Download PDF
Review  |  Open Access  |  6 Jul 2023

Physiological and pathological/ectopic mineralization: from composition to microstructure

Views: 600 |  Downloads: 95 |  Cited:   2
Microstructures 2023;3:2023030.
10.20517/microstructures.2023.05 |  © The Author(s) 2023.
Author Information
Article Notes
Cite This Article

Abstract

Biomineralization is a process that leads to the formation of hierarchically arranged structures in mineralized tissues, such as bone and teeth. Extensive research has been conducted on the crystals in bones and teeth, with the aim of understanding the underlying mechanisms of the mineralization process. Pathological/ectopic mineralization, such as kidney stones, calcific tendinitis, and skeletal fluorosis, shares some similar features but different mechanisms to physiological mineralization. A better understanding will provide new perspectives for treating pathological/ectopic mineralization-related diseases. This review provides an overview of the mechanisms of the crystallization and growth of crystals in physiological and pathological conditions from a chemistry perspective. By linking the microstructures and functions of crystals formed in both conditions, potential approaches are proposed to treat pathological/ectopic mineralization-related diseases.

Keywords

Physiological mineralization, pathological/ectopic mineralization, apatite crystals, mineral crystallinity, dental tissues, bone remodeling

INTRODUCTION

Biomineralization, an emerging interdisciplinary field, deals with the formation, structure, and mechanical strength of naturally formed mineralized tissues[1]. The skeleton of animals provides mechanical support to counteract gravitational forces on land and hydrostatic pressure in the depths of the oceans. As a highly complex and exquisitely organized organ, the skeleton is structurally, and hence mechanically, heterogeneous owing to spatial distributions in the shape, size, and composition of its constituent building blocks, the mineralized collagen fibrils[2].

Nature produces a diverse assortment of mineralized structures with a high degree of complexity. These hierarchical structures exhibit superior mechanical strength[3-5] and are, therefore, of great interest to researchers in the disciplines of biotechnology and biomedical engineering. Biomineralization on a template of organic molecules is used by many organisms to produce inorganic-organic nanocomposites that result in highly ordered multifunctional materials. The biosynthesis of bacterial magnetosome[6], eggshell[7], molluscan shell[8], dental structures[9], and skeletal system[10] are all examples of biologically controlled mineral formation through organic/inorganic recognition and interaction. In the skeleton, for example, the organic matrix consists primarily of the fibrous protein collagen and around 10% of other non-collagenous proteins (NCPs)[1,11] [Figure 1]. The inorganic phase is composed of tightly packed nanocrystals made of calcium phosphates (CaPs) with the incorporation of a number of essential trace elements.

Physiological and pathological/ectopic mineralization: from composition to microstructure

Figure 1. Physiological mineralization, mainly in bone and teeth, forms well-organized structures.

In the field of biomaterial development, the ex vivo bioactivity of the material is often predicted by examining the formation of an apatite layer on its surface in a simulated body fluid (SBF)[12]. Notably, slight differences in the SAED patterns of SBF-originated apatite on a Titanium substrate and bone apatite resulted from the random and ordered orientations, respectively, of apatite crystals [Figure 2][12-14]. However, there are other differences between SBF-derived apatite and bone apatite in terms of possible defects on the lattice of bone apatite due to the incorporation of trace elements in the body and crystal sizes affected by cells and cell-secreted bio-factors[15,16].

Physiological and pathological/ectopic mineralization: from composition to microstructure

Figure 2. Transmission electron microscope (TEM) photographs of (A) SBF-originated apatite crystals on a Titanium substrate (inset: SAED pattern) (Reproduced with permission from Takadama et al.[14]. Copyright 2001, John Wiley & Sons), (B-D) apatite crystals from mouse femur and their SAED patterns (Reproduced with permission from Taehoon et al.[13]. Copyright 2010, SpringerOpen).

Calcification has been increasingly recognized as an important component to fully understand the pathology of some diseases. For instance, the types of breast cancer have been shown to be related to the properties of calcification in the breast[17-19]. The utilization of advanced characterization techniques within the field of material science has facilitated significant advancements in our comprehension of the formation of pathological crystals, such as X-ray diffraction (XRD), spectroscopic, and electron microscopic techniques[20,21]. By comparing to crystals in physiological mineralization, understanding the characteristics and behaviors of crystals involved in pathological or ectopic calcification can offer valuable insights into the underlying mechanisms of calcification-related disorders. This knowledge may enable the development of novel disease management strategies.

Physiological mineralization is a complex process that is essential for the development of well-organized structures in bone and teeth[22]. The intricate process occurs only in specific regions[23,24]. The regulation of physiological mineralization is well-coordinated, involving both inhibitory and stimulatory factors. Some proteins, including osteopontin (OPN), matrix Gla protein, and pyrophosphate (PPi), have been identified as inhibitors of mineralization[22,25]. In contrast, other factors, such as matrix vesicles that contain calcium (Ca) and inorganic phosphate (Pi), apoptotic bodies, and tissue non-specific alkaline phosphatase, have been shown to facilitate the initiation of mineralization[26-28].

On the other hand, pathological/ectopic mineralization occurs in soft tissues and is associated with disease conditions or medical conditions, such as injury, inflammation, and aging, causing significant morbidity and mortality[29-32]. Such mineralization consists of different Ca-containing minerals, including CaPs, calcium carbonates (CaCO3), and calcium oxalates (CaOx)[20,31].

The current treatment of pathological mineralization in soft tissues includes reduced intake of Ca or reduced precipitation of Ca-containing complex. However, there are still no specific and effective treatments to prevent or counteract the condition[32,33]. Pathological/ectopic mineralization is less understood than physiological mineralization, but the commonalities between physiological and pathological/ectopic mineralization have gradually been recognized over the years[34-36]. There are still ongoing debates regarding the chemical compositions and crystal structures of pathological crystals, the dynamics of ion transport, the participation of cellular activities, the interactions of pathological crystals with surrounding tissues, and how these contribute to disease progression[36]. The compositional and structural analysis of pathological mineralization would benefit the development of better disease management through advanced treatment and prevention. In this review, bone and teeth mineralization and crystal structures are recognized as the fundamental and physiological processes in biomineralization. The comparison is made to unveil the similarities and differences in both physiological and pathological/ectopic mineralization. In addition, current approaches, with a focus on modulating crystal formation and growth on the progression of pathological mineralization, are discussed.

OVERVIEW OF THE PHYSIOLOGICAL MINERALIZATION PROCESS

The mineralization of self-assembled organic matrices results in the formation of hard tissues in the body, and its level (or the proportion of mineral contents in the tissue) varies from around 65% in the bone to close to 100% in dental enamel [Figure 1][37,38]. The fundamental form of both bone and teeth crystals is apatite. Its compositions and structures can be modified by accommodating a wide range of ion substitutions via cell uptake[39,40]. The investigation regarding the nucleation process of apatite in solution started several decades ago, attempting to unveil the mystery of mineralization[41-45]. Along with the emergence of edge-cutting equipment and the accumulation of knowledge over the years, a multi-stage nucleation process of apatite in aqueous solution was demonstrated to start with the aggregation of charged pre-nucleation complexes, [Ca(HPO4)3]4- (Ca/P(calcium/phosphorus) = 0.3-0.4)[46]. After the formation of ribbon-like calcium-deficient o phosphates (OCPs), [~Ca6(HPO4)4(PO4)22-] (Ca/P = 1.0) and elongated plate-like OCPs, [Ca8(HPO4)2(PO4)4] (Ca/P = 1.33) from amorphous calcium phosphates (ACPs), apatite is finally generated[46]. It has been suggested that the evolution of the OCP structure to apatite occurred through the elimination of the hydrated shell on the surface of the nanocrystalline[37,47]. The hydrated shell on the surface of freshly precipitated apatite contains exchangeable ionic species and some proteins[37,48]. During the maturation of apatite in the solution, ions in the hydrated shell are easily exchanged, accompanied by proteins within the shell, and it leads to reduced surface reactivity and increasingly stable apatite[37,49]. This suggests that metals (such as Ca, magnesium (Mg), sodium (Na), potassium (K), etc.), and possibly other ions are incorporated structurally into the collagenous matrices of calcified tissues[50]. In earlier investigations of the composition of trace elements in human cortical bone and beef tendons, researchers have documented a regularity in the appearance of certain trace metals, namely copper (Cu), iron (Fe), and zinc (Zn)[23,51,52]. Additionally, non-metal trace element, e.g., fluoride (F), has also been physio-chemically linked with the bone mineral matrix[53].

Mineralization of bone

A similar crystal formation mechanism was proposed in bone based on the similarity of XRD patterns of bone mineral and crystal formed from the ACP in an aqueous solution[41-43,45]. Bone minerals can be modeled as carbonate hydroxyapatite (CHAp) with a wide range of substitutions of hydroxide (OH-) and phosphate (PO43-) by carbonate (CO32-) on the lattice and the substitutions of Ca2+ by other metal ions, such as strontium (Sr2+), Mg2+, etc. [Figure 3][49,54]. Its composition can vary depending on species, location, diet, sex, age, and pathological conditions[37,55-57]. The substitution of carbonate results in the contraction of the a-axis and the expansion of the c-axis of the unit cell, as well as a decrease in crystallinity[58]. The presence of vacancies on the apatite crystals results in lower binding energies, thus, more soluble than stoichiometric hydroxyapatite (HAp)[25,59,60]. During the formation of bone minerals, ACP nanospheres formed on the site of bone formation or supplemented through blood are delivered and deposited within the collagen matrix by cells and further transformed into plate-like apatite via intermediates that resemble OCPs[44,61-63]. While Crane et al. (2006) suggest this possibility, the finding remains unreplicated and is not widely accepted[63]. However, recent evidence has shown the presence of OCPs in bone, specifically in combination with the protein osteocalcin[64]. This suggests that OCPs may play a role in bone mineral formation, although further research is needed to fully understand this process. Moreover, bone mineralization is much more complicated with the involvement of extensive biological additives, acting as promoters or inhibitors, such as proteins, trace ions, and some small organic molecules[42,65].

Physiological and pathological/ectopic mineralization: from composition to microstructure

Figure 3. A schematic illustration of the crystal lattice of bone apatite (Reproduced with permission from Ressler et al.[54]. Copyright 2021, Elsevier). (A) the unit cell of bone apatite; (B) the top view of the lattice of bone apatite.

Based on the discovery and transformation pathway from ACPs to apatite during bone mineralization, additives are believed to get involved by preventing unstable amorphous phases from crystallizing or dissolving before or during transport to the site of interest[42]. It is also known as the polymer-induced liquid precursor (PILP) process. Numerous studies aimed at identifying influencing factors in the process over the years. For example, many NCPs found in bone growth regions regulate mineralization with acidic amino acid groups through the ability to bind Ca2+, as well as high affinity for collagen fibrils, such as osteonectin and various phosphoproteins[66-68]. Organic Pi or polyphosphates, as Pi sources, participate in and affect the crystallization process of ACPs in the form of orthophosphates after being digested by alkaline phosphatase[32,69,70]. PPi, on the other hand, is a well-known inhibitor of apatite crystal formation[71].

Mineralization of teeth

In general, teeth mineral [Ca(10−x)Nax(PO4)(6−y)(CO3)(OH)(2−z)Fz] (x, cation substitutions; y, carbonate substitutions; z, fluorine substitutions) varies from both stoichiometric HAp and bone apatite[40]. Its chemical composition, the mineralization level of the organic matrix, and crystal size and orientation are tailored to meet different mechanical needs at different locations[37,72,73]. Enamel and dentin are two kinds of hard tissues in teeth. Dentin is capped by enamel and, in the root, is covered by cementum.

Dentin is less mineralized and shares a similar mineralization pattern with bone, namely matrix-mediated mineralization[74,75]. It is formed by the mineralization of the dentin matrix (composed of type I collagen), mediated by some non-collagenous matrix proteins (NCPs), such as dentin phosphoprotein (DPP), dentin sialoprotein (DSP), and dentin matrix protein-1 (DMP-1)[76]. However, unlike bone, little or no remodeling takes place in dentin. Compared to enamel, dentin has a higher capacity for F uptake through systemic ingestion due to its less crystallinity and accumulation through life[77].

Enamel is the hardest tissue in the vertebrate body, composed of crystalline HAp[78]. In contrast to the bone, which uses collagen as the substrate for mineralization and goes through remodeling throughout its lifetime, enamel does not contain collagen and does not remodel[76,79]. During enamel mineralization, the synthesized protein mixture assembles to form a matrix that regulates the precipitation of HAp[80,81]. Once the full thickness of enamel has been formed, HAp crystals expand slowly into the space previously occupied by matrix proteins and water[82]. Mature HAp crystals in enamel are ribbon-like fluoridated carbonate HAp, 50-70 nm in width and 20-25 nm in thickness[83].

Cementum is a thin layer that covers the entire root surface and anchors the roots of teeth to the jaw. Similar to the bone, it is composed of water, an organic matrix (mainly collagen, glycoproteins, and proteoglycans), and minerals and is formed through the formation of HAp in deposited type I collagen by cementoblasts[84-86]. Compared to dentin (70%) and enamel (> 95%), cementum (40%-50%) is less mineralized[86]. In addition to HAp, small amounts of ACPs were also found in cementum[85]. The presence of the amorphous phase results in a greater capacity of cementum for the adsorption of F and other elements over time and readily decalcification in acidic environments[85]. For example, cementum contains F up to 0.9% ash weight and increases with age[87]. Regulation of teeth mineralization is, in part, regulated by the ratio of Pi and PPi, as well as other factors, such as genetic modifications, age, diet, oral hygiene, and certain diseases, such as periodontal diseases, caries, root resorption, tumoral lesions, or trauma[88,89].

Otoconia

Otoconia, which is composed of CaCO3, is found in the inner ear of humans and other vertebrates. They are positioned to sense stimuli in directions and send signals to the brain to maintain bodily balance[90]. Similar to the bone, the organic matrix in otoconia determines the orientation, sizes, and shapes of crystals by providing a framework[91]. Additionally, otoconia contain high levels of Ca, Na, Mg, K, P, sulfur (S), and chloride (Cl)[90]. However, the underlying molecular etiology remains unknown, neither the functions of otoconial proteins nor the crystal formation.

PATHOLOGICAL (ECTOPIC) MINERALIZATION IN MAJOR ORGANS

Physiological mineralization is restricted within the skeletal system, including bone, teeth, and calcified cartilage[92]. In contrast, pathological (ectopic) mineralization can occur not only in the skeletal system but also in soft tissues, such as the breast, blood vessels, kidney, pancreas, and prostate, mostly composed of CaPs, CaCO3, and CaOx [Figure 4][22,31]. There are three structural types of deposits observed in ectopic calcification, single crystals, polycrystalline deposits, and calcified matrix. The specific type and characteristics of the deposits can vary, depending on the tissue and the underlying mechanisms. For example, single crystals are typically small and uniform in size and often are found in tendons and ligaments[93,94]. The formation of single crystals can be initiated by the presence of specific proteins or molecules that act as nucleation sites. Polycrystalline deposits, on the other hand, consist of multiple small crystals randomly arranged in tissues, such as blood vessels and heart valves[95]. In contrast, the calcified matrix consists of alternating layers of mineralized and non-mineralized tissues and can be found in a variety of tissues, including cartilage and blood vessels[96]. Pathological/ectopic calcification occurs in both genetic and acquired clinical conditions and affects the prognosis of diseases[32,97]. Pathological/ectopic calcification can occur through different mechanisms, depending on the involvement of cells, including cell-induced, cell-controlled, or spontaneously precipitated. For example, vascular calcification can be triggered by damage to the endothelial cells lining the blood vessels, leading to the recruitment and differentiation of smooth muscle cells into osteoblastic cells that actively deposit Ca and Pi[35]. Cell-controlled calcification occurs with specific mineralization-related factors released by cells in response to injury, inflammation, or bacterial infection[98,99]. Spontaneous deposition may happen as a result of changes in the environment without the involvement of cells, such as oxidative stress, local pH, and the supersaturation of ions[100,101]. Understanding the formation, structure, and composition of crystals deposited in the physiological and pathological conditions will aid the development of therapeutic strategies to prevent and/or regulate pathological/ectopic calcification[31,102] [Table 1].

Physiological and pathological/ectopic mineralization: from composition to microstructure

Figure 4. Pathological mineralization in soft tissues (blood vessels, brain, ocular, prostate, joint & tendon, placental, pancreas, kidney, and breast). The major forms of pathological/ectopic mineralization are calcium phosphates, calcium carbonate, and calcium oxalates.

Table 1

A summary of pathological/ectopic mineralization in terms of the composition and the involvement of cells

TissueCompositionInvolvement of cellular activitiesReferences
BoneHApCell-induced[112]
JointCPPD, HAp, TCP, OCP, whitlockite, sodium urateCell-derived articular cartilage vesicles[120,123,124,237]
TendonCarbonate HApCell-induced[94,128,130-132]
TeethBrushite, dicalcium phosphate dihydrate, OCP, HAp, whitlockiteCell-mediated[134,135]
Salivary glandHAp, whitlockite, brushite, OCPUnknown[137-139]
Heart and blood vesselHAp, whitlockiteCell-induced[20,35,140-142]
KidneyCaOx, CaP, the mixture of struvite magnesium ammonium phosphate, carbonate HApCell-mediated[151,153-155]
BrainHApUnknown[171,172]
OcularWhitlockite, HApUnknown[20,21]
BreastCaPs (carbonate HAp and whitlockite), CaOxCell-induced[182-185]
PancreasCaCO3Unknown[188,189]
ProstateCarbonate HAp, whitlockite, CaOx, Cell-mediated[191,192,238]
PlacentalCaPsCell-mediated[194,195]
Lymph nodesHAp, whitlockiteUnknown[197,198]

Bone

Bone tissue undergoes continuous remodeling to maintain its density and strength in a stable state. The bone mineralization process is tightly regulated by osteoblasts (bone-forming cells) and osteoclasts (bone resorption cells). Osteocytes, which are embedded in the bone matrix, also play important roles in regulating the bone remodeling process. The disruption in the balance between bone resorption and bone formation resulting from several pathological cues will affect bone health, metabolism, or homeostasis and lead to irregular remodeling activity and pathological bone mineralization[103]. Pathological bone mineralization can be roughly categorized as hypomineralization, hypermineralization, and other abnormal mineralization patterns.

Hypomineralization

Hypomineralization refers to insufficient mineralization of the bone matrix, resulting in reduced bone density and increased fracture risk[104]. Some several factors and conditions contribute to hypomineralization[105-107]. Rickets and osteomalacia are examples of pathological hypomineralization conditions[103]. Rickets, a disorder affecting children, and osteomalacia, its adult counterpart, can be caused by insufficient intake of essential nutrients, such as calcium and vitamin D, and hereditary factors[103]. The deficiency of calcium and vitamin D can result in a reduction in Ca that leads to impaired bone mineralization and a higher proportion of organic matrix compared to inorganic mineral content, which contributes to reduced mechanical strength and increased fracture risk[105].

Besides, certain genetic disorders can induce abnormal bone development, such as osteogenesis imperfecta, a condition characterized by fragile bones due to the mutation in genes responsible for collagen production that leads to aberrant collagen production, and hypophosphatasia, an inherited disorder that results from a deficiency in alkaline phosphatase, which is essential for proper bone mineralization[106,108].

Hypermineralization

Hypermineralization, characterized by abnormally high bone mineral density (BMD), is associated with hereditary and nonhereditary disorders[104]. For example, in hereditary disorders, such as osteopetrosis and osteosclerosis, a dramatic decrease in osteoclast number or osteoclast activity and increased bone mass and density were observed. At the crystal level, low remodeling activities alter the crystal size/shape and increase the amount of highly mineralized bone tissues, leading to decreased toughness of bone tissues and thus increasing the risk of fracture[104]. Skeletal fluorosis is another nonhereditary disorder associated with excessive ingestion of F. The accumulation of F results in abnormal bone deposition and adsorption. At the crystal level, the hydroxyl group in HAp is substituted by F, resulting in an increased crystallinity, crystal size, and elastic modulus[109].

Other abnormal mineralization patterns

Osteoporosis, one of the most prevalent pathological abnormal mineralization conditions, is mainly related to menopause and aging[104,110,111]. Osteoporosis is diagnosed based on low BMD and leads to reduced bone mass and deterioration in bone micro-architecture, which further increases the susceptibility to fracture[110]. In osteoporosis, bone resorption exceeds bone formation, resulting in altered mineralization properties, including mineral content, composition, and crystal size, which leads to reduced fracture resistance[110]. It has been reported that the bone tissue Ca/P ratio was reduced in osteoporosis patients and the imbalance increased the defect in the HAp network, thus resulting in a less rigid HAp crystal structure and further contributing to the increased fracture susceptibility in osteoporotic bones[112]. Meanwhile, the uneven distribution of minerals in the bone matrix is observed in some pathological conditions, such as Paget’s disease, which is a chronic disorder that the normal bone remodeling process is disrupted, leading to abnormally shaped, enlarged, and weakened bones that are more susceptible to fractures and deformities.

The presence of whitlockite in osteocyte lacunae (micropetrosis) is associated with pathological conditions, as it is primarily found in the pathological mineralization of various soft tissues, dental calculus, and occasionally in enamel and dentine[113]. Notably, Mg-whitlockite [Ca18Mg2(HPO4)2(PO4)12] has been detected in post-apoptotic osteocyte lacunae in human alveolar bone, but this unusual mineral has not been found in the extracellular matrix of mammalian bone under normal conditions[113]. The notion that Mg-whitlockite is a significant constituent of bone minerals remains unsubstantiated, and it is considered a pathological biomineral[114]. Therefore, contrary to some claims, biomaterials containing Mg-whitlockite do not represent a bioinspired or biomimetic approach to bone repair[113].

Joint

Cartilage calcification that has been observed in the hip and knee is pathological in almost all osteoarthritis (OA) patients[115]. The deposition of calcium pyrophosphate dihydrate (CPPD) has been found in patients suffering from acute attacks of pseudogout in the joints[116-118]. Other CaPs have also been found in synovial fluids, synovium, and cartilage of OA patients, such as carbonate HAp, tricalcium phosphate (TCP), and OCP[116,117,119,120]. Articular cartilage calcification that occurs in end-stage OA shows the accumulation of predominant HAp crystals[116,121]. The deposition of HAp initiates from articular cartilage vesicles (ACVs) and induces stress in articular chondrocytes, which ultimately results in the phenotypic change of chondrocytes and the formation of more HAp crystals[121,122]. Besides, whitlockite, another kind of CaP-based mineral, has also been found in the mineral phase of osteoarthritic articular cartilage and intervertebral disc[123,124]. Under certain circumstances, such as elevated Mg concentrations, acidic microenvironments, and the presence of specific proteins or organic molecules, Mg partly substitutes Ca, inhibits apatite originating from ACP, and aids the precipitation of whitlockite.

Tendon

Tendon mineralization (calcific tendinitis) is the cell-mediated calcification of living tissues that causes joint pain[94]. It typically affects the shoulder and hip, as well as other sites, such as the hand and wrist, foot and ankle, and neck[125]. Four well-defined phases of calcific tendinitis are formative (pre-calcific), calcific, resorptive, and reparative[94,126]. In the formative phase, a portion of the tendon undergoes a fibrocartilaginous transformation. During the calcific phase, calcium crystals are deposited in small vascular structures located between collagen fibrils[126,127]. Followed by calcification, the appearance of thin-walled vascular channels at the periphery of the calcium deposits marks the initiation of the resorptive phase, which involves Ca removal by macrophages and multinucleated giant cells[128]. Computed tomography studies showed the porous structure of calcific deposits throughout the tendon[129]. Structural and compositional analysis revealed that the chemical composition of mineralized deposits in calcific tendinitis is poorly crystallized carbonate HAp, having a Ca/P molar ratio ranging between 0.9 and 1.5[128,130-132].

Teeth

Pathological mineralization on teeth results from a variety of factors, such as changes in the oral environment, genetic predisposition, or underlying medical conditions[88,89]. In dentin, pathological mineralization happens in the form of dentinogenesis imperfecta, causing weakened teeth that are prone to cavities and fractures[133]. Dental calculus builds up on teeth (both supragingival and subgingival), is composed of inorganic components (brushite, dicalcium phosphate dihydrate, OCP, HAp, and whitlockite) and salivary proteins adsorbed on the tooth surface[134,135]. The level of dental calculus is affected by oral hygiene habits, diet, age, systemic diseases, and medications[135]. In enamel, pathological mineralization appears as spots on the surface of the tooth. It can be caused by malnutrition, high fever during childhood, or exposure to excessive F[77].

Salivary gland

A salivary gland stone (salivary calculi or sialolithiasis) found in the salivary gland consists of a central organic core and a layered cortex of inorganic components[136]. The crystals found in the salivary gland were randomly oriented carbonate HAp, whitlockite, and less frequently brushite and OCPs[137-139].

Heart and blood vessels

Cardiovascular calcification, the deposition of apatite and whitlockite, is closely associated with increased risks of several mortal diseases, such as blood vessel stenosis, ischemia, stroke, and heart disease[20,35,140-142]. Based on its histological appearance, the calcification can be distinguished as amorphous (lacking tissue architecture) and chondro-osseous (having the tissue architecture of cartilage or bone)[143].

Calcification that occurs in the intimal layer of the arteries links to arterial obstruction and atherosclerotic plaque rupture, leading to stroke or myocardial infarction[35,144]. Calcification that occurs in the medial layer, also known as Monckeberg’s medial sclerosis, is associated with vessel stiffness, systolic hypertension, and progressively increased risks of diastolic dysfunction and heart failure[35,145,146]. Even though the mechanism of cardiovascular calcification is not fully understood, it is no longer acknowledged as a passive accumulation of minerals but an active inflammatory and/or osteogenesis-related signaling process[147,148]. Several influencing factors have also been identified in the past few decades, such as the loss of inhibition, induction of osteogenesis, accumulation of nuclei, and cell death[98]. For example, cell-secreted small membrane-bound microparticles, such as apoptotic bodies and matrix vesicles, showed their capacities to concentrate Ca and Pi and initiate crystal nucleation in response to inflammatory stimuli[100,144]. Additionally, lipids and cytokines released by macrophages were also found to accelerate osteogenic differentiation and calcification of vascular smooth muscle cells (VSMCs)[100,101].

Kidney

Kidney stones are one of the most common and painful urinary disorders[149]. In most cases, they result from the nucleation and aggregation/growth of crystals from supersaturated urine [Figure 5][150,151]. Based on their composition and pathogenesis, kidney stones are commonly classified into five types: (a) Ca-containing crystals; (b) struvite or magnesium ammonium phosphate stones; (c) uric acid stones or urate; (d) cystine stones; and (e) drug-induced stones[152].

Physiological and pathological/ectopic mineralization: from composition to microstructure

Figure 5. Crystal deposition in the urinary system (Reproduced with permission from Evan et al.[150]. Copyright 2005, Elsevier). (A) The initial sites of the deposition of kidney stones in transmission electron microscopic images and (B) immunogold staining showed the localization of osteopontin in the plaque.

Ca-containing stones are predominant renal stones, typically composed of CaOx (50%), CaP (5%), or a mixture of both (45%)[153]. CaOx exists in the form of CaOx monohydrate (COM, or whewellite), CaOx dihydrate (COD, or weddellite), or a mixture of both forms[151,154]. Struvite stones in the body are shown to be a mixture of struvite magnesium ammonium phosphate [MgNH4PO4·6H2O] and carbonate HAp[155]. In addition to the most common Ca-containing stone that comprises 80% of urinary calculi, struvite accounts for 10%-15%, and uric acid stones account for 9%, leaving 1% of the rest of the stone types, including cystine stones and drug-induced stones[152,156,157].

The pathogenesis of kidney stones is a multi-step process, including crystal nucleation, aggregation, and retention. The nucleation of calculus occurs in a supersaturated liquid, where Ca and oxalate combine to form insoluble micro-clusters or start with existing nucleation centers [Figure 5][150,152,157]. Crystal aggregation and secondary nucleation of crystals account for crystal growth and interactions between crystals and cells and extracellular matrix[158]. The formation of kidney stones is affected by the supersaturation of Ca and oxalate in the urine, urinary pH, and some molecular modulators in the environment[157,159,160]. For instance, the growth of CaOx crystals is associated with a urinary pH range of 5.0 to 6.5; uric acid stones form at low urinary pH (pH < 5.05), whereas CaP and struvite favor a pH range over 7[152]. In addition to factors that affect the formation of kidney stones by altering the environment, there are several modulators actively modulating the nucleation and aggregation of pathological crystals. For example, PPi and citrate reduce the nucleation and growth of Ca-containing crystals, as well as Mg[157,160,161]; Biomolecules, such as OPN, prothrombin F1 fragment, and calgranulin, have been shown to inhibit the crystallization of CaOx and CaP by binding to Ca and hindering crystal growth[157,162-164].

Brain

Brain mineralization occurs in the basal ganglia and other regions, such as the cerebellum, thalamus, and brainstem[165,166]. Patients presenting brain calcification exhibit impaired motor and cognition, such as dystonia, parkinsonism, psychosis, and dementia[165,167,168]. It is associated with neurological or metabolic disorders (secondary hypoparathyroidism and mitochondrial diseases) and other acquired conditions, such as injuries (infection, ischemia, and trauma) and toxicity related to manganese (Mn), Fe, lead (Pb), carbon monoxide, and normal aging[165,169,170]. The major composition of brain mineralization is HAp[171,172]. Other components, such as Zn, Fe, and Mg, are also present in the deposits[173]. However, currently, there is no established method to determine whether it is a passive process, which can be attributed to normal aging, or an active process mediated by cellular activities[171,172].

Ocular

Ocular mineralization can be found in the cornea, retina, optic nerve, and Bruch’s membrane[92,174-176]. In the aging eye, the accumulation of protein- and lipid-containing deposits external to the retinal pigment epithelium (RPE) can lead to macular degradation and, consequently, blindness[92,177]. Three types of structures of minerals have been found in age-related macular degradation, spherules (whitlockite), plaques (amorphous apatite), and nodules (apatite)[20,21]. However, the mechanism of the formation of calcified nodules remains unknown.

Breast

Breast mineralization, also called microcalcification, has been suggested to be a consequence of either injury or diseases, such as chronic kidney disease, hypertension, or metabolic syndrome[178,179]. The formation of microcalcification is related to the acquisition of mesenchymal characteristics in breast cancer cells, affected by transforming growth factor beta (TGF-β) or nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB)[180,181]. Multiple phases of breast microcalcifications have been identified: CaPs (such as carbonate HAp and whitlockite), amorphous CaP, and, less commonly, CaOx[182-185]. However, the exact origins of minerals remained unclear. Nevertheless, the screening and evaluation of the morphology and distribution of microcalcification aid in determining the likelihood of whether the calcifications are benign, intermediate, or necessitate further investigation[17]. For instance, CaOx (type I calcification) is detected in benign breast lesions or lobular carcinoma, whereas HAp calcification (type II calcification) is detected in both benign and malignant breast tissues[18,19]. Compared to those formed in malignant ducts, type II microcalcification formed in benign ducts was found to contain a higher amount of CaCO3 and a lower amount of protein[186]. In some malignant specimens, Mg and Na have also been detected[187]. Elevated Na levels have been found in malignant specimens, but no correlation has been found between the level of Mg and malignancy[187].

Pancreas

Pancreatic stones (Pancreatic calculi) occur in the main ducts, side brunches, or parenchyma[188]. It contains an inner nidus core and outer shell, primarily constituting CaCO3, in the form of calcite (major) and aragonite, besides Pi and other protein content[188,189]. The inner protein nidus contained Fe, chromium (Cr), and nickel (Ni), whereas the outer calcite shell contained Ca and 17 other elements[190].

Prostate

Prostatic calculi are usually found as apatite or whitlockite, less frequently CaOx[191,192]. They are classified as primary/endogenous or secondary/extrinsic stones[193]. Endogenous stones are commonly caused by obstruction of the prostatic ducts or chronic inflammation; extrinsic stones are mainly caused by urine reflux[193]. Klimas et al. suggested two mechanisms regarding the calcification of prostatic calculi: the calcification of the corpora amylacea (type I calculi, composed of Na, S, P, Ca, and Zn) and the precipitation of prostatic secretions (type II calculi, composed of Ca and P)[191].

Placental

The placenta, a highly vascularized organ, mediates the communications between two circulatory systems. Placental calcification occurs when small calcium deposits build up on the placenta and is recognized as CaPs (the ratio of Ca/P = 2.00 ± 0.05) observed in placenta tissue[194,195]. It is often found in both preterm and term birth and serves as a predictor of adverse pregnancy outcomes[196]. However, the mechanisms of placental calcification are poorly understood[194].

Lymph nodes

Calcification in the lymph nodes occurs following diseases such as granulomatous infections[197]. Two patterns of calcification were found in the submandibular and neck regions: small blocks with different textures and islets surrounded by soft tissues, respectively[198]. In the submandibular specimen, apatite, together with a minor amount of whitlockite, was found[198]; In the neck specimen, whitlockite was found most frequently[198].

POSSIBLE PREVENTION AND TREATMENTS OF PATHOLOGICAL MINERALIZATION

Current treatments for the removal of calcified plaque include mechanical removal using rotary blade ablation or chemically dissolving in situ with acid via a catheter device[199-201]. In addition to mechanical and chemical debridement, reduced mineral ion intake from the diet or ions from bone stores have also been considered potential treatments for pathological/ectopic calcification in clinical practices[201]. Other approaches have been explored in targeting regulatory molecules in the body. For example, the monoclonal antibody Denosumab was developed to interfere with osteoclast functions and serum Ca levels, and PPi inhibited uremic vascular calcification without interfering with the mineralization of bone[202,203]. However, there are still no well-established and acknowledged treatments to prevent or counteract pathological/ectopic mineralization.

Current strategies targeting cellular mechanisms of calcification provided promising avenues for better administration of pathological mineralization in soft tissues, either reducing the nucleation of pathological crystals and crystal growth or reducing the circulating mineral ions in the serum[204]. Several endogenous calcification inhibitors have been identified, such as fetuin-A, vitamin-K dependent matrix-Gla protein (MGP), PPi, and OPN[23,33]. The serum protein fetuin-A is an important inhibitor of extra-skeletal calcification in the plasma and tissue fluids[205]. It is derived from the liver and acts as a chaperone, stabilizing excess mineral ions by forming a colloid with mineral ions or completely insoluble calciprotein particles[206-208]. MGP was the first inhibitor of artery calcification to be characterized in vivo[24]. It serves as an inhibitor of bone morphogenic protein-2 (BMP2), driving changes in cells toward an osteogenic-like phenotype and subsequent calcification[209,210]. PPi is identified as the principal inhibitor of HAp deposition, evidenced by antagonizing the ability of Pi to crystalize with Ca to inhibit biomineralization and acquired clinical conditions regarding pathological calcification upon its deficiency[32]. OPN is found to be associated with nascent mineralization foci during bone mineralization and present at interfaces where mineralization is needed to be quenched[211,212].

Several approaches to regulating the metabolisms of these natural inhibitors exhibited promising results in treating pathological mineralization. For example, Vitamin K supplementation showed the effect of reducing the progression of vascular mineralization and subsequent arterial stiffness[33,213]. Bisphosphonates, derivatives of PPi, stabilize the Pi groups by a phosphorus-carbon-phosphorus backbone. Compared to PPi, which can be easily hydrolyzed, bisphosphonates have extended plasma half-life and improved stability[32]. Bisphosphonates showed their potency to reduce fractures in postmenopausal osteoporosis[214]. They have also been found to reduce soft tissue calcification and interfere with osteoclast functions[33,204]. Alendronate, a widely prescribed bisphosphonate, has shown good tolerance and improvements in several patients with brain calcification[167]. Other bisphosphonates, such as etidronate, pamidronate, and ibandronate, have been demonstrated to inhibit aortic calcification both in vitro and in vivo[202,215-218]. In addition to bisphosphonates, a few proteins associated with the metabolism of PPi have also been identified to be promising for targeted treatment of pathological mineralization, such as ectonucleotide pyrophosphatase/phosphodiesterase-1 (ENPP1), ATP binding cassette subfamily C member 6 (ABCC6), progressive ankylosis protein (ANK), and tissue-nonspecific alkaline phosphatase (TNAP)[219-222]. Inspired by the inhibitory effect of acidic urinary macromolecules (e.g., OPN) on the formation of CaOxin vitro, polymeric carboxylic amino acids, poly-L-glutamic and aspartic acid, were found to inhibit the crystallization of CaOx[223]. A similar inhibitory effect was also found to be achieved by acidic polyanion poly (acrylic acid)[224-226]. Additionally, sodium thiosulfate (STS) treatment was found to inhibit calcium salt precipitation in calciphylaxis[227-230]. Nevertheless, these results were limited either in the long-term efficacy or potential toxicity of the STS treatment[33].

Moreover, the presence of trace elements has been shown to play a role in the crystal formation kinetics or external morphology of a growing crystal. It could be promising in designing therapeutic approaches again pathological mineralization[102,231]. For example, Mg, Zn, aluminum (Al), Fe, and Cu are indicated as growth inhibitors of CaOx at very low concentrations[232-235]. Adding AlCl3 or FeCl3 to transplanted valves effectively delayed the onset of valve calcification by blocking TNAP activity[236].

CONCLUSIONS AND FUTURE PERSPECTIVES

Our understanding of pathological/ectopic mineralization has been dramatically improved in parallel to bone biology and advanced characterization techniques widely used in the field of material science[22,92,201]. Significant progress has been made in elucidating the properties of pathological crystals and the underlying mechanisms of pathological/ectopic calcification that occurred in soft tissues over the years. By elucidating the physiological mechanisms of bone mineralization and their relationship with mineralization phenomena, we can develop targeted therapeutic interventions to prevent, manage, or treat these disorders. However, there are still some debatable questions regarding (a) the composition and structure of pathological crystals mediated by cellular activities and factors in the surrounding environment; (b) the dynamics of ion transport; (c) the involvement of cells and related cellular activities; (d) the interactions between pathological crystals and the surrounding tissues; and (e) how these contribute to disease progression[36]. Even though attempts have been made to simulate the crystallization process, it is still challenging to comprehend the in situ crystallization process due to the involvement of cellular activities. Nevertheless, a more comprehensive understanding of how the structures are formed, progressed, and adapted to changing needs enables us to conceive new insights into the progression of the pathological condition and guide future therapeutic designs to prevent and manage pathological/ectopic mineralization.

DECLARATIONS

Authors’ contributions

Conceptualization: Mu Y

Writing-original draft: Mu Y, Gao W

Writing-review and editing: Mu Y, Gao W, Zhou Y, Xiao L, Xiao Y

Availability of data and materials

Not applicable.

Financial support and sponsorship

This research is supported by the Joint Research Centre Fund from the Department of Environment and Science (2019-2023), Queensland.

Conflicts of interest

All authors declared that there are no conflicts of interest.

Ethical approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Copyright

© The Author(s) 2023.

REFERENCES

1. Weiner S, Dove PM. An overview of biomineralization processes and the problem of the vital effect. Rev Mineral Geochem 2003;54:1-29.

2. Tai K, Dao M, Suresh S, Palazoglu A, Ortiz C. Nanoscale heterogeneity promotes energy dissipation in bone. Nat Mater 2007;6:454-62.

3. Lichtenegger HC, Schöberl T, Bartl MH, Waite H, Stucky GD. High abrasion resistance with sparse mineralization: copper biomineral in worm jaws. Science 2002;298:389-92.

4. Hamm CE, Merkel R, Springer O, et al. Architecture and material properties of diatom shells provide effective mechanical protection. Nature 2003;421:841-3.

5. Tommasini SM, Wearne SL, Hof PR, Jepsen KJ. Percolation theory relates corticocancellous architecture to mechanical function in vertebrae of inbred mouse strains. Bone 2008;42:743-50.

6. Murat D, Falahati V, Bertinetti L, et al. The magnetosome membrane protein, MmsF, is a major regulator of magnetite biomineralization in Magnetospirillum magneticum AMB-1. Mol Microbiol 2012;85:684-99.

7. Gautron J, Stapane L, Le Roy N, Nys Y, Rodriguez-Navarro AB, Hincke MT. Avian eggshell biomineralization: an update on its structure, mineralogy and protein tool kit. BMC Mol Cell Biol 2021;22:11.

8. Ramos-Silva P, Benhamada S, Le Roy N, et al. Novel molluskan biomineralization proteins retrieved from proteomics: a case study with Upsalin. Chembiochem 2012;13:1067-78.

9. Beniash E, Simmer JP, Margolis HC. Structural changes in amelogenin upon self-assembly and mineral interactions. J Dent Res 2012;91:967-72.

10. Hosseini S, Naderi-Manesh H, Mountassif D, Cerruti M, Vali H, Faghihi S. C-terminal amidation of an osteocalcin-derived peptide promotes hydroxyapatite crystallization. J Biol Chem 2013;288:7885-93.

11. Weiner S, Wagner HD. The material bone: structure-mechanical function relations. Annu Rev Mater Sci 1998;28:271-98.

12. Kokubo T, Takadama H. How useful is SBF in predicting in vivo bone bioactivity? Biomaterials 2006;27:2907-15.

13. Taehoon J, Changyeon K, Eunkyung K, et al. Analysis of microstructure in mouse femur and decalcification effect on microstructure by electron microscopy. J Anal Sci Technol 2010;1:124-9. Available from: https://www.researchgate.net/publication/49587402_Analysis_of_microstructure_in_mouse_femur_and_decalcification_effect_on_microstructure_by_electron_microscopy [Last accessed on 5 July 2023].

14. Takadama H, Kim HM, Kokubo T, Nakamura T. TEM-EDX study of mechanism of bonelike apatite formation on bioactive titanium metal in simulated body fluid. J Biomed Mater Res 2001;57:441-8.

15. Baino F, Yamaguchi S. The use of simulated body fluid (SBF) for assessing materials bioactivity in the context of tissue engineering: review and challenges. Biomimetics 2020;5:57.

16. Suchý T, Bartoš M, Sedláček R, et al. Various simulated body fluids lead to significant differences in collagen tissue engineering scaffolds. Materials 2021;14:4388.

17. Bonfiglio R, Scimeca M, Urbano N, Bonanno E, Schillaci O. Breast microcalcifications: biological and diagnostic perspectives. Future Oncol 2018;14:3097-9.

18. Busing CM, Keppler U, Menges V. Differences in microcalcification in breast tumors. Virchows Arch A 1981;393:307-13.

19. Barman I, Dingari NC, Saha A, et al. Application of Raman spectroscopy to identify microcalcifications and underlying breast lesions at stereotactic core needle biopsy. Cancer Res 2013;73:3206-15.

20. Tsolaki E, Bertazzo S. Pathological mineralization: the potential of mineralomics. Materials 2019;12:3126.

21. Tan ACS, Pilgrim MG, Fearn S, et al. Calcified nodules in retinal drusen are associated with disease progression in age-related macular degeneration. Sci Transl Med 2018;10:eaat4544.

22. Kirsch T. Determinants of pathological mineralization. Curr Opin Rheumatol 2006;18:174-80.

23. Reznikov N, Steele JAM, Fratzl P, Stevens MM. A materials science vision of extracellular matrix mineralization. Nat Rev Mater 2016;1:16041.

24. Luo G, Ducy P, McKee MD, et al. Spontaneous calcification of arteries and cartilage in mice lacking matrix GLA protein. Nature 1997;386:78-81.

25. Yagami K, Suh JY, Enomoto-Iwamoto M, et al. Matrix GLA protein is a developmental regulator of chondrocyte mineralization and, when constitutively expressed, blocks endochondral and intramembranous ossification in the limb. J Cell Biol 1999;147:1097-108.

26. Yuan FL, Xu RS, Ye JX, Zhao MD, Ren LJ, Li X. Apoptotic bodies from endplate chondrocytes enhance the oxidative stress-induced mineralization by regulating PPi metabolism. J Cell Mol Med 2019;23:3665-75.

27. Wu LN, Genge BR, Dunkelberger DG, LeGeros RZ, Concannon B, Wuthier RE. Physicochemical characterization of the nucleational core of matrix vesicles. J Biol Chem 1997;272:4404-11.

28. Sekaran S, Vimalraj S, Thangavelu L. The physiological and pathological role of tissue nonspecific alkaline phosphatase beyond mineralization. Biomolecules 2021;11:1564.

29. Franklin BS, Mangan MS, Latz E. Crystal formation in inflammation. Annu Rev Immunol 2016;34:173-202.

30. Poloni LN, Ward MD. The materials science of pathological crystals. Chem Mater 2014;26:477-95.

31. Bazin D, Daudon M, Combes C, Rey C. Characterization and some physicochemical aspects of pathological microcalcifications. Chem Rev 2012;112:5092-120.

32. Ralph D, van de Wetering K, Uitto J, Li Q. Inorganic pyrophosphate deficiency syndromes and potential treatments for pathologic tissue calcification. Am J Pathol 2022;192:762-70.

33. Singh A, Tandon S, Tandon C. An update on vascular calcification and potential therapeutics. Mol Biol Rep 2021;48:887-96.

34. Fuery MA, Liang L, Kaplan FS, Mohler ER. Vascular ossification: pathology, mechanisms, and clinical implications. Bone 2018;109:28-34.

35. Durham AL, Speer MY, Scatena M, Giachelli CM, Shanahan CM. Role of smooth muscle cells in vascular calcification: implications in atherosclerosis and arterial stiffness. Cardiovasc Res 2018;114:590-600.

36. Vidavsky N, Kunitake JAMR, Estroff LA. Multiple pathways for pathological calcification in the human body. Adv Healthc Mater 2021;10:e2001271.

37. Cazalbou S, Combes C, Eichert D, Rey C. Adaptative physico-chemistry of bio-related calcium phosphates. J Mater Chem 2004;14:2148-53.

38. Zipkin I. The inorganic composition of bones and teeth. In: Schraer H, editor. Biological calcification: cellular and molecular aspects. Boston: Springer; 1970. pp. 69-103.

39. Elsharkawy S, Mata A. Hierarchical biomineralization: from nature’s designs to synthetic materials for regenerative medicine and dentistry. Adv Healthc Mater 2018;7:e1800178.

40. Abou Neel EA, Aljabo A, Strange A, et al. Demineralization-remineralization dynamics in teeth and bone. Int J Nanomed 2016;11:4743-63.

41. Eanes ED, Gillessen IH, Posner AS. Intermediate states in the precipitation of hydroxyapatite. Nature 1965;208:365-7.

42. Habraken W, Habibovic P, Epple M, Bohner M. Calcium phosphates in biomedical applications: materials for the future? Mater Today 2016;19:69-87.

43. Beniash E, Metzler RA, Lam RS, Gilbert PU. Transient amorphous calcium phosphate in forming enamel. J Struct Biol 2009;166:133-43.

44. Mahamid J, Aichmayer B, Shimoni E, et al. Mapping amorphous calcium phosphate transformation into crystalline mineral from the cell to the bone in zebrafish fin rays. Proc Natl Acad Sci USA 2010;107:6316-21.

45. Termine JD, Posner AS. Infrared analysis of rat bone: age dependency of amorphous and crystalline mineral fractions. Science 1966;153:1523-5.

46. Habraken WJ, Tao J, Brylka LJ, et al. Ion-association complexes unite classical and non-classical theories for the biomimetic nucleation of calcium phosphate. Nat Commun 2013;4:1507.

47. Brown WE, Schroeder LW, Ferris JS. Interlayering of crystalline octacalcium phosphate and hydroxylapatite. J Phys Chem 1979;83:1385-8.

48. Wang Y, Von Euw S, Fernandes FM, et al. Water-mediated structuring of bone apatite. Nat Mater 2013;12:1144-53.

49. Rey C, Combes C. What bridges mineral platelets of bone? Bonekey Rep 2014;3:586.

50. Dorvee JR, Veis A. Water in the formation of biogenic minerals: peeling away the hydration layers. J Struct Biol 2013;183:278-303.

51. Becker RO, Spadaro JA, Berg EW. The trace elements of human bone. J Bone Joint Surg Am 1968;50:326-34.

52. Ellis EH, Spadaro JA, Becker RO. Trace elements in tendon collagen. Clin Orthop Relat Res 1969;65:195-8. Available from: https://www.robertobecker.net/PDFs/BF050-CORR1969.pdf [Last accessed on 5 July 2023].

53. Grynpas MD, Rey C. The effect of fluoride treatment on bone mineral crystals in the rat. Bone 1992;13:423-9.

54. Ressler A, Žužić A, Ivanišević I, Kamboj N, Ivanković H. Ionic substituted hydroxyapatite for bone regeneration applications: a review, Open Ceram 2021;6:100122.

55. Legros R, Balmain N, Bonel G. Age-related changes in mineral of rat and bovine cortical bone. Calcif Tissue Int 1987;41:137-44.

56. Handschin RG, Stern WB. Crystallographic and chemical analysis of human bone apatite (Crista Iliaca). Clin Rheumatol 1994;13 Suppl 1:75-90.

57. Yao X, Carleton SM, Kettle AD, Melander J, Phillips CL, Wang Y. Gender-dependence of bone structure and properties in adult osteogenesis imperfecta murine model. Ann Biomed Eng 2013;41:1139-49.

58. Nelson DG. The influence of carbonate on the atomic structure and reactivity of hydroxyapatite. J Dent Res 1981;60 Spec No C:1621-9.

59. Farlay D, Panczer G, Rey C, Delmas PD, Boivin G. Mineral maturity and crystallinity index are distinct characteristics of bone mineral. J Bone Miner Metab 2010;28:433-45.

60. Lacruz RS, Habelitz S, Wright JT, Paine ML. Dental enamel formation and implications for oral health and disease. Physiol Rev 2017;97:939-93.

61. Akiva A, Malkinson G, Masic A, et al. On the pathway of mineral deposition in larval zebrafish caudal fin bone. Bone 2015;75:192-200.

62. Bennet M, Akiva A, Faivre D, et al. Simultaneous Raman microspectroscopy and fluorescence imaging of bone mineralization in living zebrafish larvae. Biophys J 2014;106:L17-9.

63. Crane NJ, Popescu V, Morris MD, Steenhuis P, Ignelzi MA Jr. Raman spectroscopic evidence for octacalcium phosphate and other transient mineral species deposited during intramembranous mineralization. Bone 2006;39:434-42.

64. Simon P, Grüner D, Worch H, et al. First evidence of octacalcium phosphate@osteocalcin nanocomplex as skeletal bone component directing collagen triple-helix nanofibril mineralization. Sci Rep 2018;8:13696.

65. Nudelman F, Lausch AJ, Sommerdijk NA, Sone ED. In vitro models of collagen biomineralization. J Struct Biol 2013;183:258-69.

66. Olszta MJ, Cheng X, Jee SS, et al. Bone structure and formation: a new perspective. Mater Sci Eng R Rep 2007;58:77-116.

67. George A, Veis A. Phosphorylated proteins and control over apatite nucleation, crystal growth, and inhibition. Chem Rev 2008;108:4670-93.

68. Gower LB. Biomimetic model systems for investigating the amorphous precursor pathway and its role in biomineralization. Chem Rev 2008;108:4551-627.

69. de Jonge LT, van den Beucken JJ, Leeuwenburgh SC, Hamers AA, Wolke JG, Jansen JA. In vitro responses to electrosprayed alkaline phosphatase/calcium phosphate composite coatings. Acta Biomater 2009;5:2773-82.

70. Omelon SJ, Grynpas MD. Relationships between polyphosphate chemistry, biochemistry and apatite biomineralization. Chem Rev 2008;108:4694-715.

71. Fleisch H, Bisaz S. Isolation from urine of pyrophosphate, a calcification inhibitor. Am J Physiol 1962;203:671-5.

72. Cuy JL, Mann AB, Livi KJ, Teaford MF, Weihs TP. Nanoindentation mapping of the mechanical properties of human molar tooth enamel. Arch Oral Biol 2002;47:281-91.

73. Tjäderhane L, Carrilho MR, Breschi L, Tay FR, Pashley DH. Dentin basic structure and composition-an overview. Endod Topics 2009;20:3-29.

74. Linde A. Dentin mineralization and the role of odontoblasts in calcium transport. Connect Tissue Res 1995;33:163-70.

75. Hao J, Ramachandran A, George A. Temporal and spatial localization of the dentin matrix proteins during dentin biomineralization. J Histochem Cytochem 2009;57:227-37.

76. Chatzistavrou X, Papagerakis S, Ma PX, Papagerakis P. Innovative approaches to regenerate enamel and dentin. Int J Dent 2012;2012:856470.

77. Vieira AP, Hancock R, Limeback H, Maia R, Grynpas MD. Is fluoride concentration in dentin and enamel a good indicator of dental fluorosis? J Dent Res 2004;83:76-80.

78. White SN, Luo W, Paine ML, Fong H, Sarikaya M, Snead ML. Biological organization of hydroxyapatite crystallites into a fibrous continuum toughens and controls anisotropy in human enamel. J Dent Res 2001;80:321-6.

79. Giacaman R, Perez V, Carrera C. 5 - Mineralization processes in hard tissues: teeth. In: Biomineralization and biomaterials. Amsterdam, The Netherlands: Elsevier; 2016. pp. 147-85.

80. Wang L, Guan X, Du C, Moradian-Oldak J, Nancollas GH. Amelogenin promotes the formation of elongated apatite microstructures in a controlled crystallization system. J Phys Chem C Nanomater Interfaces 2007;111:6398-404.

81. Robinson C, Kirkham J, Brookes SJ, Bonass WA, Shore RC. The chemistry of enamel development. Int J Dev Biol 1995;39:145-52.

82. Smith CE. Cellular and chemical events during enamel maturation. Crit Rev Oral Biol Med 1998;9:128-61.

83. Moradian-Oldak J. Protein-mediated enamel mineralization. Front Biosci 2012;17:1996-2023.

84. Yamamoto T, Hasegawa T, Yamamoto T, Hongo H, Amizuka N. Histology of human cementum: its structure, function, and development. Jpn Dent Sci Rev 2016;52:63-74.

85. Gonçalves PF, Sallum EA, Sallum AW, Casati MZ, Toledo S, Júnior FHN. Dental cementum reviewed: development, structure, composition, regeneration and potential functions. Braz J Oral Sci 2005;4:651-8. Available from: https://periodicos.sbu.unicamp.br/ojs/index.php/bjos/article/view/8641790 [Last accessed on 5 July 2023].

86. Neiders ME, Eick JD, Miller WA, Leitner JW. Electron probe microanalysis of cementum and underlying dentin in young permanent teeth. J Dent Res 1972;51:122-30.

87. Nakagaki H, Weatherell JA, Strong M, Robinson C. Distribution of fluoride in human cementum. Arch Oral Biol 1985;30:101-4.

88. Andras NL, Mohamed FF, Chu EY, Foster BL. Between a rock and a hard place: regulation of mineralization in the periodontium. Genesis 2022;60:e23474.

89. Foster BL, Nociti FH, Somerman MJ. Development and structure of cementum. In: Naji S, Rendu W, Gourichon L, editors. Dental cementum in anthropology. Cambridge: Cambridge University Press; 2022. pp. 46-64.

90. Lundberg YW, Xu Y, Thiessen KD, Kramer KL. Mechanisms of otoconia and otolith development. Dev Dyn 2015;244:239-53.

91. Zhao X, Yang H, Yamoah EN, Lundberg YW. Gene targeting reveals the role of Oc90 as the essential organizer of the otoconial organic matrix. Dev Biol 2007;304:508-24.

92. Thompson RB, Reffatto V, Bundy JG, et al. Identification of hydroxyapatite spherules provides new insight into subretinal pigment epithelial deposit formation in the aging eye. Proc Natl Acad Sci USA 2015;112:1565-70.

93. Landis WJ, Moradian-Oldak J, Weiner S. Topographic imaging of mineral and collagen in the calcifying turkey tendon. Connect Tissue Res 1991;25:181-96.

94. Siegal DS, Wu JS, Newman JS, Del Cura JL, Hochman MG. Calcific tendinitis: a pictorial review. Can Assoc Radiol J 2009;60:263-72.

95. Reynolds JL, Skepper JN, McNair R, et al. Multifunctional roles for serum protein fetuin-a in inhibition of human vascular smooth muscle cell calcification. J Am Soc Nephrol 2005;16:2920-30.

96. de Faria LL, Babler F, Ferreira LC, de Noronha Junior OA, Marsolla FL, Ferreira DL. Soft tissue calcifications: a pictorial essay. Radiol Bras 2020;53:337-44.

97. Wang D, Wang X, Huang L, et al. Unraveling an innate mechanism of pathological mineralization-regulated inflammation by a nanobiomimetic system. Adv Healthc Mater 2021;10:e2101586.

98. Giachelli CM. Vascular calcification mechanisms. J Am Soc Nephrol 2004;15:2959-64.

99. Schrijvers DM, De Meyer GR, Kockx MM, Herman AG, Martinet W. Phagocytosis of apoptotic cells by macrophages is impaired in atherosclerosis. Arterioscler Thromb Vasc Biol 2005;25:1256-61.

100. Shanahan CM. Inflammation ushers in calcification: a cycle of damage and protection? Circulation 2007;116:2782-5.

101. Pugliese G, Iacobini C, Blasetti Fantauzzi C, Menini S. The dark and bright side of atherosclerotic calcification. Atherosclerosis 2015;238:220-30.

102. Słojewski M, Czerny B, Safranow K, et al. Microelements in stones, urine, and hair of stone formers: a new key to the puzzle of lithogenesis? Biol Trace Elem Res 2010;137:301-16.

103. Bala Y, Farlay D, Boivin G. Bone mineralization: from tissue to crystal in normal and pathological contexts. Osteoporos Int 2013;24:2153-66.

104. Roschger P, Paschalis EP, Fratzl P, Klaushofer K. Bone mineralization density distribution in health and disease. Bone 2008;42:456-66.

105. Roschger P, Misof B, Paschalis E, Fratzl P, Klaushofer K. Changes in the degree of mineralization with osteoporosis and its treatment. Curr Osteoporos Rep 2014;12:338-50.

106. Kurdi MS. Chronic fluorosis: the disease and its anaesthetic implications. Indian J Anaesth 2016;60:157-62.

107. Rodan GA, Martin TJ. Therapeutic approaches to bone diseases. Science 2000;289:1508-14.

108. Faibish D, Ott SM, Boskey AL. Mineral changes in osteoporosis: a review. Clin Orthop Relat Res 2006;443:28-38.

109. Feroz S, Khan AS. 7-fluoride-substituted hydroxyapatite. In: Khan AS, Chaudhry AA, editor. Handbook of ionic substituted hydroxyapatites, Soston, UK: Woodhead, 2020; pp. 175-96.

110. Poole KE, Compston JE. Osteoporosis and its management. BMJ 2006;333:1251-6.

111. Tamimi I, Cortes ARG, Sánchez-Siles JM, et al. Composition and characteristics of trabecular bone in osteoporosis and osteoarthritis. Bone 2020;140:115558.

112. Nobakhti S, Shefelbine SJ. On the relation of bone mineral density and the elastic modulus in healthy and pathologic bone. Curr Osteoporos Rep 2018;16:404-10.

113. Shah FA. The many facets of micropetrosis - magnesium whitlockite deposition in bisphosphonate-exposed human alveolar bone with osteolytic metastasis. Micron 2023;168:103441.

114. Shah FA. Magnesium whitlockite - omnipresent in pathological mineralisation of soft tissues but not a significant inorganic constituent of bone. Acta Biomater 2021;125:72-82.

115. Meyer F, Dittmann A, Kornak U, et al. Chondrocytes from osteoarthritic and chondrocalcinosis cartilage represent different phenotypes. Front Cell Dev Biol 2021;9:622287.

116. Fuerst M, Bertrand J, Lammers L, et al. Calcification of articular cartilage in human osteoarthritis. Arthritis Rheum 2009;60:2694-703.

117. Derfus BA, Kurian JB, Butler JJ, et al. The high prevalence of pathologic calcium crystals in pre-operative knees. J Rheumatol 2002;29:570-4.

118. Halverson PB, McCarty DJ. Patterns of radiographic abnormalities associated with basic calcium phosphate and calcium pyrophosphate dihydrate crystal deposition in the knee. Ann Rheum Dis 1986;45:603-5.

119. Nalbant S, Martinez JA, Kitumnuaypong T, Clayburne G, Sieck M, Schumacher HR Jr. Synovial fluid features and their relations to osteoarthritis severity: new findings from sequential studies. Osteoarthr Cartil 2003;11:50-4. Available from: https://www.sciencedirect.com/science/article/pii/S1063458402908617 [Last accessed on 5 July 2023].

120. Rosenthal AK, Mattson E, Gohr CM, Hirschmugl CJ. Characterization of articular calcium-containing crystals by synchrotron FTIR. Osteoarthr Cartil 2008;16:1395-402.

121. Yan JF, Qin WP, Xiao BC, et al. Pathological calcification in osteoarthritis: an outcome or a disease initiator? Biol Rev Camb Philos Soc 2020;95:960-85.

122. Rosenthal AK. Articular cartilage vesicles and calcium crystal deposition diseases. Curr Opin Rheumatol 2016;28:127-32.

123. Scotchford CA, Vickers M, Ali SY. The isolation and characterization of magnesium whitlockite crystals from human articular cartilage. Osteoarthr Cartil 1995;3:79-94.

124. Lee RS, Kayser MV, Ali SY. Calcium phosphate microcrystal deposition in the human intervertebral disc. J Anat 2006;208:13-9.

125. Hayes CW, Conway WF. Calcium hydroxyapatite deposition disease. Radiographics 1990;10:1031-48.

126. Uhthoff HK, Loehr JW. Calcific tendinopathy of the rotator cuff: pathogenesis, diagnosis, and management. J Am Acad Orthop Surg 1997;5:183-91.

127. Landis WJ. A study of calcification in the leg tendons from the domestic turkey. J Ultrastruct Mol Struct Res 1986;94:217-38.

128. Oliva F, Via AG, Maffulli N. Physiopathology of intratendinous calcific deposition. BMC Med 2012;10:95.

129. McCarty DJ Jr, Gatter RA. Recurrent acute inflammation associated with focal apatite crystal deposition. Arthritis Rheum 1966;9:804-19.

130. Gärtner J, Simons B. Analysis of calcific deposits in calcifying tendinitis. Clin Orthop Relat Res 1990;254:111-20.

131. Riley GP, Harrall RL, Constant CR, Cawston TE, Hazleman BL. Prevalence and possible pathological significance of calcium phosphate salt accumulation in tendon matrix degeneration. Ann Rheum Dis 1996;55:109-15.

132. Penel G, Leroy G, Rey C, Bres E. MicroRaman spectral study of the PO4 and CO3 vibrational modes in synthetic and biological apatites. Calcif Tissue Int 1998;63:475-81.

133. Barron MJ, McDonnell ST, Mackie I, Dixon MJ. Hereditary dentine disorders: dentinogenesis imperfecta and dentine dysplasia. Orphanet J Rare Dis 2008;3:31.

134. Jin Y, Yip HK. Supragingival calculus: formation and control. Crit Rev Oral Biol Med 2002;13:426-41.

135. White DJ. Dental calculus: recent insights into occurrence, formation, prevention, removal and oral health effects of supragingival and subgingival deposits. Eur J Oral Sci 1997;105:508-22.

136. Sakae T, Yamamoto H, Hirai G. Mode of occurrence of brushite and whitlockite in a sialolith. J Dent Res 1981;60:842-4.

137. Anneroth G, Eneroth CM, Isacsson G. Crystalline structure of salivary calculi. a microradiographic and microdiffractometric study. J Oral Pathol 1975;4:266-72.

138. Burnstein LS, Boskey AL, Tannenbaum PJ, Posner AS, Mandel ID. The crystal chemistry of submandibular and parotid salivary gland stones. J Oral Pathol 1979;8:284-91.

139. Boskey AL, Burstein LS, Mandel ID. Phospholipids associated with human parotid gland sialoliths. Arch Oral Biol 1983;28:655-7.

140. Nicoll R, Henein MY. The predictive value of arterial and valvular calcification for mortality and cardiovascular events. Int J Cardiol Heart Vessel 2014;3:1-5.

141. Higgins CL, Marvel SA, Morrisett JD. Quantification of calcification in atherosclerotic lesions. Arterioscler Thromb Vasc Biol 2005;25:1567-76.

142. Reid JD, Andersen ME. Medial calcification (whitlockite) in the aorta. Atherosclerosis 1993;101:213-24.

143. Sage AP, Tintut Y, Demer LL. Regulatory mechanisms in vascular calcification. Nat Rev Cardiol 2010;7:528-36.

144. You AYF, Bergholt MS, St-Pierre JP, et al. Raman spectroscopy imaging reveals interplay between atherosclerosis and medial calcification in the human aorta. Sci Adv 2017;3:e1701156.

145. Chow B, Rabkin SW. The relationship between arterial stiffness and heart failure with preserved ejection fraction: a systemic meta-analysis. Heart Fail Rev 2015;20:291-303.

146. Barasch E, Gottdiener JS, Marino Larsen EK, Chaves PH, Newman AB. Cardiovascular morbidity and mortality in community-dwelling elderly individuals with calcification of the fibrous skeleton of the base of the heart and aortosclerosis (The Cardiovascular Health Study). Am J Cardiol 2006;97:1281-6.

147. Aikawa E, Nahrendorf M, Figueiredo JL, et al. Osteogenesis associates with inflammation in early-stage atherosclerosis evaluated by molecular imaging in vivo. Circulation 2007;116:2841-50.

148. Wick G, Grundtman C. Inflammation and atherosclerosis. Boston: Springer Science & Business Media; 2011.

149. Saita A, Bonaccorsi A, Motta M. Stone composition: where do we stand? Urol Int 2007;79 Suppl 1:16-9.

150. Evan AP, Coe FL, Rittling SR, et al. Apatite plaque particles in inner medulla of kidneys of calcium oxalate stone formers: osteopontin localization. Kidney Int 2005;68:145-54.

151. He JY, Deng SP, Ouyang JM. Morphology, particle size distribution, aggregation, and crystal phase of nanocrystallites in the urine of healthy persons and lithogenic patients. IEEE Trans Nanobiosci 2010;9:156-63.

152. Alelign T, Petros B. Kidney stone disease: an update on current concepts. Adv Urol 2018;2018:3068365.

153. Chaudhary A, Singla SK, Tandon C. In vitro evaluation of terminalia arjuna on calcium phosphate and calcium oxalate crystallization. Indian J Pharm Sci 2010;72:340-5.

154. Bensatal A, Ouahrani MR. Inhibition of crystallization of calcium oxalate by the extraction of Tamarix gallica L. Urol Res 2008;36:283-7.

155. Griffith DP. Struvite stones. Kidney Int 1978;13:372-82.

156. Luigia M, Summ V. A review of pathological biomineral analysis techniques and classification schemes. In: Aydinalp C, editor. An introduction to the study of mineralogy. London: InTech; 2012.

157. Coe FL, Evan A, Worcester E. Kidney stone disease. J Clin Invest 2005;115:2598-608.

158. Aggarwal KP, Narula S, Kakkar M, Tandon C. Nephrolithiasis: molecular mechanism of renal stone formation and the critical role played by modulators. Biomed Res Int 2013;2013:292953.

159. Coe FL, Parks JH, Asplin JR. The pathogenesis and treatment of kidney stones. N Engl J Med 1992;327:1141-52.

160. Fleisch H. Inhibitors and promoters of stone formation. Kidney Int 1978;13:361-71.

161. Ebrahimpour A, Perez L, Nancollas GH. Induced crystal growth of calcium oxalate monohydrate at hydroxyapatite surfaces. the influence of human serum albumin, citrate, and magnesium. Langmuir 1991;7:577-83.

162. Asplin JR, Arsenault D, Parks JH, Coe FL, Hoyer JR. Contribution of human uropontin to inhibition of calcium oxalate crystallization. Kidney Int 1998;53:194-9.

163. Stapleton A, Ryall U. Blood coagulation proteins and urolithiasis are linked: crystal matrix protein is the Fl activation peptide of human prothrombin. Br J Urol 1995;75:712-9.

164. Pillay SN, Asplin JR, Coe FL. Evidence that calgranulin is produced by kidney cells and is an inhibitor of calcium oxalate crystallization. Am J Physiol 1998;275:F255-61.

165. Loeb JA, Sohrab SA, Huq M, Fuerst DR. Brain calcifications induce neurological dysfunction that can be reversed by a bone drug. J Neurol Sci 2006;243:77-81.

166. Smeyers-Verbeke J, Michotte Y, Pelsmaeckers J, et al. The chemical composition of idiopathic nonarteriosclerotic cerebral calcifications. Neurology 1975;25:48-57.

167. Oliveira JR, Oliveira MF. Primary brain calcification in patients undergoing treatment with the biphosphanate alendronate. Sci Rep 2016;6:22961.

168. Lemos RR, Ferreira JBMM, Keasey MP, Oliveira JRM. Chapter Fourteen - an update on primary familial brain calcification. In: Bhatia KP, Schneider SA, editors. International review of neurobiology. Cambridge: Academic Press. 2013; pp. 349-71.

169. Jankovic J. Searching for a relationship between manganese and welding and Parkinson’s disease. Neurology 2005;64:2021-8.

170. Bekiesinska-Figatowska M, Mierzewska H, Jurkiewicz E. Basal ganglia lesions in children and adults. Eur J Radiol 2013;82:837-49.

171. Simoni M, Pantoni L, Pracucci G, et al. Prevalence of CT-detected cerebral abnormalities in an elderly Swedish population sample. Acta Neurol Scand 2008;118:260-7.

172. Yamada M, Asano T, Okamoto K, et al. High frequency of calcification in basal ganglia on brain computed tomography images in Japanese older adults. Geriatr Gerontol Int 2013;13:706-10.

173. Deng H, Zheng W, Jankovic J. Genetics and molecular biology of brain calcification. Ageing Res Rev 2015;22:20-38.

174. Anderson SB, de Souza RF, Hofmann-Rummelt C, Seitz B. Corneal calcification after amniotic membrane transplantation. Br J Ophthalmol 2003;87:587-91.

175. Nicholson BP, Lystad LD, Emch TM, Singh AD. Idiopathic dural optic nerve sheath calcification. Br J Ophthalmol 2011;95:290, 299.

176. Russell SR, Mullins RF, Schneider BL, Hageman GS. Location, substructure, and composition of basal laminar drusen compared with drusen associated with aging and age-related macular degeneration. Am J Ophthalmol 2000;129:205-14.

177. Hageman GS, Anderson DH, Johnson LV, et al. A common haplotype in the complement regulatory gene factor H (HF1/CFH) predisposes individuals to age-related macular degeneration. Proc Natl Acad Sci USA 2005;102:7227-32.

178. Castellanos MR, Paramanathan K, El-Sayegh S, Forte F, Buchbinder S, Kleiner M. Breast cancer screening in women with chronic kidney disease: the unrecognized effects of metastatic soft-tissue calcification. Nat Clin Pract Nephrol 2008;4:337-41.

179. Yildiz S, Toprak H, Aydin S, et al. The association of breast arterial calcification and metabolic syndrome. Clinics 2014;69:841-6.

180. Scimeca M, Giannini E, Antonacci C, Pistolese CA, Spagnoli LG, Bonanno E. Microcalcifications in breast cancer: an active phenomenon mediated by epithelial cells with mesenchymal characteristics. BMC Cancer 2014;14:286.

181. Scimeca M, Antonacci C, Colombo D, Bonfiglio R, Buonomo OC, Bonanno E. Emerging prognostic markers related to mesenchymal characteristics of poorly differentiated breast cancers. Tumour Biol 2016;37:5427-35.

182. Hassler O. Microradiographic investigations of calcifications of the female breast. Cancer 1969;23:1103-9.

183. Scott R, Stone N, Kendall C, Geraki K, Rogers K. Relationships between pathology and crystal structure in breast calcifications: an in situ X-ray diffraction study in histological sections. NPJ Breast Cancer 2016;2:16029.

184. Kunitake JAMR, Choi S, Nguyen KX, et al. Correlative imaging reveals physiochemical heterogeneity of microcalcifications in human breast carcinomas. J Struct Biol 2018;202:25-34.

185. Lakhdar A, Daudon M, Mathieu M, Kellum A, Balleyguier C, Bazin D. Underlining the complexity of the structural and chemical characteristics of ectopic calcifications in breast tissues through FE-SEM and μFTIR spectroscopy. Comptes Rendus Chimie 2016;19:1610-24.

186. Haka AS, Shafer-Peltier KE, Fitzmaurice M, Crowe J, Dasari RR, Feld MS. Identifying microcalcifications in benign and malignant breast lesions by probing differences in their chemical composition using Raman spectroscopy. Cancer Res 2002;62:5375-80.

187. Scott R, Kendall C, Stone N, Rogers K. Elemental vs. phase composition of breast calcifications. Sci Rep 2017;7:136.

188. Tandan M, Talukdar R, Reddy DN. Management of pancreatic calculi: an update. Gut Liver 2016;10:873-80.

189. Narasimhulu KV, Gopal NO, Rao JL, et al. Structural studies of the biomineralized species of calcified pancreatic stones in patients suffering from chronic pancreatitis. Biophys Chem 2005;114:137-47.

190. Pitchumoni CS, Viswanathan KV, Gee Varghese PJ, Banks PA. Ultrastructure and elemental composition of human pancreatic calculi. Pancreas 1987;2:152-8.

191. Klimas R, Bennett B, Gardner WA Jr. Prostatic calculi: a review. Prostate 1985;7:91-6.

192. Sutor DJ, Wooley SE. The crystalline composition of prostatic calculi. Br J Urol 1974;46:533-5.

193. Hyun JS. Clinical significance of prostatic calculi: a review. World J Mens Health 2018;36:15-21.

194. Wallingford MC, Benson C, Chavkin NW, Chin MT, Frasch MG. Placental vascular calcification and cardiovascular health: it is time to determine how much of maternal and offspring health is written in stone. Front Physiol 2018;9:1044.

195. Poggi SH, Bostrom KI, Demer LL, Skinner HC, Koos BJ. Placental calcification: a metastatic process? Placenta 2001;22:591-6.

196. Chen KH, Chen LR, Lee YH. Exploring the relationship between preterm placental calcification and adverse maternal and fetal outcome. Ultrasound Obstet Gynecol 2011;37:328-34.

197. Marchiori E, Hochhegger B, Zanetti G. Lymph node calcifications. J Bras Pneumol 2018;44:83.

198. Sakae T, Yamamoto H. Crystals and calcification patterns in two lymph node calcifications. J Oral Pathol 1987;16:456-62.

199. Kim MH, Kim HJ, Kim NN, Yoon HS, Ahn SH. A rotational ablation tool for calcified atherosclerotic plaque removal. Biomed Microdevices 2011;13:963-71.

200. Kaul A, Dhalla PS, Bapatla A, et al. Current treatment modalities for calcified coronary artery disease: a review article comparing novel intravascular lithotripsy and traditional rotational atherectomy. Cureus 2020;12:e10922.

201. Jahnen-Dechent W, Schäfer C, Ketteler M, McKee MD. Mineral chaperones: a role for fetuin-A and osteopontin in the inhibition and regression of pathologic calcification. J Mol Med 2008;86:379-89.

202. Schibler D, Russell RG, Fleisch H. Inhibition by pyrophosphate and polyphosphate of aortic calcification induced by vitamin D3 in rats. Clin Sci 1968;35:363-72.

203. O’Neill WC, Lomashvili KA, Malluche HH, Faugere MC, Riser BL. Treatment with pyrophosphate inhibits uremic vascular calcification. Kidney Int 2011;79:512-7.

204. Wu M, Rementer C, Giachelli CM. Vascular calcification: an update on mechanisms and challenges in treatment. Calcif Tissue Int 2013;93:365-73.

205. Schinke T, Amendt C, Trindl A, Pöschke O, Müller-Esterl W, Jahnen-Dechent W. The serum protein α2-HS glycoprotein/fetuin inhibits apatite formation in vitro and in mineralizing calvaria cells. J Biol Chem 1996;271:20789-96.

206. Heiss A, Jahnen-Dechent W, Endo H, Schwahn D. Structural dynamics of a colloidal protein-mineral complex bestowing on calcium phosphate a high solubility in biological fluids. Biointerphases 2007;2:16-20.

207. Heiss A, DuChesne A, Denecke B, et al. Structural basis of calcification inhibition by α2-HS glycoprotein/fetuin-A. J Biol Chem 2003;278:13333-41.

208. Jahnen-Dechent W, Heiss A, Schäfer C, Ketteler M. Fetuin-A regulation of calcified matrix metabolism. Circ Res 2011;108:1494-509.

209. Zebboudj AF, Imura M, Boström K. Matrix GLA protein, a regulatory protein for bone morphogenetic protein-2. J Biol Chem 2002;277:4388-94.

210. Hruska KA, Mathew S, Saab G. Bone morphogenetic proteins in vascular calcification. Circ Res 2005;97:105-14.

211. Mckee M, Nanci A. Osteopontin at mineralized tissue interfaces in bone, teeth, and osseointegrated implants: Ultrastructural distribution and implications for mineralized tissue formation, turnover, and repair. Microsc Res Tech 1996;33:141-64.

212. McKee MD, Nanci A. Osteopontin: an interfacial extracellular matrix protein in mineralized tissues. Connect Tissue Res 1996;35:197-205.

213. Oikonomaki T, Papasotiriou M, Ntrinias T, et al. The effect of vitamin K2 supplementation on vascular calcification in haemodialysis patients: a 1-year follow-up randomized trial. Int Urol Nephrol 2019;51:2037-44.

214. Eastell R, Walsh JS, Watts NB, Siris E. Bisphosphonates for postmenopausal osteoporosis. Bone 2011;49:82-8.

215. Francis MD, Russell RG, Fleisch H. Diphosphonates inhibit formation of calcium phosphate crystals in vitro and pathological calcification in vivo. Science 1969;165:1264-6.

216. van der Sluis IM, Boot AM, Vernooij M, Meradji M, Kroon AA. Idiopathic infantile arterial calcification: clinical presentation, therapy and long-term follow-up. Eur J Pediatr 2006;165:590-3.

217. Persy V, De Broe M, Ketteler M. Bisphosphonates prevent experimental vascular calcification: treat the bone to cure the vessels? Kidney Int 2006;70:1537-8.

218. Shiraishi N, Kitamura K, Miyoshi T, et al. Successful treatment of a patient with severe calcific uremic arteriolopathy (calciphylaxis) by etidronate disodium. Am J Kidney Dis 2006;48:151-4.

219. Jansen RS, Duijst S, Mahakena S, et al. ABCC6-mediated ATP secretion by the liver is the main source of the mineralization inhibitor inorganic pyrophosphate in the systemic circulation-brief report. Arterioscler Thromb Vasc Biol 2014;34:1985-9.

220. Li Q, Huang J, Pinkerton AB, et al. Inhibition of tissue-nonspecific alkaline phosphatase attenuates ectopic mineralization in the Abcc6(-/-) mouse model of PXE but not in the enpp1 mutant mouse models of GACI. J Invest Dermatol 2019;139:360-8.

221. Pomozi V, Brampton C, van de Wetering K, et al. Pyrophosphate supplementation prevents chronic and acute calcification in ABCC6-deficient mice. Am J Pathol 2017;187:1258-72.

222. Zhao J, Kingman J, Sundberg JP, Uitto J, Li Q. Plasma PPi deficiency is the major, but not the exclusive, cause of ectopic mineralization in an Abcc6(-/-) mouse model of PXE. J Invest Dermatol 2017;137:2336-43.

223. Garti N, Tibika F, Sarig S, Perlberg S. The inhibitory effect of polymeric carboxylic amino-acids and urine on calcium oxalate crystallization. Biochem Biophys Res Commun 1980;97:1154-62.

224. Kleinman JG, Alatalo LJ, Beshensky AM, Wesson JA. Acidic polyanion poly(acrylic acid) prevents calcium oxalate crystal deposition. Kidney Int 2008;74:919-24.

225. Worcester EM, Blumenthal SS, Beshensky AM, Lewand DL. The calcium oxalate crystal growth inhibitor protein produced by mouse kidney cortical cells in culture is osteopontin. J Bone Miner Res 1992;7:1029-36.

226. Wesson JA, Johnson RJ, Mazzali M, et al. Osteopontin is a critical inhibitor of calcium oxalate crystal formation and retention in renal tubules. J Am Soc Nephrol 2003;14:139-47.

227. Cohen GF, Vyas NS. Sodium thiosulfate in the treatment of calciphylaxis. J Clin Aesthet Dermatol 2013;6:41-4.

228. Guerra G, Shah RC, Ross EA. Rapid resolution of calciphylaxis with intravenous sodium thiosulfate and continuous venovenous haemofiltration using low calcium replacement fluid: case report. Nephrol Dial Transplant 2005;20:1260-2.

229. Cicone JS, Petronis JB, Embert CD, Spector DA. Successful treatment of calciphylaxis with intravenous sodium thiosulfate. Am J Kidney Dis 2004;43:1104-8.

230. Yatzidis H. Successful sodium thiosulphate treatment for recurrent calcium urolithiasis. Clin Nephrol 1985;23:63-7.

231. Sun XY, Xu M, Ouyang JM. Effect of crystal shape and aggregation of calcium oxalate monohydrate on cellular toxicity in renal epithelial cells. ACS Omega 2017;2:6039-52.

232. Bazin D, Chevallier P, Matzen G, Jungers P, Daudon M. Heavy elements in urinary stones. Urol Res 2007;35:179-84.

233. Meyer J, Angino E. The role of trace metals in calcium urolithiasis. Invest Urol 1977;14:347-50.

234. Muñoz JA, Valiente M. Effects of trace metals on the inhibition of calcium oxalate crystallization. Urol Res 2005;33:267-72.

235. Atakan IH, Kaplan M, Seren G, Aktoz T, Gül H, Inci O. Serum, urinary and stone zinc, iron, magnesium and copper levels in idiopathic calcium oxalate stone patients. Int Urol Nephrol 2007;39:351-6.

236. Levy RJ, Schoen FJ, Flowers WB, Staelin ST. Initiation of mineralization in bioprosthetic heart valves: studies of alkaline phosphatase activity and its inhibition by AlCl3 or FeCl3 preincubations. J Biomed Mater Res 1991;25:905-35.

237. Ali SY. Apatite-type crystal deposition in arthritic cartilage. Scan Electron Microsc 1985;Pt 4:1555-66.

238. Usai D, Maritan L, Dal Sasso G, et al. Late pleistocene/early holocene evidence of prostatic stones at Al khiday cemetery, central sudan. PLoS One 2017;12:e0169524.

Cite This Article

Export citation file: BibTeX | RIS

OAE Style

Mu Y, Gao W, Zhou Y, Xiao L, Xiao Y. Physiological and pathological/ectopic mineralization: from composition to microstructure. Microstructures 2023;3:2023030. http://dx.doi.org/10.20517/microstructures.2023.05

AMA Style

Mu Y, Gao W, Zhou Y, Xiao L, Xiao Y. Physiological and pathological/ectopic mineralization: from composition to microstructure. Microstructures. 2023; 3(4): 2023030. http://dx.doi.org/10.20517/microstructures.2023.05

Chicago/Turabian Style

Mu, Yuqing, Wendong Gao, Yinghong Zhou, Lan Xiao, Yin Xiao. 2023. "Physiological and pathological/ectopic mineralization: from composition to microstructure" Microstructures. 3, no.4: 2023030. http://dx.doi.org/10.20517/microstructures.2023.05

ACS Style

Mu, Y.; Gao W.; Zhou Y.; Xiao L.; Xiao Y. Physiological and pathological/ectopic mineralization: from composition to microstructure. Microstructures. 2023, 3, 2023030. http://dx.doi.org/10.20517/microstructures.2023.05

About This Article

© The Author(s) 2023. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Data & Comments

Data

Views
600
Downloads
95
Citations
2
Comments
0
2

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.

0
Download PDF
Cite This Article 17 clicks
Like This Article 2 likes
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Microstructures
ISSN 2770-2995 (Online)
Follow Us

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/