase activity and providing guidance for future clean, renewable resource generation. Drawing inspiration from plant photosynthesis, Chen et al. extracted thylakoids from spinach leaves and used cell membrane nanocoating technology to deliver these thylakoids to aging cells in animals, enabling stable and controlled production of ATP and NADPH within the cells, which holds promise for tackling the global challenge of delivering energy to cells.
4.4.3.3. Cell Engineering
Researchers have sought to construct stable and controllable functional biomimetic artificial organelles using artificial cell membranes and bioactive substances that form enclosed compartments. Artificial organelles can replicate certain features and behaviors of biological cells, incorporate nonbiological processes and modify cellular functions through tailored design. Zhang et al. created a new type of stable synthetic DNA membraneless organelle (MO) with controlled biological activity based on DNA coacervates. This self-stabilizing and fastener-bound gain-of-function technique enabled the manipulation of vital biological processes and provided fresh insights into the synthesis of membraneless organelles. Wu et al. stabilized ACC using carboxylated cellulose nanofibrils, achieving superior stabilization and enabling the fabrication of transparent composite films with remarkable mechanical properties. Tang et al. proposed a strategy involving the self-assembly of CaP mineralization layers on yeast surfaces, and subsequently, Jin et al. applied this biomimetic approach to encapsulate tetracycline on osteoclast surfaces, enhancing their activity around ectopic calcified tissue and significantly reducing ectopic bone volume. Inspired by this approach, Huang et al. developed a multifunctional polymer to facilitate continuous mineral deposition on tumor cell surfaces through cell surface engineering, aiming to block nutrient, oxygen, and information exchange between tumors and their microenvironment for treatment, addressing the lack of natural mineralization sites on tumor cell.
4.4.3.4. 3D Printing
The use of biological 3D printing in bone tissue engineering has emerged as a recent research highlight. Materials like HA and β-tricalcium phosphate, which have compositions similar to human bone, can be blended with polymers such as polycaprolactone, polylactic acid, polylactic acid-hydroxyacetic acid copolymer, polybenzyl glutamate, and polytrimethylene carbonate to create 3D printable bioinks for repairing bone defects. This not only maintains the biological activity of the materials but also allows for customized repairs suited to specific bone defects, anatomical structures, physiological functions, and patient-specific needs. Shi et al. developed a 3D graft using layered gelatin methacryloyl (GelMA), creating uniform organic–inorganic nanoinks by adding ultrasmall CaP oligomers and bone morphogenetic protein 2 to GelMA precursors. These biomimetic 3D printing technologies can enhance the mechanical properties of bioactive materials to match those of natural bone tissue and facilitate the repair of uniquely shaped bone defects.
4.4.3.5. Concrete Restoration
The biomineralization application of microbial self-healing concrete is crucial, but current research is limited. Adding healing agents with nutrients and bacteria can prolong the setting time of cement and potentially reduce concrete’s mechanical properties. A two-component healing system incorporating alkali-resistant Bacillus spores with calcium lactate precursors has been developed for concrete integration prior to casting, demonstrating mechanical property enhancement through high bacterial concentrations (109 CFU/mL). This approach addresses the inherent vulnerability of concrete to defect formation and crack propagation under prolonged environmental stress. Furthermore, microbial consolidation techniques using B. sphaericus have shown effective limestone prism stabilization and surface protection, confirming biological preservation potential in architectural heritage applications. Kalhori et al. and Bang et al. found that curing with urea and CaCl2 solutions or using Siran beads loaded with Sporosarcina pasteurii in a urea-CaCl2 medium increased concrete’s compressive strength to around 35 MPa. Luo et al. found that microbial healing agents effectively repair cracks in cement paste, highlighting the limitation of narrow crack width and depth for effective microbial repair. Wang et al. developed a pH-responsive chitosan hydrogel to encapsulate B. sphaericus spores, which maintained compressive strength and bridged 32% of cracks in concrete. The selection of carrier materials in microbial self-healing technology significantly influences repair efficacy. Studies demonstrate that vermiculite as a carrier can achieve near-complete crack healing.
5. Conclusion and Prospect
In the past decade, significant progress has been made in understanding the genetic basis and downstream biochemical and mineralogical processes of biomineralization. This field has seen a remarkable convergence of multidisciplinary approaches, integrating biology, chemistry, physics, and engineering to unravel the intricate mechanisms that govern the formation of biominerals. Over the next ten years, a variety of advanced imaging, physical, chemical, and genomic/transcriptomic tools are expected to address unresolved questions related to the molecular mechanisms of biomineralization. These tools include high-resolution microscopy, improved proteomic mass spectrometry, high-resolution spectroscopy, various omics strategies, and computational simulations. It is anticipated that these technologies will soon be integrated with genetic tools, such as reverse genetics, to elucidate the biochemical mechanisms of biomineralization in more complex and advanced organisms.
A new field of artificial intelligence-based biomineralization emerged, a cutting-edge interdisciplinary area that integrates artificial intelligence, biotechnology, and materials science. It has a broad and promising future. With the continuous development of artificial intelligence technology, especially breakthroughs in image recognition, data analysis, and machine learning, new possibilities have been provided for the optimization and control of the biomineralization process. Precise control of the biomineralization process was successfully achieved by combining artificial intelligence with peptide biomimetic materials, creating multilevel fine structures that are difficult to achieve in nature. Moreover, the application of artificial intelligence in biomineralization is also reflected in the dynamic monitoring and prediction of the biomineralization process, for example, through deep learning algorithms to analyze three-dimensional imaging data, thereby more accurately understanding the mechanism of biomineralization. In the field of resource recovery and sustainable development, the combination of artificial intelligence and biomineralization technology also shows great potential. For instance, microbial biomineralization technology has been used in the research of biomineralization of lead and selenium, where bacteria convert soluble metals into stable amorphous mineral products, thereby achieving the removal of heavy metals and resource recovery. At the same time, artificial intelligence technology has also played an important role in optimizing the parameters and conditions of the biomineralization process, for example, through machine learning algorithms to optimize the formulation of biomineral ink to improve its printability and biological activity in 3D bioprinting. These technological advancements not only help improve resource utilization efficiency but also reduce negative impacts on the environment, promoting the realization of green mining and sustainable development.
The integration of these advanced techniques and genetic tools will not only deepen our fundamental understanding of biomineralization but also pave the way for practical applications that can significantly impact human health and environmental sustainability. By harnessing the power of biomineralization, we can develop innovative solutions that address some of the most pressing global challenges, from healthcare to environmental conservation. This multidisciplinary approach promises to open new avenues for scientific discovery and technological innovation, ultimately leading to a more sustainable and resilient future.
Acknowledgments
We thank Chengliang Li with the help of figures. Created in BioRender. Li, C. (2025) https://BioRender.com/dfhtx61.
We are grateful to the National Natural Science Foundation of China (No. 32271298 and T2241002), the National Key Research and Development Program of China (No. 2021YFA1200400), and Wenzhou Institute, University of Chinese Academy of Sciences (No. WIUCASQD2021003 and WIUCASQD2023012).
The authors declare no competing financial interest.
References
- Tang R., Suzuki M.. Special Issue on Biomineralization: From Principles to Practices. ACS Biomater Sci. Eng. 2023;9(4):1730–1732. doi: 10.1021/acsbiomaterials.3c00258. [DOI] [PubMed] [Google Scholar]
- Debnath A., Mitra S., Ghosh S., Sen R.. Understanding microbial biomineralization at the molecular level: recent advances. World J. Microbiol. Biotechnol. 2024;40(10):320. doi: 10.1007/s11274-024-04132-6. [DOI] [PubMed] [Google Scholar]
- Arnold A., Dennison E., Kovacs C. S., Mannstadt M., Rizzoli R., Brandi M. L., Clarke B., Thakker R. V.. Hormonal regulation of biomineralization. Nat. Rev. Endocrinol. 2021;17(5):261–275. doi: 10.1038/s41574-021-00477-2. [DOI] [PubMed] [Google Scholar]
- Ma L., Sun Y., Cheng Q., Yang Z., Wang J., Xu Z., Yang M., Shuai Y.. Silk Protein-Mediated Biomineralization: From Bioinspired Strategies and Advanced Functions to Biomedical Applications. ACS Appl. Mater. Interfaces. 2023;15(28):33191–33206. doi: 10.1021/acsami.3c04067. [DOI] [PubMed] [Google Scholar]
- Kovacs C. S., Chaussain C., Osdoby P., Brandi M. L., Clarke B., Thakker R. V.. The role of biomineralization in disorders of skeletal development and tooth formation. Nat. Rev. Endocrinol. 2021;17(6):336–349. doi: 10.1038/s41574-021-00488-z. [DOI] [PubMed] [Google Scholar]
- Erben K. H.. Ultrastrukturen und Mineralisation rezenter und fossiler Eischalen bei Vogeln und Reptilien. Biominer. Forschungsber. 1970;1:1–65. [Google Scholar]
- Dauphin Y.. A Brief History of Biomineralization Studies. ACS Biomater. Sci. Eng. 2023;9(4):1774–1790. doi: 10.1021/acsbiomaterials.1c01385. [DOI] [PubMed] [Google Scholar]
- Harting, P. Recherches de morphologie synthétique sur la production artificielle de quelques formations calcaires organiques. Van der post, 1872.
- Frémy E.. Recherches chimiques sur les os. Ann. Chim. Phys. 1855;43:47–107. [Google Scholar]
- Boggild O. B.. The Shell Structure of the Mollusks. K. Dan. Vidensk. Selsk. Skr., Naturvidensk. Math. Afd. 1930;(2):235–326. [Google Scholar]
- a Schmidt C.. Zur vergleichenden Physiologie der wirbellosen Thiere. Justus Liebigs Ann. Chem. 1845;54(3):284–330. doi: 10.1002/jlac.18450540303. [DOI] [Google Scholar]; b Schmidt, W. J. ; Keil, A. . Polarizing Microscopy of Dental Tissues: Theory, Methods, and Results from the Structural Analysis of Normal and Diseased Hard, Dental Tissues and Tissues Associated with Them in Man and Other Vertebrates; Pergamon Press, 1971. [Google Scholar]
- Anderson H. C.. Electron microscopic studies of induced cartilage development and calcification. J. Cell Biol. 1967;35(1):81–101. doi: 10.1083/jcb.35.1.81. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kitano Y., Hood D. W.. The influence of organic material on the polymorphic crystallization of calcium carbonate. Geochim. Cosmochim. Acta. 1965;29(1):29–41. doi: 10.1016/0016-7037(65)90075-X. [DOI] [Google Scholar]
- Greenawalt J. W., Rossi C. S., Lehninger A. L.. Effect of active accumulation of calcium and phosphate ions on the structure of rat liver mitochondria. J. Cell Biol. 1964;23(1):21–38. doi: 10.1083/jcb.23.1.21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Miles, A. E. W. Structural and Chemical Organization of Teeth Vol. II; Academic Press, 1967. [Google Scholar]
- Posner A. S., Betts F.. Synthetic amorphous calcium phosphate and its relation to bone mineral structure. Acc. Chem. Res. 1975;8(8):273–281. doi: 10.1021/ar50092a003. [DOI] [Google Scholar]
- Lowenstam H. A.. Minerals Formed by Organisms. Science. 1981;211(4487):1126–1131. doi: 10.1126/science.7008198. [DOI] [PubMed] [Google Scholar]
- Mann S., Archibald D. D., Didymus J. M., Douglas T., Heywood B. R., Meldrum F. C., Reeves N. J.. Crystallization at Inorganic-organic Interfaces: Biominerals and Biomimetic Synthesis. Science. 1993;261(5126):1286–1292. doi: 10.1126/science.261.5126.1286. [DOI] [PubMed] [Google Scholar]
- Boonrungsiman S., Gentleman E., Carzaniga R., Evans N. D., McComb D. W., Porter A. E., Stevens M. M.. The role of intracellular calcium phosphate in osteoblast-mediated bone apatite formation. Proc. Natl. Acad. Sci. U.S.A. 2012;109(35):14170–14175. doi: 10.1073/pnas.1208916109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gilbert P. U. P. A., Porter S. M., Sun C. Y., Xiao S., Gibson B. M., Shenkar N., Knoll A. H.. Biomineralization by particle attachment in early animals. Proc. Natl. Acad. Sci. U.S.A. 2019;116(36):17659–17665. doi: 10.1073/pnas.1902273116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Albéric M., Caspi E. N., Bennet M., Ajili W., Nassif N., Azaïs T., Berner A., Fratzl P., Zolotoyabko E., Bertinetti L., Politi Y.. Interplay between Calcite, Amorphous Calcium Carbonate, and Intracrystalline Organics in Sea Urchin Skeletal Elements. Cryst. Growth Des. 2018;18(4):2189–2201. doi: 10.1021/acs.cgd.7b01622. [DOI] [Google Scholar]
- a De Yoreo J. J., Gilbert P. U., Sommerdijk N. A., Penn R. L., Whitelam S., Joester D., Zhang H., Rimer J. D., Navrotsky A., Banfield J. F.. et al. CRYSTAL GROWTH. Crystallization by particle attachment in synthetic, biogenic, and geologic environments. Science. 2015;349(6247):aaa6760. doi: 10.1126/science.aaa6760. [DOI] [PubMed] [Google Scholar]; b Meldrum F. C., O’Shaughnessy C.. Crystallization in Confinement. Adv. Mater. 2020;32(31):2001068. doi: 10.1002/adma.202001068. [DOI] [PubMed] [Google Scholar]; c Xu Y., Nudelman F., Eren E. D., Wirix M. J. M., Cantaert B., Nijhuis W. H., Hermida-Merino D., Portale G., Bomans P. H. H., Ottmann C.. et al. Intermolecular channels direct crystal orientation in mineralized collagen. Nat. Commun. 2020;11(1):5068. doi: 10.1038/s41467-020-18846-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jin B., Chen Y., Pyles H., Baer M. D., Legg B. A., Wang Z., Washton N. M., Mueller K. T., Baker D., Schenter G. K.. et al. Formation, chemical evolution and solidification of the dense liquid phase of calcium (bi)carbonate. Nat. Mater. 2025;24(1):125–132. doi: 10.1038/s41563-024-02025-5. [DOI] [PubMed] [Google Scholar]
- De Muynck W., De Belie N., Verstraete W.. Microbial carbonate precipitation in construction materials: A review. Ecol. Eng. 2010;36(2):118–136. doi: 10.1016/j.ecoleng.2009.02.006. [DOI] [Google Scholar]
- a Bazylinski D. A., Frankel R. B.. Biologically Controlled Mineralization in Prokaryotes. Rev. Mineral. Geochem. 2003;54(1):217–247. doi: 10.2113/0540217. [DOI] [Google Scholar]; b Lefèvre C. T., Pósfai M., Abreu F., Lins U., Frankel R. B., Bazylinski D. A.. Morphological features of elongated-anisotropic magnetosome crystals in magnetotactic bacteria of the Nitrospirae phylum and the Deltaproteobacteria class. Earth Planet. Sci. Lett. 2011;312(1):194–200. doi: 10.1016/j.epsl.2011.10.003. [DOI] [Google Scholar]
- a Abu Tawila Z. M., Ismail S., Dadrasnia A., Usman M. M.. Production and Characterization of a Bioflocculant Produced by Bacillus salmalaya 139SI-7 and Its Applications in Wastewater Treatment. Molecules. 2018;23(10):2689. doi: 10.3390/molecules23102689. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Dupraz C., Reid R. P., Braissant O., Decho A. W., Norman R. S., Visscher P. T.. Processes of carbonate precipitation in modern microbial mats. Earth-Sci. Rev. 2009;96(3):141–162. doi: 10.1016/j.earscirev.2008.10.005. [DOI] [Google Scholar]; c Ronholm J., Schumann D., Sapers H. M., Izawa M., Applin D., Berg B., Mann P., Vali H., Flemming R. L., Cloutis E. A., Whyte L. G.. A mineralogical characterization of biogenic calcium carbonates precipitated by heterotrophic bacteria isolated from cryophilic polar regions. Geobiology. 2014;12(6):542–556. doi: 10.1111/gbi.12102. [DOI] [PubMed] [Google Scholar]
- Li H., Yao Q.-Z., Yu S.-H., Huang Y.-R., Chen X.-D., Fu S.-Q., Zhou G.-T.. Bacterially mediated morphogenesis of struvite and its implication for phosphorus recovery. Am. Mineral. 2017;102(2):381–390. doi: 10.2138/am-2017-5687. [DOI] [Google Scholar]
- Rabadjieva D., Gergulova R., Titorenkova R., Tepavitcharova S., Dyulgerova E., Balarew C., Petrov O.. Biomimetic transformations of amorphous calcium phosphate: kinetic and thermodynamic studies. J. Mater. Sci.: Mater. Med. 2010;21:2501–2509. doi: 10.1007/s10856-010-4103-8. [DOI] [PubMed] [Google Scholar]
- a Nudelman F., Pieterse K., George A., Bomans P. H., Friedrich H., Brylka L. J., Hilbers P. A., de With G., Sommerdijk N. A.. The role of collagen in bone apatite formation in the presence of hydroxyapatite nucleation inhibitors. Nat. Mater. 2010;9(12):1004–1009. doi: 10.1038/nmat2875. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Dey A., Bomans P. H., Müller F. A., Will J., Frederik P. M., de With G., Sommerdijk N. A.. The role of prenucleation clusters in surface-induced calcium phosphate crystallization. Nat. Mater. 2010;9(12):1010–1014. doi: 10.1038/nmat2900. [DOI] [PubMed] [Google Scholar]
- Habraken W. J. E. M., Tao J., Brylka L. J., Friedrich H., Bertinetti L., Schenk A. S., Verch A., Dmitrovic V., Bomans P. H., Frederik P. M.. et al. Ion-association complexes unite classical and non-classical theories for the biomimetic nucleation of calcium phosphate. Nat. Commun. 2013;4(1):1507. doi: 10.1038/ncomms2490. [DOI] [PubMed] [Google Scholar]
- Lotsari A., Rajasekharan A. K., Halvarsson M., Andersson M.. Transformation of amorphous calcium phosphate to bone-like apatite. Nat. Commun. 2018;9(1):4170. doi: 10.1038/s41467-018-06570-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Onuma K., Ito A.. Cluster Growth Model for Hydroxyapatite. Chem. Mater. 1998;10(11):3346–3351. doi: 10.1021/cm980062c. [DOI] [Google Scholar]
- Wang L., Li S., Ruiz-Agudo E., Putnis C. V., Putnis A.. Posner’s cluster revisited: direct imaging of nucleation and growth of nanoscale calcium phosphate clusters at the calcite-water interface. CrystEngComm. 2012;14(19):6252–6256. doi: 10.1039/c2ce25669j. [DOI] [Google Scholar]
- Yan X., Zhang Q., Ma X., Zhong Y., Tang H., Mai S.. The mechanism of biomineralization: Progress in mineralization from intracellular generation to extracellular deposition. Jpn. Dent. Sci. Rev. 2023;59:181–190. doi: 10.1016/j.jdsr.2023.06.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Prakriya M., Lewis R. S.. Store-operated calcium channels. Physiol. Rev. 2015;95(4):1383–1436. doi: 10.1152/physrev.00020.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Berridge M. J.. The inositol trisphosphate/calcium signaling pathway in health and disease. Physiol. Rev. 2016;96(4):1261–1296. doi: 10.1152/physrev.00006.2016. [DOI] [PubMed] [Google Scholar]
- Lee S., Lee K.-S., Huh S., Liu S., Lee D.-Y., Hong S. H., Yu K., Lu B.. Polo kinase phosphorylates miro to control ER-mitochondria contact sites and mitochondrial Ca2+ homeostasis in neural stem cell development. Dev. Cell. 2016;37(2):174–189. doi: 10.1016/j.devcel.2016.03.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- De Stefani D., Bononi A., Romagnoli A., Messina A., De Pinto V., Pinton P., Rizzuto R.. VDAC1 selectively transfers apoptotic Ca2+ signals to mitochondria. Cell Death Differ. 2012;19(2):267–273. doi: 10.1038/cdd.2011.92. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dong Z., Shanmughapriya S., Tomar D., Siddiqui N., Lynch S., Nemani N., Breves S. L., Zhang X., Tripathi A., Palaniappan P.. et al. Mitochondrial Ca2+ uniporter is a mitochondrial luminal redox sensor that augments MCU channel activity. Mol. Cell. 2017;65(6):1014–1028. doi: 10.1016/j.molcel.2017.01.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Tang C., Wei Y., Gu L., Zhang Q., Li M., Yuan G., He Y., Huang L., Liu Y., Zhang Y.. Biomineral precursor formation is initiated by transporting calcium and phosphorus clusters from the endoplasmic reticulum to mitochondria. Adv. Sci. 2020;7(8):1902536. doi: 10.1002/advs.201902536. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Xia W., Meng X., Jiang Y., Li M., Xing X., Pang L., Wang O., Pei Y., Yu L.-Y., Sun Y.. et al. Three novel mutations of the PHEX gene in three Chinese families with X-linked dominant hypophosphatemic rickets. Calcif. Tissue Int. 2007;81:415–420. doi: 10.1007/s00223-007-9067-4. [DOI] [PubMed] [Google Scholar]
- Gerin I., Noël Gt., Van Schaftingen E.. Novel arguments in favor of the substrate-transport model of glucose-6-phosphatase. Diabetes. 2001;50(7):1531–1538. doi: 10.2337/diabetes.50.7.1531. [DOI] [PubMed] [Google Scholar]
- Nordlie R. C., Scott H., Waddell I., Hume R., Burchell A.. Analysis of human hepatic microsomal glucose-6-phosphatase in clinical conditions where the T2 pyrophosphate/phosphate transport protein is absent. Biochem. J. 1992;281(3):859–863. doi: 10.1042/bj2810859. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Seifert E. L., Ligeti E., Mayr J. A., Sondheimer N., Hajnóczky G.. The mitochondrial phosphate carrier: role in oxidative metabolism, calcium handling and mitochondrial disease. Biochem. Biophys. Res. Commun. 2015;464(2):369–375. doi: 10.1016/j.bbrc.2015.06.031. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shapiro I. M., Lee N. H.. Calcium accumulation by chondrocyte mitochondria. Clin. Orthop. Relat. Res. 1975;106(1976–2007):323–329. doi: 10.1097/00003086-197501000-00044. [DOI] [PubMed] [Google Scholar]
- Brookes P. S., Yoon Y., Robotham J. L., Anders M., Sheu S.-S.. Calcium, ATP, and ROS: a mitochondrial love-hate triangle. Am. J. Physiol.-Cell Physiol. 2004;287(4):C817–C833. doi: 10.1152/ajpcell.00139.2004. [DOI] [PubMed] [Google Scholar]
- Hasegawa T., Hongo H., Yamamoto T., Abe M., Yoshino H., Haraguchi-Kitakamae M., Ishizu H., Shimizu T., Iwasaki N., Amizuka N.. Matrix vesicle-mediated mineralization and osteocytic regulation of bone mineralization. Int. J. Mol. Sci. 2022;23(17):9941. doi: 10.3390/ijms23179941. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tadross M. R., Tsien R. W., Yue D. T.. Ca2+ channel nanodomains boost local Ca2+ amplitude. Proc. Natl. Acad. Sci. U.S.A. 2013;110(39):15794–15799. doi: 10.1073/pnas.1313898110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Melcrová A., Pokorna S., Pullanchery S., Kohagen M., Jurkiewicz P., Hof M., Jungwirth P., Cremer P. S., Cwiklik L.. The complex nature of calcium cation interactions with phospholipid bilayers. Sci. Rep. 2016;6(1):38035. doi: 10.1038/srep38035. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Orriss I. R.. Extracellular pyrophosphate: The body’s ″water softener″. Bone. 2020;134:115243. doi: 10.1016/j.bone.2020.115243. [DOI] [PubMed] [Google Scholar]
- Briolay A., Bessueille L., Magne D.. TNAP: A New Multitask Enzyme in Energy Metabolism. Int. J. Mol. Sci. 2021;22(19):22. doi: 10.3390/ijms221910470. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mitton-Fitzgerald E., Gohr C. M., Bettendorf B., Rosenthal A. K.. The Role of ANK in Calcium Pyrophosphate Deposition Disease. Curr. Rheumatol. Rep. 2016;18(5):25. doi: 10.1007/s11926-016-0574-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Liu Y., Wu Y., Jiang M.. The emerging roles of PHOSPHO1 and its regulated phospholipid homeostasis in metabolic disorders. Front. Physiol. 2022;13:935195. doi: 10.3389/fphys.2022.935195. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Staines K. A., Myers K., Little K., Ralston S. H., Farquharson C.. Proton Pump Inhibitors Inhibit PHOSPHO1 Activity and Matrix Mineralisation In Vitro. Calcif. Tissue Int. 2021;109(6):696–705. doi: 10.1007/s00223-021-00882-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jun Y. S., Zhu Y., Wang Y., Ghim D., Wu X., Kim D., Jung H.. Classical and Nonclassical Nucleation and Growth Mechanisms for Nanoparticle Formation. Annu. Rev. Phys. Chem. 2022;73:453–477. doi: 10.1146/annurev-physchem-082720-100947. [DOI] [PubMed] [Google Scholar]
- Jahnen-Dechent W., Heiss A., Schäfer C., Ketteler M.. Fetuin-A regulation of calcified matrix metabolism. Circ. Res. 2011;108(12):1494–1509. doi: 10.1161/CIRCRESAHA.110.234260. [DOI] [PubMed] [Google Scholar]
- DiMeglio L. A., Imel E. A.. Calcium and phosphate: hormonal regulation and metabolism. Basic Appl. Bone Biol. 2019:257–282. doi: 10.1016/B978-0-12-813259-3.00013-0. [DOI] [Google Scholar]
- Campbell A. K.. Calcium as an intracellular regulator. Proc. Nutr. Soc. 1990;49(1):51–56. doi: 10.1079/PNS19900008. [DOI] [PubMed] [Google Scholar]
- George A., Veis A.. Phosphorylated proteins and control over apatite nucleation, crystal growth, and inhibition. Chem. Rev. 2008;108(11):4670–4693. doi: 10.1021/cr0782729. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boskey A. L., Villarreal-Ramirez E.. Intrinsically disordered proteins and biomineralization. Matrix Biol. 2016;52–54:43–59. doi: 10.1016/j.matbio.2016.01.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Staines K. A., MacRae V. E., Farquharson C.. The importance of the SIBLING family of proteins on skeletal mineralisation and bone remodelling. J. Endocrinol. 2012;214(3):241–255. doi: 10.1530/JOE-12-0143. [DOI] [PubMed] [Google Scholar]
- Habchi J., Tompa P., Longhi S., Uversky V. N.. Introducing protein intrinsic disorder. Chem. Rev. 2014;114(13):6561–6588. doi: 10.1021/cr400514h. [DOI] [PubMed] [Google Scholar]
- Holt C., Lenton S., Nylander T., Sørensen E. S., Teixeira S. C.. Mineralisation of soft and hard tissues and the stability of biofluids. J. Struct. Biol. 2014;185(3):383–396. doi: 10.1016/j.jsb.2013.11.009. [DOI] [PubMed] [Google Scholar]
- Guibert C., Landoulsi J.. Enzymatic Approach in Calcium Phosphate Biomineralization: A Contribution to Reconcile the Physicochemical with the Physiological View. Int. J. Mol. Sci. 2021;22(23):12957. doi: 10.3390/ijms222312957. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Leiendecker A., Witzleben S., Schulze M., Tobiasch E.. Template-Mediated Biomineralization for Bone Tissue Engineering. Curr. Stem Cell Res. Ther. 2016;12(2):103–123. doi: 10.2174/1574888X11666160217154436. [DOI] [PubMed] [Google Scholar]
- Schloßmacher U., Wiens M., Schröder H. C., Wang X., Jochum K. P., Müller W. E.. Silintaphin-1--interaction with silicatein during structure-guiding bio-silica formation. FEBS J. 2011;278(7):1145–1155. doi: 10.1111/j.1742-4658.2011.08040.x. [DOI] [PubMed] [Google Scholar]
- Zhao Y., Chen Z., Wang H., Wang J.. Crystallization Control of CaCO3 by Ionic Liquids in Aqueous Solution. Cryst. Growth Des. 2009;9(11):4984–4986. doi: 10.1021/cg900771c. [DOI] [Google Scholar]
- Xiong R., Wu W., Lu C., Cölfen H.. Bioinspired Chiral Template Guided Mineralization for Biophotonic Structural Materials (Adv. Mater. 51/2022) Adv. Mater. 2022;34(51):2270355. doi: 10.1002/adma.202270355. [DOI] [PubMed] [Google Scholar]
- Luan P. P., Jiang Y. J., Zhang S. P., Gao J., Su Z. G., Ma G. H., Zhang Y. F.. Chitosan-mediated formation of biomimetic silica nanoparticles: an effective method for manganese peroxidase immobilization and stabilization. J. Biosci. Bioeng. 2014;118(5):575–582. doi: 10.1016/j.jbiosc.2014.05.003. [DOI] [PubMed] [Google Scholar]
- Mehrali M., Shirazi S. F. S., Baradaran S., Mehrali M., Metselaar H. S., Kadri N. A., Osman N. A.. Facile synthesis of calcium silicate hydrate using sodium dodecyl sulfate as a surfactant assisted by ultrasonic irradiation. Ultrason. Sonochem. 2014;21(2):735–742. doi: 10.1016/j.ultsonch.2013.08.012. [DOI] [PubMed] [Google Scholar]
- Gopi D., Kanimozhi K., Bhuvaneshwari N., Indira J., Kavitha L.. Novel banana peel pectin mediated green route for the synthesis of hydroxyapatite nanoparticles and their spectral characterization. Spectrochim. Acta, Part A. 2014;118:589–597. doi: 10.1016/j.saa.2013.09.034. [DOI] [PubMed] [Google Scholar]
- Cao C. Y., Mei M. L., Li Q. L., Lo E. C., Chu C. H.. Methods for biomimetic remineralization of human dentine: a systematic review. Int. J. Mol. Sci. 2015;16(3):4615–4627. doi: 10.3390/ijms16034615. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhong B., Peng C., Wang G., Tian L., Cai Q., Cui F.. Contemporary research findings on dentine remineralization. J. Tissue Eng. Regener. Med. 2015;9(9):1004–1016. doi: 10.1002/term.1814. [DOI] [PubMed] [Google Scholar]
- Ben-Nissan, B. Advances in Calcium Phosphate Biomaterials; Springer, 2014. [Google Scholar]
- a Milan A. M., Sugars R. V., Embery G., Waddington R. J.. Adsorption and interactions of dentine phosphoprotein with hydroxyapatite and collagen. Eur. J. Oral Sci. 2006;114(3):223–231. doi: 10.1111/j.1600-0722.2006.00347.x. [DOI] [PubMed] [Google Scholar]; b Yarbrough D. K., Hagerman E., Eckert R., He J., Choi H., Cao N., Le K., Hedger J., Qi F., Anderson M.. et al. Specific binding and mineralization of calcified surfaces by small peptides. Calcif. Tissue Int. 2010;86(1):58–66. doi: 10.1007/s00223-009-9312-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Prasad M., Butler W. T., Qin C.. Dentin sialophosphoprotein in biomineralization. Connect. Tissue Res. 2010;51(5):404–417. doi: 10.3109/03008200903329789. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Q. Q., Miao L., Zhang H., Wang S. Q., Li Q., Sun W.. A novel amphiphilic oligopeptide induced the intrafibrillar mineralisation via interacting with collagen and minerals. J. Mater. Chem. B. 2020;8(11):2350–2362. doi: 10.1039/C9TB02928A. [DOI] [PubMed] [Google Scholar]
- Fan M., Zhang M., Xu H. H. K., Tao S., Yu Z., Yang J., Yuan H., Zhou X., Liang K., Li J.. Remineralization effectiveness of the PAMAM dendrimer with different terminal groups on artificial initial enamel caries in vitro. Dent. Mater. 2020;36(2):210–220. doi: 10.1016/j.dental.2019.11.015. [DOI] [PubMed] [Google Scholar]
- Chen L., Liang K., Li J., Wu D., Zhou X., Li J.. Regeneration of biomimetic hydroxyapatite on etched human enamel by anionic PAMAM template in vitro. Arch. Oral Biol. 2013;58(8):975–980. doi: 10.1016/j.archoralbio.2013.03.008. [DOI] [PubMed] [Google Scholar]
- Sharma V., Srinivasan A., Nikolajeff F., Kumar S.. Biomineralization process in hard tissues: The interaction complexity within protein and inorganic counterparts. Acta Biomater. 2021;120:20–37. doi: 10.1016/j.actbio.2020.04.049. [DOI] [PubMed] [Google Scholar]
- Wei Y., Liu S., Xiao Z., Zhao H., Luo J., Deng X., Guo L.. Enamel Repair with Amorphous Ceramics. Adv. Mater. 2020;32(7):1907067. doi: 10.1002/adma.201907067. [DOI] [PubMed] [Google Scholar]
- Prajapati S., Ruan Q., Mukherjee K., Nutt S., Moradian-Oldak J.. The Presence of MMP-20 Reinforces Biomimetic Enamel Regrowth. J. Dent. Res. 2018;97(1):84–90. doi: 10.1177/0022034517728504. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dogan S., Fong H., Yucesoy D. T., Cousin T., Gresswell C., Dag S., Huang G., Sarikaya M.. Biomimetic Tooth Repair: Amelogenin-Derived Peptide Enables in Vitro Remineralization of Human Enamel. ACS Biomater. Sci. Eng. 2018;4(5):1788–1796. doi: 10.1021/acsbiomaterials.7b00959. [DOI] [PubMed] [Google Scholar]
- Segman-Magidovich S., Grisaru H., Gitli T., Levi-Kalisman Y., Rapaport H.. Matrices of acidic β-sheet peptides as templates for calcium phosphate mineralization. Adv. Mater. 2008;20(11):2156–2161. doi: 10.1002/adma.200702125. [DOI] [Google Scholar]
- Kirkham J., Firth A., Vernals D., Boden N., Robinson C., Shore R. C., Brookes S. J., Aggeli A.. Self-assembling peptide scaffolds promote enamel remineralization. J. Dent. Res. 2007;86(5):426–430. doi: 10.1177/154405910708600507. [DOI] [PubMed] [Google Scholar]
- Philip N.. State of the Art Enamel Remineralization Systems: The Next Frontier in Caries Management. Caries Res. 2019;53(3):284–295. doi: 10.1159/000493031. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang J., Cao S., Li J., Xin J., Chen X., Wu W., Xu F., Li J.. Staged self-assembly of PAMAM dendrimers into macroscopic aggregates with a microribbon structure similar to that of amelogenin. Soft Matter. 2013;9(31):7553–7559. doi: 10.1039/c3sm51510a. [DOI] [Google Scholar]
- Shao C., Jin B., Mu Z., Lu H., Zhao Y., Wu Z., Yan L., Zhang Z., Zhou Y., Pan H.. et al. Repair of tooth enamel by a biomimetic mineralization frontier ensuring epitaxial growth. Sci. Adv. 2019;5(8):eaaw9569. doi: 10.1126/sciadv.aaw9569. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Liu Z., Shao C., Jin B., Zhang Z., Zhao Y., Xu X., Tang R.. Crosslinking ionic oligomers as conformable precursors to calcium carbonate. Nature. 2019;574(7778):394–398. doi: 10.1038/s41586-019-1645-x. [DOI] [PubMed] [Google Scholar]; b Mass T., Giuffre A. J., Sun C. Y., Stifler C. A., Frazier M. J., Neder M., Tamura N., Stan C. V., Marcus M. A., Gilbert P.. Amorphous calcium carbonate particles form coral skeletons. Proc. Natl. Acad. Sci. U.S.A. 2017;114(37):E7670–E7678. doi: 10.1073/pnas.1707890114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gordon L. M., Cohen M. J., MacRenaris K. W., Pasteris J. D., Seda T., Joester D.. Dental materials. Amorphous intergranular phases control the properties of rodent tooth enamel. Science. 2015;347(6223):746–750. doi: 10.1126/science.1258950. [DOI] [PubMed] [Google Scholar]
- de Melo Pereira D., Habibovic P.. Biomineralization-Inspired Material Design for Bone Regeneration. Adv. Healthcare Mater. 2018;7(22):1800700. doi: 10.1002/adhm.201800700. [DOI] [PubMed] [Google Scholar]
- Palmer L. C., Newcomb C. J., Kaltz S. R., Spoerke E. D., Stupp S. I.. Biomimetic systems for hydroxyapatite mineralization inspired by bone and enamel. Chem. Rev. 2008;108(11):4754–4783. doi: 10.1021/cr8004422. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lotsari A., Rajasekharan A. K., Halvarsson M., Andersson M.. Transformation of amorphous calcium Phosphate to bone-like apatite. Nat. Commun. 2018;9(1):4170. doi: 10.1038/s41467-018-06570-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang, X.-M. ; Qiu, Z.-Y. ; Cui, H. . Mineralized Collagen Bone Graft Substitutes; Woodhead Publishing, 2019. [Google Scholar]
- Erez J.. The source of ions for biomineralization in foraminifera and their implications for paleoceanographic proxies. Rev. Mineral. Geochem. 2003;54(1):115–149. doi: 10.2113/0540115. [DOI] [Google Scholar]
- Wang G., Li X., Mo L., Song Z., Chen W., Deng Y., Zhao H., Qin E., Qin C., Tang R.. Eggshell-inspired biomineralization generates vaccines that do not require refrigeration. Angew. Chem., Int. Ed. 2012;51(42):10576–10579. doi: 10.1002/anie.201206154. [DOI] [PubMed] [Google Scholar]
- Ma X., Chen H., Yang L., Wang K., Guo Y., Yuan L.. Construction and potential applications of a functionalized cell with an intracellular mineral scaffold. Angew. Chem., Int. Ed. 2011;50(32):7414–7417. doi: 10.1002/anie.201100126. [DOI] [PubMed] [Google Scholar]
- Yu Y., Guo Z., Zhao Y., Kong K., Pan H., Xu X., Tang R., Liu Z.. A Flexible and Degradable Hybrid Mineral as a Plastic Substitute. Adv. Mater. 2022;34(9):2107523. doi: 10.1002/adma.202107523. [DOI] [PubMed] [Google Scholar]
- Wang B., Liu P., Liu Z., Pan H., Xu X., Tang R.. Biomimetic construction of cellular shell by adjusting the interfacial energy. Biotechnol. Bioeng. 2014;111(2):386–395. doi: 10.1002/bit.25016. [DOI] [PubMed] [Google Scholar]
- Zhang Z. B., Gao H. L., Wen S. M., Pang J., Zhang S. C., Cui C., Wang Z. Y., Yu S. H.. Scalable Manufacturing of Mechanical Robust Bioinspired Ceramic-Resin Composites with Locally Tunable Heterogeneous Structures. Adv. Mater. 2023;35(14):2209510. doi: 10.1002/adma.202209510. [DOI] [PubMed] [Google Scholar]
- Liu X., Zhang F., Jing X., Pan M., Liu P., Li W., Zhu B., Li J., Chen H., Wang L.. et al. Complex silica composite nanomaterials templated with DNA origami. Nature. 2018;559(7715):593–598. doi: 10.1038/s41586-018-0332-7. [DOI] [PubMed] [Google Scholar]
- Wassermann L. M., Scheckenbach M., Baptist A. V., Glembockyte V., Heuer-Jungemann A.. Full Site-Specific Addressability in DNA Origami-Templated Silica Nanostructures. Adv. Mater. 2023;35(23):2212024. doi: 10.1002/adma.202212024. [DOI] [PubMed] [Google Scholar]
- Lv H., Xie N., Li M., Dong M., Sun C., Zhang Q., Zhao L., Li J., Zuo X., Chen H.. et al. DNA-based programmable gate arrays for general-purpose DNA computing. Nature. 2023;622(7982):292–300. doi: 10.1038/s41586-023-06484-9. [DOI] [PubMed] [Google Scholar]
- Danesi A. L., Athanasiadou D., Aderinto A. O., Rasie P., Chou L. Y. T., Carneiro K. M. M.. Peptide-Decorated DNA Nanostructures Promote Site-Specific Hydroxyapatite Growth. ACS Appl. Mater. Interfaces. 2022;14(1):1692–1698. doi: 10.1021/acsami.1c19271. [DOI] [PubMed] [Google Scholar]
- He Z., Shi K., Li J., Chao J.. Self-assembly of DNA origami for nanofabrication, biosensing, drug delivery, and computational storage. iScience. 2023;26(5):106638. doi: 10.1016/j.isci.2023.106638. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lei C., Wang K. Y., Ma Y. X., Hao D. X., Zhu Y. N., Wan Q. Q., Zhang J. S., Tay F. R., Mu Z., Niu L. N.. Biomimetic Self-Maturation Mineralization System for Enamel Repair. Adv. Mater. 2024;36(16):2311659. doi: 10.1002/adma.202311659. [DOI] [PubMed] [Google Scholar]
- Yao S., Jin B., Liu Z., Shao C., Zhao R., Wang X., Tang R.. Biomineralization: From Material Tactics to Biological Strategy. Adv. Mater. 2017;29(14):1605903. doi: 10.1002/adma.201605903. [DOI] [PubMed] [Google Scholar]
- Zhao C., Toufigh V., Zhang J., Liu Y., Fan W., He X., Cao B., Xiao Y.. Enhancing biomineralization process efficiency with trained bacterial strains: A technical perspective. Biogeotechnics. 2023;1(2):100017. doi: 10.1016/j.bgtech.2023.100017. [DOI] [Google Scholar]
- Özcan S., Çiftçioğlu M.. Enhanced model protein adsorption of nanoparticulate hydroxyapatite thin films on silk sericin and fibroin surfaces. J. Mater. Sci.: Mater. Med. 2022;33(1):6. doi: 10.1007/s10856-021-06632-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guo Y., Xu Z., Liu M., Zu S., Yang Y., Wang Q., Yu Z., Zhang Z., Ren L.. The corrosion resistance, biocompatibility and biomineralization of the dicalcium phosphate dihydrate coating on the surface of the additively manufactured NiTi alloy. J. Mater. Res. Technol. 2022;17:622–635. doi: 10.1016/j.jmrt.2022.01.063. [DOI] [Google Scholar]
- Cao S., Ivanov T., Heuer J., Ferguson C. T. J., Landfester K., Caire da Silva L.. Dipeptide coacervates as artificial membraneless organelles for bioorthogonal catalysis. Nat. Commun. 2024;15(1):39. doi: 10.1038/s41467-023-44278-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dai J., Wang L., Wang Y., Tian S., Tian X., Xie A., Zhang R., Yan Y., Pan J.. Robust Nacrelike Graphene Oxide-Calcium Carbonate Hybrid Mesh with Underwater Superoleophobic Property for Highly Efficient Oil/Water Separation. ACS Appl. Mater. Interfaces. 2020;12(4):4482–4493. doi: 10.1021/acsami.9b18664. [DOI] [PubMed] [Google Scholar]
- Chen W., Zhang P., Yu S., Zang R., Xu L., Wang S., Wang B., Meng J.. Nacre-inspired underwater superoleophobic films with high transparency and mechanical robustness. Nat. Protoc. 2022;17(11):2647–2667. doi: 10.1038/s41596-022-00725-3. [DOI] [PubMed] [Google Scholar]
- a Andersson M., Jia Q., Abella A., Lee X. Y., Landreh M., Purhonen P., Hebert H., Tenje M., Robinson C. V., Meng Q.. et al. Biomimetic spinning of artificial spider silk from a chimeric minispidroin. Nat. Chem. Biol. 2017;13(3):262–264. doi: 10.1038/nchembio.2269. [DOI] [PubMed] [Google Scholar]; b Chen W. C., Wang R. C., Yu S. K., Chen J. L., Kao Y. H., Wang T. Y., Chang P. Y., Sheu H. S., Chen S. C., Liu W. R.. Self-Healable Spider Dragline Silk Materials. Adv. Funct. Mater. 2023;33(44):2303571. doi: 10.1002/adfm.202303571. [DOI] [Google Scholar]
- Arndt T., Chatterjee U., Shilkova O., Francis J., Lundkvist J., Johansson D., Schmuck B., Greco G., Nordberg Å. E., Li Y.. et al. Tuneable Recombinant Spider Silk Protein Hydrogels for Drug Release and 3D Cell Culture. Adv. Funct. Mater. 2024;34(35):2303622. doi: 10.1002/adfm.202303622. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Anton A. M., Heidebrecht A., Mahmood N., Beiner M., Scheibel T., Kremer F.. Foundation of the Outstanding Toughness in Biomimetic and Natural Spider Silk. Biomacromolecules. 2017;18(12):3954–3962. doi: 10.1021/acs.biomac.7b00990. [DOI] [PubMed] [Google Scholar]
- Yu Y., He Y., Mu Z., Zhao Y., Kong K., Liu Z., Tang R.. Biomimetic mineralized organic–inorganic hybrid macrofiber with spider silk-like supertoughness. Adv. Funct. Mater. 2020;30(6):1908556. doi: 10.1002/adfm.201908556. [DOI] [Google Scholar]
- Wan Q., Yang M., Hu J., Lei F., Shuai Y., Wang J., Holland C., Rodenburg C., Yang M.. Mesoscale structure development reveals when a silkworm silk is spun. Nat. Commun. 2021;12(1):3711. doi: 10.1038/s41467-021-23960-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu S.-D., Chuang W.-T., Ho J.-C., Wu H.-C., Hsu S.-h.. Self-healing of recombinant spider silk gel and coating. Polymers. 2023;15(8):1855. doi: 10.3390/polym15081855. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Koga T., Morishita T., Harumoto Y., Nishimura S.-n., Higashi N.. Spider silk-inspired peptide multiblock hybrid copolymers for self-healable thin film materials. Mater. Adv. 2021;2(24):7851–7860. doi: 10.1039/D1MA00823D. [DOI] [Google Scholar]
- Lyu R., Chen Y., Shuai Y., Wang J., Lu L., Cheng Q., Cai J., Mao C., Yang M.. Novel biomaterial-binding/osteogenic bi-functional peptide binds to silk fibroin membranes to effectively induce osteogenesis in vitro and in vivo. ACS Appl. Mater. Interfaces. 2023;15(6):7673–7685. doi: 10.1021/acsami.2c17554. [DOI] [PubMed] [Google Scholar]
- Li C., Wu J., Shi H., Xia Z., Sahoo J. K., Yeo J., Kaplan D. L.. Fiber-based biopolymer processing as a route toward sustainability. Adv. Mater. 2022;34(1):2105196. doi: 10.1002/adma.202105196. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Veiga A., Castro F., L Oliveira A., Rocha F.. High efficient strategy for the production of hydroxyapatite/silk sericin nanocomposites. J. Chem. Technol. Biotechnol. 2021;96(1):241–248. doi: 10.1002/jctb.6532. [DOI] [Google Scholar]
- Cheng C., Shao Z., Vollrath F.. Silk fibroin-regulated crystallization of calcium carbonate. Adv. Funct. Mater. 2008;18(15):2172–2179. doi: 10.1002/adfm.200701130. [DOI] [Google Scholar]
- Zhang X., Fan Z., Lu Q., Huang Y., Kaplan D. L., Zhu H.. Hierarchical biomineralization of calcium carbonate regulated by silk microspheres. Acta Biomater. 2013;9(6):6974–6980. doi: 10.1016/j.actbio.2013.03.004. [DOI] [PubMed] [Google Scholar]
- Weiss I. M., Tuross N., Addadi L., Weiner S.. Mollusc larval shell formation: amorphous calcium carbonate is a precursor phase for aragonite. J. Exp. Zool. 2002;293(5):478–491. doi: 10.1002/jez.90004. [DOI] [PubMed] [Google Scholar]
- Cheng M., Liu M., Chang L., Liu Q., Wang C., Hu L., Zhang Z., Ding W., Chen L., Guo S.. et al. Overview of structure, function and integrated utilization of marine shell. Sci. Total Environ. 2023;870:161950. doi: 10.1016/j.scitotenv.2023.161950. [DOI] [PubMed] [Google Scholar]
- Rodrigues J. R., Alves N. M., Mano J. F.. Nacre-inspired nanocomposites produced using layer-by-layer assembly: Design strategies and biomedical applications. Mater. Sci. Eng.: C. 2017;76:1263–1273. doi: 10.1016/j.msec.2017.02.043. [DOI] [PubMed] [Google Scholar]
- Yu Y., Kong K., Tang R., Liu Z.. A bioinspired ultratough composite produced by integration of inorganic ionic oligomers within polymer networks. ACS Nano. 2022;16(5):7926–7936. doi: 10.1021/acsnano.2c00663. [DOI] [PubMed] [Google Scholar]
- Liang K., Zhao C., Song C., Zhao L., Qiu P., Wang S., Zhu J., Gong Z., Liu Z., Tang R.. et al. In Situ Biomimetic Mineralization of Bone-Like Hydroxyapatite in Hydrogel for the Acceleration of Bone Regeneration. ACS Appl. Mater. Interfaces. 2023;15(1):292–308. doi: 10.1021/acsami.2c16217. [DOI] [PubMed] [Google Scholar]
- Chen K., Huang J., Ding L., Yao Y., Wei Y., Zhong H., Yang X., Liu Z., Deng X., Guo L.. Preparation, microstructure, and mechanical properties of enamel-like nanocomposites for dental repair application. Innovation Mater. 2025;3(1):100111. doi: 10.59717/j.xinn-mater.2024.100111. [DOI] [Google Scholar]
- a Olszta M. J., Cheng X., Jee S. S., Kumar R., Kim Y.-Y., Kaufman M. J., Douglas E. P., Gower L. B.. Bone structure and formation: A new perspective. Mater. Sci. Eng., R. 2007;58(3):77–116. doi: 10.1016/j.mser.2007.05.001. [DOI] [Google Scholar]; b Marelli B., Ghezzi C. E., Alessandrino A., Barralet J. E., Freddi G., Nazhat S. N.. Silk fibroin derived polypeptide-induced biomineralization of collagen. Biomaterials. 2012;33(1):102–108. doi: 10.1016/j.biomaterials.2011.09.039. [DOI] [PubMed] [Google Scholar]
- a Yang M., Shuai Y., Zhang C., Chen Y., Zhu L., Mao C., OuYang H.. Biomimetic Nucleation of Hydroxyapatite Crystals Mediated by Antheraea pernyi Silk Sericin Promotes Osteogenic Differentiation of Human Bone Marrow Derived Mesenchymal Stem Cells. Biomacromolecules. 2014;15(4):1185–1193. doi: 10.1021/bm401740x. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Xu Z., Shi L., Hu D., Hu B., Yang M., Zhu L.. Formation of hierarchical bone-like apatites on silk microfiber templates via biomineralization. RSC Adv. 2016;6(80):76426–76433. doi: 10.1039/C6RA17199K. [DOI] [Google Scholar]
- Miller J.. Advancing Bioenergetic Research in Biomineralization with Antibody Functionalized AFM Imaging. J. Bio Energ. 2024;12(4):1000281 [Google Scholar]
- a Maleki H., Shahbazi M. A., Montes S., Hosseini S. H., Eskandari M. R., Zaunschirm S., Verwanger T., Mathur S., Milow B., Krammer B., Hüsing N.. Mechanically Strong Silica-Silk Fibroin Bioaerogel: A Hybrid Scaffold with Ordered Honeycomb Micromorphology and Multiscale Porosity for Bone Regeneration. ACS Appl. Mater. Interfaces. 2019;11(19):17256–17269. doi: 10.1021/acsami.9b04283. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Maleki H., Montes S., Hayati-Roodbari N., Putz F., Huesing N.. Compressible, Thermally Insulating, and Fire Retardant Aerogels through Self-Assembling Silk Fibroin Biopolymers Inside a Silica Structure-An Approach towards 3D Printing of Aerogels. ACS Appl. Mater. Interfaces. 2018;10(26):22718–22730. doi: 10.1021/acsami.8b05856. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yu C., Ying X., Shahbazi M. A., Yang L., Ma Z., Ye L., Yang W., Sun R., Gu T., Tang R.. et al. A nano-conductive osteogenic hydrogel to locally promote calcium influx for electro-inspired bone defect regeneration. Biomaterials. 2023;301:122266. doi: 10.1016/j.biomaterials.2023.122266. [DOI] [PubMed] [Google Scholar]
- a Matrajt L., Eaton J., Leung T., Dimitrov D., Schiffer J. T., Swan D. A., Janes H.. Optimizing vaccine allocation for COVID-19 vaccines shows the potential role of single-dose vaccination. Nat. Commun. 2021;12(1):3449. doi: 10.1038/s41467-021-23761-1. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Ye Z., Harmon J., Ni W., Li Y., Wich D., Xu Q.. The mRNA Vaccine Revolution: COVID-19 Has Launched the Future of Vaccinology. ACS Nano. 2023;17(16):15231–15253. doi: 10.1021/acsnano.2c12584. [DOI] [PubMed] [Google Scholar]; c Yeh M. T., Smith M., Carlyle S., Konopka-Anstadt J. L., Burns C. C., Konz J., Andino R., Macadam A.. Genetic stabilization of attenuated oral vaccines against poliovirus types 1 and 3. Nature. 2023;619(7968):135–142. doi: 10.1038/s41586-023-06212-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen X., Fernando G. J., Raphael A. P., Yukiko S. R., Fairmaid E. J., Primiero C. A., Frazer I. H., Brown L. E., Kendall M. A.. Rapid kinetics to peak serum antibodies is achieved following influenza vaccination by dry-coated densely packed microprojections to skin. J. Controlled Release. 2012;158(1):78–84. doi: 10.1016/j.jconrel.2011.10.026. [DOI] [PubMed] [Google Scholar]
- Ai L., Li Y., Zhou L., Yao W., Zhang H., Hu Z., Han J., Wang W., Wu J., Xu P.. et al. Lyophilized mRNA-lipid nanoparticle vaccines with long-term stability and high antigenicity against SARS-CoV-2. Cell Discovery. 2023;9(1):9. doi: 10.1038/s41421-022-00517-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang X., Sun C., Li P., Wu T., Zhou H., Yang D., Liu Y., Ma X., Song Z., Nian Q.. et al. Vaccine Engineering with Dual-Functional Mineral Shell: A Promising Strategy to Overcome Preexisting Immunity. Adv. Mater. 2016;28(4):694–700. doi: 10.1002/adma.201503740. [DOI] [PubMed] [Google Scholar]
- Wang X., Yang D., Li S., Xu X., Qin C. F., Tang R.. Biomineralized vaccine nanohybrid for needle-free intranasal immunization. Biomaterials. 2016;106:286–294. doi: 10.1016/j.biomaterials.2016.08.035. [DOI] [PubMed] [Google Scholar]
- Hao H., Wu S., Lin J., Zheng Z., Zhou Y., Zhang Y., Guo Q., Tian F., Zhao M., Chen Y.. et al. Immunization against Zika by entrapping live virus in a subcutaneous self-adjuvanting hydrogel. Nat. Biomed. Eng. 2023;7(7):928–942. doi: 10.1038/s41551-023-01014-4. [DOI] [PubMed] [Google Scholar]
- Qiao Q., Li X., Ou X., Liu X., Fu C., Wang Y., Niu B., Kong L., Yang C., Zhang Z.. Hybrid biomineralized nanovesicles to enhance inflamed lung biodistribution and reduce side effect of glucocorticoid for ARDS therapy. J. Controlled Release. 2024;369:746–764. doi: 10.1016/j.jconrel.2024.04.015. [DOI] [PubMed] [Google Scholar]
- Wang W., Liu X., Zheng X., Jin H. J., Li X.. Biomineralization: An Opportunity and Challenge of Nanoparticle Drug Delivery Systems for Cancer Therapy. Adv. Healthcare Mater. 2020;9(22):2001117. doi: 10.1002/adhm.202001117. [DOI] [PubMed] [Google Scholar]
- a Nong Y., Ren Y., Wang P., Zhou M., Yu Y., Yuan J., Xu B., Wang Q.. A facile strategy for the preparation of photothermal silk fibroin aerogels with antibacterial and oil-water separation abilities. J. Colloid Interface Sci. 2021;603:518–529. doi: 10.1016/j.jcis.2021.06.134. [DOI] [PubMed] [Google Scholar]; b Xue J., Gao H. L., Wang X. Y., Qian K. Y., Yang Y., He T., He C., Lu Y., Yu S. H.. Bioinspired Unidirectional Silk Fibroin–Silver Compound Nanowire Composite Scaffold via Interface-Mediated In Situ Synthesis. Angew. Chem. 2019;131(40):14290–14294. doi: 10.1002/ange.201907708. [DOI] [PubMed] [Google Scholar]
- a Potiyaraj P., Kumlangdudsana P., Dubas S. T.. Synthesis of silver chloride nanocrystal on silk fibers. Mater. Lett. 2007;61(11–12):2464–2466. doi: 10.1016/j.matlet.2006.09.039. [DOI] [Google Scholar]; b Su H., Han J., Dong Q., Xu J., Chen Y., Gu Y., Song W., Zhang D.. In situ bioinspired synthesis of silver chloride nanocrystals on silk fibroin fibers. Appl. Phys. A. 2011;102:429–434. doi: 10.1007/s00339-010-6137-7. [DOI] [Google Scholar]
- Findik M., Bingol H., Erdem A.. Electrochemical detection of interaction between daunorubicin and DNA by hybrid nanoflowers modified graphite electrodes. Sens. Actuators, B. 2021;329:129120. doi: 10.1016/j.snb.2020.129120. [DOI] [PubMed] [Google Scholar]
- Wang J., Zhang Y., Jin N., Mao C., Yang M.. Protein-induced gold nanoparticle assembly for improving the photothermal effect in cancer therapy. ACS Appl. Mater. Interfaces. 2019;11(12):11136–11143. doi: 10.1021/acsami.8b21488. [DOI] [PubMed] [Google Scholar]
- Shuai Y., Yang S., Li C., Zhu L., Mao C., Yang M.. In situ protein-templated porous protein–hydroxylapatite nanocomposite microspheres for pH-dependent sustained anticancer drug release. J. Mater. Chem. B. 2017;5(21):3945–3954. doi: 10.1039/C7TB00208D. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Yang R., Fu S., Li R., Zhang L., Xu Z., Cao Y., Cui H., Kang Y., Xue P.. Facile engineering of silk fibroin capped AuPt bimetallic nanozyme responsive to tumor microenvironmental factors for enhanced nanocatalytic therapy. Theranostics. 2021;11(1):107. doi: 10.7150/thno.50486. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Yu H., He M., Li Y., Liu Y., Xu Z., Zhang L., Kang Y., Xue P.. A robust Au@ Cu2-xS nanoreactor assembled by silk fibroin for enhanced intratumoral glucose depletion and redox dyshomeostasis. Biomaterials. 2023;293:121970. doi: 10.1016/j.biomaterials.2022.121970. [DOI] [PubMed] [Google Scholar]
- Yang R., Hou M., Gao Y., Lu S., Zhang L., Xu Z., Li C. M., Kang Y., Xue P.. Biomineralization-inspired crystallization of manganese oxide on silk fibroin nanoparticles for in vivo MR/fluorescence imaging-assisted tri-modal therapy of cancer. Theranostics. 2019;9(21):6314–6333. doi: 10.7150/thno.36252. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhao R., Wang B., Yang X., Xiao Y., Wang X., Shao C., Tang R.. A Drug-Free Tumor Therapy Strategy: Cancer-Cell-Targeting Calcification. Angew. Chem., Int. Ed. 2016;55(17):5225–5229. doi: 10.1002/anie.201601364. [DOI] [PubMed] [Google Scholar]
- Tang N., Li H., Zhang L., Zhang X., Chen Y., Shou H., Feng S., Chen X., Luo Y., Tang R., Wang B.. A Macromolecular Drug for Cancer Therapy via Extracellular Calcification. Angew. Chem., Int. Ed. 2021;60(12):6509–6517. doi: 10.1002/anie.202016122. [DOI] [PubMed] [Google Scholar]
- Panwar V., Dutta T.. Diatom Biogenic Silica as a Felicitous Platform for Biochemical Engineering: Expanding Frontiers. ACS Appl. Bio Mater. 2019;2(6):2295–2316. doi: 10.1021/acsabm.9b00050. [DOI] [PubMed] [Google Scholar]
- Rea I., De Stefano L.. Recent Advances on Diatom-Based Biosensors. Sensors. 2019;19(23):5208. doi: 10.3390/s19235208. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim J.-N., Lee J., Lee H., Oh I.. Stretchable and self-healable catechol-chitosan-diatom hydrogel for triboelectric generator and self-powered tremor sensor targeting at Parkinson disease. Nano Energy. 2021;82:105705. doi: 10.1016/j.nanoen.2020.105705. [DOI] [Google Scholar]
- Buiculescu R., Chaniotakis N. A.. The stabilization of Au NP-AChE nanocomposites by biosilica encapsulation for the development of a thiocholine biosensor. Bioelectrochemistry. 2012;86:72–77. doi: 10.1016/j.bioelechem.2012.02.005. [DOI] [PubMed] [Google Scholar]
- Choi O., Kim B. C., An J. H., Min K., Kim Y. H., Um Y., Oh M. K., Sang B. I.. A biosensor based on the self-entrapment of glucose oxidase within biomimetic silica nanoparticles induced by a fusion enzyme. Enzyme Microb. Technol. 2011;49(5):441–445. doi: 10.1016/j.enzmictec.2011.07.005. [DOI] [PubMed] [Google Scholar]
- Lei Z., Wang Q., Sun S., Zhu W., Wu P.. A Bioinspired Mineral Hydrogel as a Self-Healable, Mechanically Adaptable Ionic Skin for Highly Sensitive Pressure Sensing. Adv. Mater. 2017;29(22):1700321. doi: 10.1002/adma.201700321. [DOI] [PubMed] [Google Scholar]
- Liu J.-h., Mao Z., Chen Y., Long Y., Wu H., Shen J., Zhang R., Yeung O. W. H., Zhou B., Zhi C.. et al. Amorphous biomineral-reinforced hydrogels with dramatically enhanced toughness for strain sensing. Chem. Eng. J. 2023;468:143735. doi: 10.1016/j.cej.2023.143735. [DOI] [Google Scholar]
- Dieudonné A., Prévéral S., Pignol D.. A Sensitive Magnetic Arsenite-Specific Biosensor Hosted in Magnetotactic Bacteria. Appl. Environ. Microbiol. 2020;86(14):e00803-20. doi: 10.1128/AEM.00803-20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu S., Tian J., Xue X., Ma F., Li Q. X., Morisseau C., Hammock B. D., Xu T.. Biosynthesis of magnetosome-nanobody complex in Magnetospirillum gryphiswaldense MSR-1 and a magnetosome-nanobody-based enzyme-linked immunosorbent assay for the detection of tetrabromobisphenol A in water. Anal. Bioanal. Chem. 2024;416(1):141–149. doi: 10.1007/s00216-023-05005-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu S., Ma F., He J., Li Q. X., Hammock B. D., Tian J., Xu T.. Fusion expression of nanobodies specific for the insecticide fipronil on magnetosomes in Magnetospirillum gryphiswaldense MSR-1. J. Nanobiotechnol. 2021;19(1):27. doi: 10.1186/s12951-021-00773-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Sannigrahi S., Arumugasamy S. K., Mathiyarasu J., Sudhakaran R., Suthindhiran K.. Detection of white spot syndrome virus in seafood samples using a magnetosome-based impedimetric biosensor. Arch. Virol. 2021;166(10):2763–2778. doi: 10.1007/s00705-021-05187-8. [DOI] [PubMed] [Google Scholar]; b Wu L., Gao B., Zhang F., Sun X., Zhang Y., Li Z.. A novel electrochemical immunosensor based on magnetosomes for detection of staphylococcal enterotoxin B in milk. Talanta. 2013;106:360–366. doi: 10.1016/j.talanta.2012.12.053. [DOI] [PubMed] [Google Scholar]
- Hartati Y. W., Irkham I., Zulqaidah S., Syafira R. S., Kurnia I., Noviyanti A. R., Topkaya S. N.. Recent advances in hydroxyapatite-based electrochemical biosensors: Applications and future perspectives. Sens. Bio-Sens. Res. 2022;38:100542. doi: 10.1016/j.sbsr.2022.100542. [DOI] [Google Scholar]
- Jiang W., Tian D., Zhang L., Guo Q., Cui Y., Yang M.. Dual signal amplification strategy for amperometric aptasensing using hydroxyapatite nanoparticles. Application to the sensitive detection of the cancer biomarker platelet-derived growth factor BB. Microchim. Acta. 2017;184:4375–4381. doi: 10.1007/s00604-017-2471-1. [DOI] [Google Scholar]
- Chen Z., Ai Y., Yang W., Niu L., Wang Q., Gao F.. Specific adsorption of magnetic Fe3O4@ hydroxyapatite nanocomposites to a hybridized duplex for immobilization and label-free electrochemical biosensing of MicroRNA. ACS Appl. Nano Mater. 2023;6(10):8584–8592. doi: 10.1021/acsanm.3c00951. [DOI] [Google Scholar]
- Syafira R. S., Devi M. J., Gaffar S., Irkham, Kurnia I., Arnafia W., Einaga Y., Syakir N., Noviyanti A. R., Hartati Y. W.. Hydroxyapatite-Gold Modified Screen-Printed Carbon Electrode for Selective SARS-CoV-2 Antibody Immunosensor. ACS Appl. Bio Mater. 2024;7(2):950–960. doi: 10.1021/acsabm.3c00953. [DOI] [PubMed] [Google Scholar]
- He Y., Lv C., Hou X., Wu L.. Mono-dispersed nano-hydroxyapatite based MRI probe with tetrahedral DNA nanostructures modification for in vitro tumor cell imaging. Anal. Chim. Acta. 2020;1138:141–149. doi: 10.1016/j.aca.2020.09.006. [DOI] [PubMed] [Google Scholar]
- Qin J., Rosen B. P., Zhang Y., Wang G., Franke S., Rensing C.. Arsenic detoxification and evolution of trimethylarsine gas by a microbial arsenite S-adenosylmethionine methyltransferase. Proc. Natl. Acad. Sci. U.S.A. 2006;103(7):2075–2080. doi: 10.1073/pnas.0506836103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vera-Espíndola F., Jeison D., Gentina J. C., Muñoz J., González E.. Reviewing arsenic biomineralization: An upcoming strategy for mining wastewater treatment. Sci. Total Environ. 2024;954:176538. doi: 10.1016/j.scitotenv.2024.176538. [DOI] [PubMed] [Google Scholar]
- Xiong W., Zhao X., Zhu G., Shao C., Li Y., Ma W., Xu X., Tang R.. Silicification-induced cell aggregation for the sustainable production of H2 under aerobic conditions. Angew. Chem. 2015;127(41):12129–12133. doi: 10.1002/ange.201504634. [DOI] [PubMed] [Google Scholar]
- Chen P., Liu X., Gu C., Zhong P., Song N., Li M., Dai Z., Fang X., Liu Z., Zhang J.. et al. A plant-derived natural photosynthetic system for improving cell anabolism. Nature. 2022;612(7940):546–554. doi: 10.1038/s41586-022-05499-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Tang T. D., Huang X.. Artificial Cells: From Basic Assembly to Directed Functionality. Small Methods. 2023;7(12):2301446. doi: 10.1002/smtd.202301446. [DOI] [PubMed] [Google Scholar]; b Einfalt T., Witzigmann D., Edlinger C., Sieber S., Goers R., Najer A., Spulber M., Onaca-Fischer O., Huwyler J., Palivan C. G.. Biomimetic artificial organelles with in vitro and in vivo activity triggered by reduction in microenvironment. Nat. Commun. 2018;9(1):1127. doi: 10.1038/s41467-018-03560-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu W., Lu Z., Lu C., Sun X., Ni B., Cölfen H., Xiong R.. Bioinspired Stabilization of Amorphous Calcium Carbonate by Carboxylated Nanocellulose Enables Mechanically Robust, Healable, and Sensing Biocomposites. ACS Nano. 2023;17(7):6664–6674. doi: 10.1021/acsnano.2c12385. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu Z., Xu X., Tang R.. Improvement of biological organisms using functional material shells. Adv. Funct. Mater. 2016;26(12):1862–1880. doi: 10.1002/adfm.201504480. [DOI] [Google Scholar]
- a Introduction: 3D Printing for Biomaterials Chem. Rev. 2020, 120 19 10545 10546. [DOI] [PubMed] [Google Scholar]; b Pati F., Gantelius J., Svahn H. A.. 3D Bioprinting of Tissue/Organ Models. Angew. Chem., Int. Ed. 2016;55(15):4650–4665. doi: 10.1002/anie.201505062. [DOI] [PubMed] [Google Scholar]; c Wang L., Li D., Huang Y., Mao R., Zhang B., Luo F., Gu P., Song P., Ge X., Lu J.. et al. Bionic Mineralized 3D-Printed Scaffolds with Enhanced In Situ Mineralization for Cranial Bone Regeneration. Adv. Funct. Mater. 2024;34(3):2309042. doi: 10.1002/adfm.202309042. [DOI] [Google Scholar]
- Pati F., Gantelius J., Svahn H. A.. 3D bioprinting of tissue/organ models. Angew. Chem., Int. Ed. 2016;55(15):4650–4665. doi: 10.1002/anie.201505062. [DOI] [PubMed] [Google Scholar]
- Shi Y., Shi J., Sun Y., Liu Q., Zhang C., Shao C., Yu K., Ge M., Mi R., Gu J.. et al. A Hierarchical 3D Graft Printed with Nanoink for Functional Craniofacial Bone Restoration. Adv. Funct. Mater. 2023;33(40):2301099. doi: 10.1002/adfm.202301099. [DOI] [Google Scholar]
- Mallick S., Pradhan T., Das S.. Bacterial biomineralization of heavy metals and its influencing factors for metal bioremediation. J. Environ. Manage. 2025;373:123977. doi: 10.1016/j.jenvman.2024.123977. [DOI] [PubMed] [Google Scholar]
- Castro-Alonso M. J., Montañez-Hernandez L. E., Sanchez-Muñoz M. A., Franco M. R. M., Narayanasamy R., Balagurusamy N.. Microbially Induced Calcium Carbonate Precipitation (MICP) and Its Potential in Bioconcrete: Microbiological and Molecular Concepts. Front. Mater. 2019;6:00126. doi: 10.3389/fmats.2019.00126. [DOI] [Google Scholar]
- Jonkers H. M., Schlangen E.. Development of a bacteria-based self healing concrete. Tailor-Made Concr. Struct. 2008;1:425–430. [Google Scholar]
- De Muynck W., Verbeken K., De Belie N., Verstraete W.. Influence of temperature on the effectiveness of a biogenic carbonate surface treatment for limestone conservation. Appl. Microbiol. Biotechnol. 2013;97:1335–1347. doi: 10.1007/s00253-012-3997-0. [DOI] [PubMed] [Google Scholar]
- Kalhori H., Bagherpour R.. Application of carbonate precipitating bacteria for improving properties and repairing cracks of shotcrete. Constr. Build. Mater. 2017;148:249–260. doi: 10.1016/j.conbuildmat.2017.05.074. [DOI] [Google Scholar]
- Bang S. S., Lippert J. J., Yerra U., Mulukutla S., Ramakrishnan V.. Microbial calcite, a bio-based smart nanomaterial in concrete remediation. Int. J. Smart Nano Mater. 2010;1(1):28–39. doi: 10.1080/19475411003593451. [DOI] [Google Scholar]
- Luo M., Qian C.-x., Li R.-y.. Factors affecting crack repairing capacity of bacteria-based self-healing concrete. Constr. Build. Mater. 2015;87:1–7. doi: 10.1016/j.conbuildmat.2015.03.117. [DOI] [Google Scholar]
- Wang X., Zhang L., Liu Z., Zeng Q., Jiang G., Yang M.. Probing the surface structure of hydroxyapatite through its interaction with hydroxyl: a first-principles study. RSC Adv. 2018;8(7):3716–3722. doi: 10.1039/C7RA13121F. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Su Y., Jin P.. Application of encapsulated expanded vermiculites as carriers of microorganisms and nutrients in self-repairing concrete. Biochem. Eng. J. 2022;187(2022):108672. doi: 10.1016/j.bej.2022.108672. [DOI] [Google Scholar]
- Lopez-Ayuso C. A., Aranda-Herrera B., Guzman-Rocha D., Chavez-Granados P. A., Garcia-Contreras R.. Synthetic Proteins in Dental Applications. SynBio. 2024;2(1):1–20. doi: 10.3390/synbio2010001. [DOI] [Google Scholar]





