Next Article in Journal
The Intestinal Immune Defense System in Insects
Next Article in Special Issue
Proteomic and Transcriptomic Responses Enable Clams to Correct the pH of Calcifying Fluids and Sustain Biomineralization in Acidified Environments
Previous Article in Journal
Breast Cancer Vaccine Containing a Novel Toll-like Receptor 7 Agonist and an Aluminum Adjuvant Exerts Antitumor Effects
Previous Article in Special Issue
Fabrication of Multilayered Biofunctional Material with an Enamel-like Structure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Shedding Light on the Role of Na,K-ATPase as a Phosphatase during Matrix-Vesicle-Mediated Mineralization †

by
Heitor Gobbi Sebinelli
1,
Luiz Henrique Silva Andrilli
1,
Bruno Zoccaratto Favarin
1,
Marcos Aantonio Eufrasio Cruz
1,
Maytê Bolean
1,
Michele Fiore
2,
Carolina Chieffo
2,
David Magne
2,
Andrea Magrini
3,
Ana Paula Ramos
1,
José Luis Millán
4,
Saida Mebarek
2,
Rene Buchet
2,
Massimo Bottini
1,5,* and
Pietro Ciancaglini
1,5,*
1
Departamento de Química, Faculdade de Filosofia, Ciências e Letras de Ribeirão Preto da Universidade de São Paulo (FFCLRP-USP), Ribeirão Preto, São Paulo 14040-900, Brazil
2
University Lyon, Université. Claude Bernard Lyon 1, CNRS UMR 5246, ICBMS, F-69622 Lyon, France
3
Department of Biomedicine and Prevention, University of Rome Tor Vergata, 00133 Rome, Italy
4
Sanford Burnham Prebys, La Jolla, CA 92037, USA
5
Department of Experimental Medicine, University of Rome Tor Vergata, 00133 Rome, Italy
*
Authors to whom correspondence should be addressed.
The authors dedicated this article in memoriam of Dr. Mathieu Borel, friend and researcher from Univ. Lyon 1.
Int. J. Mol. Sci. 2022, 23(23), 15072; https://doi.org/10.3390/ijms232315072
Submission received: 12 October 2022 / Revised: 26 November 2022 / Accepted: 29 November 2022 / Published: 1 December 2022
(This article belongs to the Special Issue Organic, Inorganic and Natural Molecules in Biomineralization)

Abstract

:
Matrix vesicles (MVs) contain the whole machinery necessary to initiate apatite formation in their lumen. We suspected that, in addition to tissue-nonspecific alkaline phosphatase (TNAP), Na,K,-ATPase (NKA) could be involved in supplying phopshate (Pi) in the early stages of MV-mediated mineralization. MVs were extracted from the growth plate cartilage of chicken embryos. Their average mean diameters were determined by Dynamic Light Scattering (DLS) (212 ± 19 nm) and by Atomic Force Microcopy (AFM) (180 ± 85 nm). The MVs had a specific activity for TNAP of 9.2 ± 4.6 U·mg−1 confirming that the MVs were mineralization competent. The ability to hydrolyze ATP was assayed by a colorimetric method and by 31P NMR with and without Levamisole and SBI-425 (two TNAP inhibitors), ouabain (an NKA inhibitor), and ARL-67156 (an NTPDase1, NTPDase3 and Ecto-nucleotide pyrophosphatase/phosphodiesterase 1 (NPP1) competitive inhibitor). The mineralization profile served to monitor the formation of precipitated calcium phosphate complexes, while IR spectroscopy allowed the identification of apatite. Proteoliposomes containing NKA with either dipalmitoylphosphatidylcholine (DPPC) or a mixture of 1:1 of DPPC and dipalmitoylphosphatidylethanolamine (DPPE) served to verify if the proteoliposomes were able to initiate mineral formation. Around 69–72% of the total ATP hydrolysis by MVs was inhibited by 5 mM Levamisole, which indicated that TNAP was the main enzyme hydrolyzing ATP. The addition of 0.1 mM of ARL-67156 inhibited 8–13.7% of the total ATP hydrolysis in MVs, suggesting that NTPDase1, NTPDase3, and/or NPP1 could also participate in ATP hydrolysis. Ouabain (3 mM) inhibited 3–8% of the total ATP hydrolysis by MVs, suggesting that NKA contributed only a small percentage of the total ATP hydrolysis. MVs induced mineralization via ATP hydrolysis that was significantly inhibited by Levamisole and also by cleaving TNAP from MVs, confirming that TNAP is the main enzyme hydrolyzing this substrate, while the addition of either ARL-6715 or ouabain had a lesser effect on mineralization. DPPC:DPPE (1:1)-NKA liposome in the presence of a nucleator (PS-CPLX) was more efficient in mineralizing compared with a DPPC-NKA liposome due to a better orientation of the NKA active site. Both types of proteoliposomes were able to induce apatite formation, as evidenced by the presence of the 1040 cm−1 band. Taken together, the findings indicated that the hydrolysis of ATP was dominated by TNAP and other phosphatases present in MVs, while only 3–8% of the total hydrolysis of ATP could be attributed to NKA. It was hypothesized that the loss of Na/K asymmetry in MVs could be caused by a complete depletion of ATP inside MVs, impairing the maintenance of symmetry by NKA. Our study carried out on NKA-liposomes confirmed that NKA could contribute to mineral formation inside MVs, which might complement the known action of PHOSPHO1 in the MV lumen.

1. Introduction

Biomineralization consists of the accumulation of minerals, composed mainly of phosphate and calcium ions that form hydroxyapatite (Ca10(PO4)6(OH)2) or related carbonated apatites when the hydroxyls are substituted by other anions and propagate them onto the collagenous matrix. The ossification process mediated by osteoblasts (in the formation of flat bones) or by odontoblasts (in tooth formation) is clearly distinct from that which occurs during the calcification of epiphyseal cartilage (endochondral formation) which is mediated by hypertrophic chondrocytes [1]. In all of these tissues, mineralization depends on the homeostasis of calcium (Ca2+) and inorganic phosphate (Pi) ions, the adequate protein composition of the extracellular matrix, and the absence and/or removal of mineralization inhibitors [1]. The initiation of mineral formation by matrix vesicles (MVs) in the growth plate cartilage is supported by several pieces of evidence [2,3,4,5,6,7,8,9,10,11]. (i) In the early 1960s, lipids were discovered at the sites of calcification in the growth plate [12,13,14,15]. Later, it was evidenced by transmission electron microscopy that the lipids formed vesicular structures, with diameters in the 100–300 nm range [16,17,18,19], named MVs [2] that were able to bind strongly to collagen [20,21]. (ii) MVs were shown to be released from the polarized apical side of plasma membranes of hypertrophied chondrocytes [22]. (iii) and to possess the ability to accumulate calcium and phosphate ions in their lumen [23,24,25,26,27,28,29], which can nucleate [30,31,32,33] and form apatites [29]. (iv) Once released from MVs, either by mechanical force and/or degradation of phospholipids [22,34,35,36,37,38], these apatitic crystals continue to grow onto collagen fibers in the extracellular matrix [2,3]. MVs released from osteoblasts and odontoblasts have similar properties and are also able to initiate apatite within their lumen [2,3,10].
MVs harbor the complete machinery necessary to sustain the supply of Pi in the lumen of MVs and to control the homeostasis of extracellular inorganic pyrophosphate (PPi) [2,3,4,5,6,7] (Figure 1). PPi inhibits the formation of apatite crystals [39] as evidenced by hypomorphic mutations of tissue-nonspecific alkaline phosphatase (TNAP) gene (ALPL), which impair PPi hydrolysis and lead to the accumulation of extracellular PPi that cause the soft bones disease known as hypophosphatasia [40,41]. The function of MVs is to accumulate Pi and Ca2+ inside their lumen, which provides an optimum environment to nucleate and form apatite and to release apatite in the extracellular medium, where the PPi has to be removed to sustain the second step of mineral formation [2,3,4,5,6,7,10] (Figure 1). High levels of TNAP, the enzyme responsible for PPi hydrolysis, have been found in osteoblast membranes during bone tissue mineralization [2,42,43,44] and in MVs [28,45,46]. The enzyme is allosterically regulated by ATP [47] and competitively inhibited by the reaction product, Pi [48], suggesting that the relative levels of PPi, an inhibitor of biomineralization, and Pi, a promoter of mineralization but also an inhibitor of TNAP, present in the matrix extracellular fluid also play an important role in regulating the physiological mineralization process.
Another extracellular phosphatase is Nucleotide Pyrophosphataes/Phosphodiesterase 1(NPP1), which produces PPi and Pi from ATP [39] (Figure 1). NPP1, also known as glycoprotein-1 (PC-1), is bound to the plasma membrane, while another member of the NPP family, autotaxin (NPP2), is secreted, and B10 (NPP3) is abundant in intracellular spaces [39]. These three isoenzymes are expressed in a wide variety of tissues, including bone and cartilage [40], and all have the common ability to hydrolyze diesters of phosphoric acid to phosphomonoesters, primarily ATP to AMP and/or ADP to adenosine [41]. NPP1, which has its active site facing the extracellular milieu, is found in high concentrations on the surfaces of osteoblasts and chondrocytes, as well as in the membranes of their MVs [42,43] (Figure 1). NPP1 inhibits apatite precipitation by its PPi-generating property. This hypothesis has been confirmed by in vitro studies, where cells transfected with the NPP1 cDNA demonstrated high levels of PPi in osteoblast-derived MVs, accompanied by reduced matrix mineralization [43,44]. Abnormal precipitation of Ca2+ pyrophosphate dehydrate (CPPD) has been observed in association with TNAP deficiency. While excess PPi in patients with hypophosphatasia primarily leads to rickets and osteomalacia, patients with the disease can also develop the pathological formation of CPPD crystals in articular cartilage [45]. NPP1 also acts as a phosphatase in the absence of TNAP [41,46,47,48,49]. The nucleoside triphosphate diphosphohydrolase 1 (CD39) also hydrolyses extracellular ATP or ADP forming AMP with the release of one Pi ion [50].
On the other hand, Pi needs to be concentrated or generated to sustain the formation of apatitic minerals inside MVs. Pi is thought to accumulate into MV via a Na+-dependent Pi transport system [26], although the substitution of Li+ or K+ for Na+ has a minimal effect [27]. One likely Pi transporter could be the phosphate transporter 1 (PiT-1) [28] (Figure 1). Alternatively, there are enzymes that hydrolyze phosphoesters, including phosphocholine and phosphoethanolamine, inside MVs to generate Pi. High levels of PHOSPHO1 in MVs isolated from chondrocytes were revealed by immunoblotting [51] (Figure 1). By comparing wild-type and Phospho1−/− mice, it has been shown that the function of PHOSPHO1, another prominent phosphatase present in MVs, is to produce intraluminal Pi and induce apatite mineral [52]. Nevertheless, the gradient of Na+ generated by Na,K-ATPase (NKA), along with its own ATP hydrolysis, could be theoretically sufficient to produce a burst of Pi inside of the MVs to form the nucleational core, using the auxiliary action of PiT-1 cotransporter [53,54] (Figure 1).
NKA belongs to the P-Type-II ATPase family; it is ubiquitous at the cellular level for animal tissues and vital for these organisms [55]. The functions performed by the protein in the cells are related to the maintenance of the ionic composition of the cytosolic medium [56]. Thus, several tissues use NKA to carry out the transport of three Na+ ions out of the cell and two K+ ions into the cell through the plasma membrane with a perpetual energy expenditure of one ATP molecule [57]. This uneven movement of ionic transport generates an electrolytic gradient that allows the cell to reuse energy by propelling secondary transport of molecules such as sugars, neurotransmitters, metabolite amino acids, and other H+, Ca2+, Cl, and other ions [57].
To assess NKA contribution in the production of luminal Pi in MVs, ATP hydrolysis of MVs extracted from epiphyseal cartilage from femurs of chicken embryos was determined on solubilized MVs by using Levamisole (a TNAP inhibitor of low specificity), and SBI-425 (a TNAP inhibitor of high specificity) [58,59] ouabain (an NKA inhibitor) and without any inhibitors. The ability to induce apatite formation was assessed in native MVs and in MVs devoid of TNAP through cleavage of their GPI anchor moiety. Additionally, NKA reconstituted in liposomes was used as a biomimetic model to study the role of NKA during mineral formation. In this respect, the advantage of a proteoliposome system is to have control over lipid and protein composition and observe NKA activity in isolation [60,61]. NKA was solubilized by using a non-ionic detergent and reconstituted in liposomes containing a 1:1 (w/w) mixture of DPPC:DPPE in a 1:3 (w/w) lipid:protein ratio resulting in proteoliposomes with 89% recovery of incorporated protein, with a mean diameter of 140 nm and 79% recovered ATPase activity. The reconstituted NKA has the ATP hydrolysis site located on the outside of the proteoliposome vesicles, called inside-out orientation [62,63]. The inside-out proteoliposomes allow a series of biophysical studies on enzyme stability and structure. Although the functional unit of NKA in vivo is not yet clearly elucidated, it is known that the oligomerization process occurs in natural membranes, and it cannot be ruled out that this behavior may be a natural mechanism developed by organisms to act in the metabolic regulation of various functions in the plasma membrane of cells [64,65]. Atomic force microscopy (AFM) was applied in intermittent contact mode to obtain morphological and topographical information on how NKA is organized in proteoliposomes composed of DPPC and DPPC:DPPE (1:1 molar ratio), to assess the accessibility of the active site and to optimize mimetic models to match their behavior of biological systems [66].

2. Results and Discussion

2.1. Characterization of MV

Twenty-five different fresh MV samples were extracted, as reported by [66]. The average protein concentration of MVs was around 2.9 ± 0.9 mg·mL−1. The specific TNAP activity of MVs was 9.2 ± 4.6 U·mg−1, which was significantly higher than that of plasma membranes of chondrocytes which amounted to 5.1 ± 3.6 U·mg−1, consistent with the significant reported enrichment of TNAP in MVs compared to plasma membranes [66]. The mean diameter of 10 distinct samples of MVs was 212 ± 19 nm, as revealed by DLS, which falls within the reported 100–300 nm diameter range [16,17,18,19]. However, DLS analysis revealed a higher polydispersity index (PI) than the ideal for a monodisperse solution, suggesting the presence of a mixture of different-sized vesicles (Figure 1). A mean diameter of 180 ± 85 nm was obtained (Figure 2), in accordance with that revealed by Dynamic Light Scattering (DLS) analysis and with AFM values found in the literature [29]. This technique allows the investigation of surface morphology from nanostructures or microdomains composed of lipids and proteins. The phase contrast shift became more evident when using synthetic vesicles such as liposomes and proteoliposomes but was not evident in the MV surface [29,66]. The sum of the infinite interferences arising from the unlimited variety of proteins and lipids probably dilutes the equipment signal and reduces the visualization of structures on the surface. Thus, in the phase images, the MVs have low-range phase shifts (Figure 2B, color scale). Even protrusions, a characteristic of the several proteins present in the membrane, are difficult to be observed (Figure 2A, 3D topographic image).

2.2. ATP Hydrolysis by MV

In order to comprehend the role of NKA in MVs, ATP hydrolysis was measured by 31P NMR and a colorimetric assay. To determine the ATPase activity inside and outside MVs, the MVs were solubilized by the addition of 0.2% of C12E8 or Nonidet P-40. MVs without detergent have a mean diameter of 206 nm and PI of 0.14, while MVs in the presence of the detergent resulted in a diameter of 79.9 nm and PI of 1.82. Therefore, MVs were destabilized by the addition of the detergent, exposing ATP to all regions of MVs and enabling the measurement of ATP hydrolysis inside MVs. The addition of 2 mM of ouabain was sufficient to inhibit 90% of the total activity of NKA and 99.9% of the activity when used at 5.0 mM [62,63]. The addition of 3 mM of ouabain in MVs resulted in a very small decrease in total ATP hydrolysis to around 3–7% (Table 1), which was consistent with the small ATP hydrolysis by ATPase in MVs using [γ-32P]-ATP and Pi/molibdate complex determination [53,67].
Levamisole and SBI-425 were more efficient in inhibiting ATP hydrolysis as they act mainly and exclusively, respectively, on TNAP activity (Table 1). The addition of 5 mM of Levamisole, the stereoisomer of tetramisole, a potent uncompetitive inhibitor [68], led to a decrease of around 72–69% of the total ATP hydrolysis by MVs (Table 2) compared to the control experiment. SBI-425 is a potent pharmacological uncompetitive TNAP inhibitor that, unlike Levamisole, acts specifically on TNAP without cross-inhibition of other phosphatases [58,59]. Two concentrations of SBI-425 (5 μM and 10 μM), at much lower concentrations than those used for Levamisole, resulted in inhibitions of 21.5 and 35.4%, respectively (Table 1), which confirmed that TNAP is hydrolyzing ATP.
6-N,N-Diethyl-D-beta-gamma-dibromomethylene adenosine triphosphate (ARL-67156) is a competitive inhibitor of the ectoenzymes NTPDase1 and NTPDase3 along with NPP1 [70]. The inhibition of CD39 (an NTPDase1) activity has been reported to induce the osteogenic potential of mesenchymal stem cells derived from gingival tissue by regulating the balance between osteoclasts and osteoblasts through the Wnt/β-Catenin pathway in an osteopenic model induced by ovariectomies in mice, and therefore such inhibition has potential therapeutic value for osteoporosis [71,72]. The addition of 0.1 mM of ARL 67156 inhibited 8–13.7% of the total ATP hydrolysis in MVs, which indicated that TNAP was more effective in hydrolyzing ATP than the other ectoenzymes.

2.3. Mineral Formation Monitored by Turbidity

For the determination of the influence of phosphatases other than GPI-anchored TNAP, a phosphatidylinositol phospholipase C (PI-PLC) treatment that can remove TNAP activity from MVs (66–92%) [73,74] was used. The use of PI-PLC resulted in the partial cleavage of TNAP and generated cTNAP (TNAP released from MVs) and cMVs (MVs without GPI-anchored TNAP). The integrity of the vesicles was preserved, although a minor increase in the diameter was observed in cMVs (286 ± 61 nm) compared to native MVs (217 ± 23 nm) after the PI-PLC treatment and ultracentrifugation process. The percentage of TNAP activity removed was around 66%, which was consistent with previous reports in the literature [75] (Table 2).
Table 2. Biochemical and biophysical characterization of MV; cMV and cTNAP obtained after treatment with PI-PLC as described in Materials and Methods. MV were treated with PI-PLC during 3 h, ultracentrifuged at 80,000× g. The supernatant contained cTNAP, and the pellet composed of cMV was resuspended in SCL without phosphate. Protein determination was assayed by the Bradford method [76]. TNAP activity in MVs isolated from chondrocytes, was measured continuously using para-nitrophenyl phosphate (pNPP, ε = 17,600 M−1·cm−1, pH 13.0, 1.0 M) [75]. Vesicles sizes were determined by Dynamic light scattering (DLS) at 25 °C using the unimodal distribution.
Table 2. Biochemical and biophysical characterization of MV; cMV and cTNAP obtained after treatment with PI-PLC as described in Materials and Methods. MV were treated with PI-PLC during 3 h, ultracentrifuged at 80,000× g. The supernatant contained cTNAP, and the pellet composed of cMV was resuspended in SCL without phosphate. Protein determination was assayed by the Bradford method [76]. TNAP activity in MVs isolated from chondrocytes, was measured continuously using para-nitrophenyl phosphate (pNPP, ε = 17,600 M−1·cm−1, pH 13.0, 1.0 M) [75]. Vesicles sizes were determined by Dynamic light scattering (DLS) at 25 °C using the unimodal distribution.
SampleProtein
(mg·mL−1)
pNPP
(U·mg−1)
Diameter (nm)PI
MV (native) 3.7 ± 0.34.4 ± 0.9217 ± 230.4
cMV (partial cleaved TNAP) 1.9 ± 0.31.5 ± 0.3286 ± 610.5
cTNAP (cleaved from MV) 0.4 ± 0.045.1 ± 1.8----
SBI-425 inhibited a maximum of 33.4% of the total activity of the cTNAP, removed from the MV. This value is in agreement with the results of the experiments carried out previously with ATP at 3.0 mM as a substrate (Table 1) but less than expected based on the inhibition properties of human and mouse TNAP. The active site pocket of chicken TNAP has a Glutamine at position 434, rather than a Histidine present in human and mouse TNAPs, and this substitution explains the lower sensitivity of chicken TNAP to inhibition by SBI-425, as well as Levamisole [77]. In this sense, the inhibitory effect is not observed in MV since it has a high content of TNAP, while in cMV, its presence only delays mineralization (Figure 3).
In order to evaluate if NKA could affect mineralization, turbidity assays were performed to monitor the formation of calcium–phosphate precipitation by measuring the absorbance at 340 nm. ATP, as PPi, is a strong inhibitor of mineral formation and can delay the induction time of mineralization [74,78,79,80]. The addition of 3.42 mM of ATP delayed the starting induction time to around 20 h due to the time needed to hydrolyze ATP and to initiate mineralization (Figure 3A, control black symbols).
The control experiment (curve without inhibitor) revealed an increase in turbidity with a sigmoidal shape. It started (starting time, ti) after 20 h of incubation and ended at 50.1 Abs340nm·mg−1 (Umax) at the end of the 120 h of the experiment (tf of Table 3). The addition of Levamisole (which inhibits TNAP-mediated ATP hydrolysis) almost eliminated the induction of mineralization due to the presence of excess ATP (conserved by the lack of ATPase activity) and, mostly, the absence of Pi (Figure 3A, red). The addition of ouabain with or without SBI-425 (Figure 3A, violet and green symbols) and SBI-425 alone (Figure 3A, cyan) slightly affected the starting induction time, which remained around 14–20 h (Table 3). However, the hydrolysis of ATP was not completely inhibited by ouabain and/or SBI-425 as compared to the control, since their tmax rate (58.8 to 74.4 h), Umax (Abs340nm·mg−1) (54.6–71.6) PMP (0.73–1.15 h−1) were almost similar and even slightly higher compared to the control values (tmax rate= 57.9 ± 0.8 h; Umax (Abs340nm·mg−1) = 50.1 ± 0.8 and PMP = 0.86 h−1). This indicated that an amount of ATP was sufficiently hydrolyzed even in the presence of ouabain and/or SBI-425.
The binding site pocket for the uncompetitive inhibitor SBI-425 is not perfectly conserved for chicken TNAP, as compared to that in human or mouse TNAP, due to the functionally significant H434Q substitution [77]. Furthermore, ATPase activity assays in Table 1 showed that 10 μM SBI-425 inhibited about 35% of MV total activity, whereas 5 mM Levamisole inhibited 70%. In this sense, the inhibitory effect cannot be observed in MV since there is a high abundance of TNAP that cannot be totally inhibited by SBI-425. Other phosphatases act as a backup mechanism for Pi production, in a manner analogous to the documented phosphatase activity of ENPP1 rather than its more traditional pyrophosphohydrolase activity, in the complete absence of TNAP activity [46].
Levamisole efficiently prevented mineral propagation. The inhibition is attributed mostly to TNAP. Chicken MVs do not have the same susceptibility to SBI-425 as for Levamisole since it appears that SBI-425 does not completely inhibit TNAP from chicken embryos, probably due to a distinct topology of the inhibition site of TNAP from chicken embryos as compared to that of human TNAP. Ouabain did not show a great effect on mineralization since the contribution of NKA compared to the hydrolysis of TNAP by MV was around 5% or less, and/or the ATP binding site of NKA would be mostly occluded inside of the vesicle. In order to verify the main role of TNAP in hydrolyzing ATP, the ability to hydrolyze ATP in cMVs was depleted by around 65% of TNAP, as compared to native MV. In Figure 3B, all mineralization curves associated with cMVs presented a delay of about 40 h (ti in Table 3) when compared directly with the native MVs, but they retained the sigmoidal behavior.
The delay corresponds to the step in the reaction, where ATP substrate and its derivatives were being hydrolyzed; however, there is still not enough Pi concentration for mineral formation. The addition of Levamisole completely abolished the mineralization. Even after removing 65% of TNAP (Figure 3B, Table 3), it was still impossible to observe any ouabain effect on the propagation and mineralization. The combined effect of Levamisole and ouabain could be explained in native MVs (Figure 3A) by the effect of Levamisole on TNAP activity since ouabain does not reach the binding site of NKA. The little increment in turbidimetry, when compared with Levamisole alone, could be explained by unspecific interaction and reducing the effect of Levamisole. In the cMV (Figure 3B), this effect was not observed since the contribution of TNAP activity was reduced by 70%.

2.4. Na,K-ATPase Proteoliposomes as MV Biomimetic Models

Due to the presence of several types of phosphatases in MVs and to the occluded NKA active site (not exposed to the external medium), it was unclear if NKA itself could contribute to Pi production inside MVs and to subsequent mineralization. In order to address this question, DPPC:DPPE proteoliposomes were constructed containing solubilized NKA from membrane fractions prepared from rabbit kidneys [62]. The concentration found for the solubilized NKA was 0.19 mg·mL−1, and its catalytic activity was 0.336 U·mg−1 (Table 4).
The hydrodynamic diameter of the enzyme is 19 ± 1 nm (Table 4), obtained by the DLS analysis [64]. The crystal structures of the αβ and (αβ)2 conformations have a hydrodynamic radius of 5.0 and 6.3 nm (according to PDBs 3WGV and 3A3Y for the monomer and 3WGV for the dimer). Therefore, there is a balance between monomers, oligomers, and aggregates in solution as determined by analytical centrifugation [64].
Taking into account the mean diameter of the MVs and that of NKA, it is possible to determine the spatial contribution of the enzyme. The NKA diameters comprise 10, 5, and 6.3% of the mean diameter of the MVs when calculated by DLS or by crystallography, respectively. However, when these values are evaluated in surface area, 1134 nm2 is determined by the average found by DLS, 314 nm2 for the αβ form, 499 nm2 for the (αβ)2 form against 125,667 nm2 for the MVs with 200 nm of diameter. Therefore, in these calculations, the contribution of NKA to the total area of MVs varies from 0.9 to 0.25% of the area of MVs, which explains the difficulty in finding them during topographic analysis by AFM [66,81].
Mineral propagation parameters for TNAP and NPP1 liposomes in different lipid compositions indicated that the best results were obtained for the DPPC compositions [41,82], consistent with the presence of 41.8% phosphatidylcholine relative to the total phospholipid found in MVs [6]. The second largest phospholipid type is phosphatidylethanolamine which amounted to 14.9% of the total phospholipid composition in MVs [6]. The lipid composition can influence the catalytic activities of the enzymatic machinery. Another advantage of DPPE is that it induces an inside-out orientation, with the NKA active site facing the outside of the vesicles [63]. In this configuration, the active site of the enzyme is in an inside-out position and has free access to ATP from the medium [63,66]; moreover, as indicated by the similar activity values for the NKA (0.336 U·mg−1) and NKA-DPPC:DPPE (0.29 U·mg−1, corresponding to 86% of NKA activity) in contrast to NKA:DPPC (0.141 U·mg−1, corresponding to 61% of NKA activity), consistent with previous findings [63,66] (Table 4).
The mean diameter of DPPC liposomes measured by DLS was 365 ± 35 nm, while its polydispersity index (PI) was 0.7 (Table 4). The DPPC:DPPE liposome mean diameter was 504 ± 50 nm with a PI of 1.3 (Table 4). Such size discrepancies may reflect the nature of phospholipids and their respective PI values [66,83]. The NKA-DPPC has a mean diameter of 634 ± 60 nm, while the NKA-DPPC:DPPE has an average diameter of 899 ± 90 nm (Table 4). In order to induce mineral formation, a phosphtidylserine complex containing calcium and phosphate (PS-CPLX) was added since the enzyme alone was not able to nucleate. The solubilized protein, by hydrolyzing ATP (51.4 Abs340nm·mg−1), produced turbidity in the presence of PS-CPLX nucleators and Ca2+ ions, which amounted to 51.4 Abs340nm·mg−1 and had the same propagation capacity as NKA-DPPC:DPPE which was around 55.8 Abs340nm·mg−1 (Table 4). DPPC-NKA liposome by hydrolyzing ATP in a medium containing CPLX nucleators and Ca2+ ions induced a smaller nucleation profile which was 38.4 Abs340nm·mg−1 than the NKA-DPPC:DPPE. This is due to the more favorable orientation of the active site of NKA facing outside in NKA-DPPC:DPPE (90% inside-out orientation) as compared to that of NKA:DPPC (61%). Larger amounts of ATP are hydrolyzed by NKA-DPPC:DPPE than by NKA-DPPC. Other factors, such as interactions between the phospholipid DPPE and the DPPS used in the preparation of the PS-CPLX nucleator, cannot be completely ruled out. The inhibition of NKA by ouabain in proteoliposomes and in this assay condition has been extensively discussed by our group [62,63,84,85,86,87]. Nevertheless, the addition of ouabain did not completely prevent the ATP nucleation process, probably due to the occluded ouabain binding site, which turned to the lumen of proteoliposomes, but even the addition of detergent was not able to improve the inhibition and abrogate mineral formation.

2.5. Spectroscopic Analysis of Minerals Formed by Na, K-ATPase-Liposomes

The ATP-FTIR spectra of the minerals formed from the incubation of NKA-DPPC (Figure 4A1–A3) and of NKA-DPPC:DPPE (Figure 4B1–B3) have similar profiles. The remaining activity of the NKA-liposomes after the inhibition by ouabain did not affect the mineral quality, even after the addition of C12E8 to solubilize the vesicles. That is, the residual activity was sufficient to generate phosphate and mineral formation. The 1040 cm−1 band (as indicated by the dashed yellow line) could correspond to apatite, suggesting that the proteoliposome formed by NKA and either DPPC or DPPE:DPPE could promote mineral formation.

3. Materials and Methods

3.1. Materials

Solutions were prepared with Millipore Direct-Q ultrapure apyrogenic water. All the reagents were of the highest commercial purity available. Bovine serum albumin (BSA), Collagenase type 1 (clostridium histolyticum), Tris[hydroxymethyl]aminomethane (Tris), N-(2-hydroxyethyl) piperazine-N’-ethanesulfonic acid (HEPES), adenosine 5′-Triphosphate tris salt (ATP), dodecyloctaglycol (C12E8), Glutaraldehyde (Grade I, specially purified for use as an electron microscopy fixative), Octylphenoxy poly(ethyleneoxy)ethanol, (branched), p-Nitrophenyl phosphate disodium hexahydrate, Phospholipase-C (bacillus cereus), Levamisole (L(−)-2,3,5,6-Tetrahydro-6- phenylimidazo [2,1-b]thiazole hydrochloride), ouabain (1β,3β,5β,11α,14,19-Hexahydroxycard- 20(22)-enolide 3-(6-deoxy-α-L-mannopyranoside)), ARL-67156 trisodium salt, a selective inhibitor of ecto-ATPase, D-glucose and sucrose were purchased from Sigma Chemical Co. (St. Louis, MO). 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) and 1,2-dipalmitoyl-sn-glycero-3-phosphoethanolamine (DPPE) were acquired from Avanti Polar Lipids. Ethylenediaminetetracetic acid (EDTA), magnesium chloride (MgCl2), potassium chloride (KCl), sodium chloride (NaCl), sodium hydrogen carbonate (NaHCO3), di-sodium sulfate (Na2SO4) and trichloroacetic acid (TCA) were from Merck. Biobeads and Bradford reagents were acquired from Bio-Rad. The TNAP inhibitor SBI-425 was kindly provided by Professor Dr. J.L. Milan of the Sanford Health Institute in La Jolla, California.

3.2. Matrix Vesicles Isolated from Chicken Embryo Femurs

The MVs were prepared with a collagenase digestion step [75]. Briefly, 20 chicken embryos (17 days after fertilization) were sacrificed by decapitation. The femurs were dissected, and slices (1–3 mm thick) of the epiphyses/growth plates finely cut and digested for 3 h, at 37 °C, in synthetic cartilage buffer (SCL) buffer supplemented with 1 mM CaCl2 and type I collagenase, concentration of 300 U/g of tissue. SCL, a buffer that mimics the native environment of MVs in the cartilage, was composed of 1.83 mM NaHCO3, 12.7 mM KCl, 0.57 mM MgCl2, 5.55 mM D-glucose, 63.5 mM sucrose, 16.5 mM Tris (2-Amino-2-hydroxymethyl-propane-1,3-diol), 100 mM NaCl, 0.7 mM Na2SO4 in water at pH 7.6 [31]. After digestion, the suspension was filtered through a nylon membrane (100 μm) and centrifuged at 600× g to remove all cell debris. The supernatant was subjected to consecutive centrifugations at 20,000× g for 30 min and 80,000× g for 1 h, both at 4 °C. The final pellet was homogenized in 200 μL of SCL with 2 mM NaCl2 and stored at 4 °C. Protein determination was assayed with 2 uL of MV suspension and add 798 μL pure water, and 200 μL Bradford [76]. All procedures involved in the euthanasia of animal embryos were approved by the ethics committee of FFCLRP protocol 19.1.842.59.13.

3.3. TNAP Enzymatically Cleaved from MV by PI-PLC

Aliquots of native MV were incubated with phosphatidylcholine phospholipase C (PI-PLC) (0.1 U) in SCL buffer for 3 h under constant rotating agitation at 37 °C [73]. Samples were further centrifuged at 80,000× g for 1 h at 4 °C. The supernatant containing TNAP cleaved from its GPI anchor was reserved. The pellet containing MV with partial cleated TNAP (cMV) was resuspended in the same volume of SCL buffer. Both samples had protein concentration and specific activity quantified by [76] and [78], respectively. The size of cMV was analyzed by DLS.

3.4. Mineralization Assay for MV and for cMV

Mineralization for MV and cMV samples were performed without phosphatidylserine-calcium complex phosphatidylserine calcium complex nucleator (PS-CPLX): MV and cMV were incubated in 96 wells plates with SCL for 120 h at 37 °C in a hydrated incubation chamber. The SCL medium was supplemented with the inhibitors Levamisole 10.0 mM, SBI-425 μM, and ouabain 3.0 mM either alone or in combination as indicated. Samples were pre-incubated for 30 min before 3.41 mM ATP addition. The absorbance at 340 nm was measured within 24-h intervals and normalized by protein concentration generating Abs340nm/mg data. Curves of Abs340nm/mg versus time were plotted, and the sigmoidal tendency of mineral formation was evaluated by the mathematical approach described [88]. Mineralization-related parameters obtained were: the initial mineralization time (ti) is characterized by a rapid increase in U; the final mineralization time (tf) is characterized by a decrease in U; the time in which the maximum rate of mineral formation is reached (tmax rate) corresponds to the maximum of the dU/dt curve; Umax is the maximum of turbidity (Abs340nm/mg); Umax/tmax rate is the potential of mineral propagation (PMP) that is a measure of the tendency to form mineral [88]. Data shown in Table 3.

3.5. Preparation of Na,K-ATPase

Solubilized NKA was obtained from dark red medulla of rabbit kidney as described [62]. Protein concentration was estimated in the presence of 2% SDS (0.2 g/mL) [63], and BSA was used as standard. ATPase activity was assayed discontinuously at 37 °C by quantifying phosphate release using standard conditions: 50.0 mM HEPES buffer, pH 7.5, containing 3.0 mM ATP, 10.0 mM KCl, 5.0 mM MgCl2 and 50.0 mM NaCl in a final volume of 1.0 mL. The reaction was initiated by addition of the enzyme, stopped with 0.5 mL of cold 30% TCA at appropriate time intervals, followed by 4000× g centrifugation immediately prior to phosphate measurement as described [62,63,69]. All procedures involved in the euthanasia of animals were approved by the ethics committee of FFCLRP protocol 17.5.801.59.9.

3.6. Liposomes and Na,K-ATPase Proteoliposomes

DPPC and 1:1 DPPC:DPPE (w/w) multilayer liposomes (1.0 mg/mL) were sonicated during 1 min per milliliter of the sample at 200 W (VibraCell, VC-600 with a microtip) to obtain large unillamelar vesicles (LUVs) [73]. DPPC and 1:1 DPPC:DPPE (w/w) NKA-liposomes were prepared by co-solubilization as previously described [63,66]. Vesicle size was determined by Dynamic light scattering (DLS) at 25 °C using the unimodal distribution. Determination of the protein incorporation in proteoliposomes was performed according to method previously described by Cornelius et al. [61], and BSA was used as standard.

3.7. Mineralization Assay for Na,K-ATPase Proteoliposomes

Mineralization was assayed with a nucleator PS-CPLX. Synthesis of PS-CPLX was performed as previously described [41,82,88,89]. Solubilized NKA, NKA-liposomes (NKA-DPPC and NKA-DPPC:DPPE), and MV were incubated with PS-CPLX in the 96 wells plates for 48 h in synthetic cartilage lymph (SCL) buffer with PS-CPLX. To start the reaction 3.41 mM of ATP was added to the medium. Only initial and end-point were measured to observe the total contribution of NKA. Absorbance was measured at 340 nm and normalized by protein concentration in the assay resulting in Abs340nm·mg−1 information. Abs340nm·mg−1 = [Δ(Abs340nm)/(NKA per well)] was used to compare the ability of mineralization per mg of enzyme used in the assay. Experiments were performed in triplets and are presented in Table 4. Mineralization was also assayed for NKA-liposomes and MV without PS-CPLX in the same conditions described above, but using 3.0 mM ouabain and/or 0.2% C12E8, always with pre-incubated for 30 min before ATP addition.

3.8. Atomic Force Microscopy

MV samples were prepared with SCL buffer. Samples were stabilized by adding 1:10 (v/v) of glutaraldehyde solution (25%), and 5 μL of the sample was placed onto freshly cleaved mica substrate. After stabilization, samples were imaged as described [9]. AFM micrographs were obtained by Shimadzu SPM-9600 Scanning Probe Microscopy (Shimadzu Corporation, Kyoto, Japan) operating in tapping mode (called Phase Contrast Mode by Shimadzu), which is a dynamic-based mode. Scanning was performed in air at 25 °C by using standard 4-sided pyramidal silicon probes with a resonance frequency ranging from 324 to 369 kHz (Nanosensors™, Neuchatel, Switzerland). The scan rate was set at 0.2–0.3 Hz to prevent tip-induced vesicle deformations and/or damages. The values of the spring constants of the cantilevers were approximately 38 ± 8 N/m. Tips radius was 7 nm (guaranteed by the manufacturer >10 nm) and 10–15 μm height. Their resonance frequency values were approximately 336 ± 67 kHz. The vesicle sizes were determined by SPM Offline software from Shimadzu. For each sample, N = 100 vesicles were analyzed.

3.9. Determination of ATP Hydrolysis by Colorimetric Assay and by 31P NMR

ATPase activity was colorimetrically determined discontinuously at 37 °C, and 31P NMR activity was determined by measuring inorganic phosphate released at 25 °C in 50.0 mM HEPES reaction medium, pH 7.5, containing 3.0 mM ATP, 10.0 mM KCl, 5.0 mM MgCl2 and 50.0 mM NaCl and 10% 2H2O for 31P NMR assay [62,63,69]. Evolution 60S UV-Visible Spectrophotometer (Thermo Scientific) spectrophotometer was used for colorimetry. 31P NMR spectra were measured with a 300 ultrashied Bruker spectrometer. One unit of enzyme (1 U) was arbitrarily defined as 1 μmol of phosphate released per minute under standard test conditions (when specified, U/mg was used). Results represent the mean of determinations performed in three independent measurements.

3.10. Determination of p-Nitrophenolphosphate Hydrolysis by Colorimetric Assay

TNAP activity in MVs isolated from chondrocytes, was measured continuously using para-nitrophenyl phosphate (pNPP, ε = 17,600 M−1·cm−1, pH 13.0, 1.0 M) [75]. Aliquots of MVs were mixed with 200 μL of 10 mM pNPP dissolved in 56 mM 2-Amino-2-methyl-1-propanol buffer (Ampol) pH = 10.0 and 2 mM MgCl2 at 37 °C. The absorbance was recorded at 410 nm for 120 s using a 96-well plate reader. Three independent measurements were performed. TNAP activity was calculated as the number of units (U) per mg of total protein (1 U corresponds to 1 μmol of hydrolyzed pNPP per minute). Alternatively, TNAP activity measurement was performed discontinuously. For these experiments, when at optimal pH, 10 mM pNPP dissolved in Ampol 50 mM pH 10.0, 2 mM MgCl2 buffer was used at 37 °C in a final volume of 0.5 mL. The reaction was started by adding an aliquot of the samples and stopped by adding 0.5 mL of 1.0 M NaOH at appropriate time intervals [90,91]. When specified, U/mg was used.

3.11. FTIR Chemical Analysis of Mineral Formed

The chemical composition of the dried precipitates from the mineralization assays was investigated by means of Fourier-transformed infrared spectroscopy using an attenuated total reflectance accessory (ATR-FTIR, germanium crystal with acquisition from 4000–600 cm−1) model IRPrestige-21, Shimadzu Co., Tokyo, Japan.

3.12. Statistical Analysis

Kinetics and mineralization data are reported as the mean ± SD of triplicate measurements of three different proteoliposome preparations, which was considered to be statistically significant at p ≤ 0.05 or p ≤ 0.001, as indicated with “*” in the tables.

4. Conclusions

Extracellular ATP is hydrolyzed mostly by TNAP in MVs, since 5 mM Levamisole inhibited around 69–72% of the total ATP hydrolysis, while around 8–13% was hydrolyzed by other phosphatases, including NTPDase1, NTPDase3, and NPP1 as revealed by the addition of 0.1 mM ARL which can inhibit these phosphatases. TNAP’s main function in MVs is to hydrolyze PPi, which is a strong inhibitor of apatite formation [80,92]. ATP and ADP can also inhibit apatite formation [93]. Therefore, TNAP is an essential enzyme to induce mineralization by depleting PPi, ATP, and ADP [80,92]. To sustain internal apatite formation in the lumen of MVs, Pi has to be either transported into the MVs [26,27,28] and/or produced inside MVs (Figure 1). Several enzymes, including PHOSPHO1 [51,52] and ATPases [53,54], could contribute to the production of internal Pi (Figure 1). Here, NKA accounted for not more than 3–8% of the total ATP hydrolysis, consistent with other reports [53,67]. On the other hand, 164 mM Na+ and 19 mM K+ were found in nascent MVs in contrast to 59 mM Na+ and 101 mM K+ in hypertrophied chondrocytes, from which MVs originate [6]. This suggests that all the ATP was depleted inside MVs and that NKA in MVs could not fulfill its function to maintain ion asymmetry.
The question remains whether NKA could induce apatite formation inside MVs with a sufficient amount of ATP, corresponding to the very early stage of mineralization, before the complete depletion of ATP. Since the ATP hydrolysis was dominated by TNAP and other phosphatases, liposomes were prepared to reconstitute NKA to address the question. Solubilized NKA, as well as NKA-DPPC and NKA-DPPC:DPPE liposomes, were able to propagate minerals in SCL buffer in 48 h in the presence of ATP, Ca2+, and the PS-CPLX nucleator. The NKA-DPPC:DPPE liposome induced greater mineralization, as compared to NKA-DPPC liposomes, because of the better orientation of the ATP active site, which was mostly oriented outside due to DPPE presence [63,66,87]. ATR-FTIR analysis of the mineral formed by NKA proteoliposomes indicated the presence of apatite. The findings support the contention that ATP could be a source of internal Pi, promoting the formation of calcium phosphate complexes, including apatite. It is tempting to suggest that all ATP inside MVs is hydrolyzed by NKA, providing a part of the Pi necessary for apatite formation just after their release from hypertrophied chondrocytes, but MVs do not have the mitochondrial machinery to continuously supply ATP, which may cause the loss of Na+/K+ asymmetry in MVs [6]. Other internal enzymes, such as PHOSPHO1 [51,52], can continue to produce a sufficient amount of luminal Pi to sustain apatite formation. Thus, we propose the putative function of Na,K-ATPase is to transport Na+ out MVs, which could be used by the PiT-1 cotransporter in the initial stages of mineralization (sufficient for nucleation). The Pi produced by Na,K-ATPase from ATP could contribute to the pool of Pi produced by PHOSPHO1.

Author Contributions

Performed the research, H.G.S., L.H.S.A., M.F., C.C., M.A.E.C., D.M. and B.Z.F., conceptualization, M.B. (Massimo Bottini), R.B., S.M., J.L.M. and P.C.; funding acquisition, supervision, and validation, M.B. (Massimo Bottini); J.L.M., A.P.R. and P.C.; investigation, H.G.S., P.C., S.M.; resources, P.C., A.P.R. and M.B. (Massimo Bottini); writing, original draft, H.G.S., A.P.R., R.B., M.B. (Massimo Bottini) and P.C.; writing, review and editing, M.B. (Massimo Bottini), L.H.S.A., H.G.S., B.Z.F., M.F., C.C., A.M., A.P.R., J.L.M., M.B. (Maytê Bolean) and P.C. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP, grants 2017/20846-2; 2019/08568-2; 2019/25054-2; 2021/13140-1 and 2022/05026-7) Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES, Finance Code 001, 1738449, 88887.320304/2019-00 and 88887.368020/2019-00) and Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq, 304021/2017-2 and 305426/2021-4), PVE-Print CAPES-USP 2020-88887.569449/2020-00 and Bando Visiting Professor—Prot. n. 0058868 12/12/2021—Tor Vergata Università degli Studi di Roma. We acknowledge the Brazil-France USP-COFECUB Uc Sv 184/20 grant. Maytê Bolean received a CAPES scholarship. P.C. and A.P.R. are CNPq researchers.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ADPadenosine 5′-Diphosphate
AFMAtomic Force Microcopy
AMPAdenosine 5′-Monophosphate
ATPAdenosine 5′-Triphosphate
ATR-FTIRFourier transformed infrared spectroscopy using an attenuated total reflectance accessory
BSABovine Serum Albumin
C12E8dodecyloctaglycol
CD39nucleoside triphosphate diphosphohydrolase 1
CPPDCalcium pyrophosphate dehydrate
DLSDynamic Light Scattering
DPPC1,2-dipalmitoyl-sn-glycero-3-phosphocholine
DPPE1,2-dipalmitoyl-sn-glycero-3-phosphoethanolamine
GPIGlycosylphosphatidylinositol
HAapatite
HEPESN-(2-hydroxyethyl) piperazine-N′-ethanesulfonic acid
LUVslarge unillamelar vesicles
MVsMatrix vesicles
cMVMatrix vesicles with partial cleaved NAP (Treated with PI-PL-C)
NKANa,K,-ATPase
NPP1Nucleotide Pyrophosphatases/Phosphodiesterases
PCphosphocholine
PC-1glycoprotein-1
PHOSPHO1Orphan phosphatase 1
Piphosphate
PIpolydispersity index
PI-PLCphosphatidylinositol phospholipase C
PiT-1phosphate transporter 1
PMPpotential of mineral propagation
PPiPyrophosphate
PS-CPLXphosphtidylserine complex containing calcium and phosphate
SCLSynthetic Cartilage Lymph
SDSsodium dodecyl sulfate
SMsphingomyelin
SMPD3sphingomyelin phosphodiesterase 3
TCAtrichloroacetic acid
TNAPtissue-nonspecific alkaline phosphatase
TrisTris[hydroxymethyl]aminomethane

References

  1. Schlesinger, P.H.; Braddock, D.T.; Larrouture, Q.C.; Ray, E.C.; Riazanski, V.; Nelson, D.J.; Tourkova, I.L.; Blair, H.C.; Louis, S.; New, Y.; et al. Phylogeny and Chemistry of Biological Mineral Transport. Bone 2020, 141, 115621. [Google Scholar] [CrossRef] [PubMed]
  2. Anderson, H.C. Molecular Biology of Matrix Vesicles. Clin. Orthop. Relat. Res. 1995, 314, 266–280. [Google Scholar] [CrossRef]
  3. Anderson, H.C.; Garimella, R.; Tague, S.E. The Role of Matrix Vesicles in Growth Plate Development and Biomineralization. Front. Biosci. 2005, 10, 822–837. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Golub, E.E. Role of Matrix Vesicles in Biomineralization. Biochim. Biophys. Acta—Gen. Subj. 2009, 1790, 1592–1598. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Golub, E.E. Biomineralization and Matrix Vesicles in Biology and Pathology. Semin. Immunopathol. 2011, 33, 409–417. [Google Scholar] [CrossRef] [Green Version]
  6. Wuthier, R.E.; Lipscomb, G.F. Matrix Vesicles: Structure, Composition, Formation and Function in Calcification. Front. Biosci. 2011, 16, 2812–2902. [Google Scholar] [CrossRef]
  7. Cui, L.; Houston, D.A.; Farquharson, C.; MacRae, V.E. Characterisation of Matrix Vesicles in Skeletal and Soft Tissue Mineralisation. Bone 2016, 87, 147–158. [Google Scholar] [CrossRef] [Green Version]
  8. Ciancaglini, P.; Simão, A.M.S.; Camolezi, F.L.; Millán, J.L.; Pizauro, J.M. Contribution of Matrix Vesicles and Alkaline Phosphatase to Ectopic Bone Formation. Braz. J. Med. Biol. Res. 2006, 39, 603–610. [Google Scholar] [CrossRef] [Green Version]
  9. Bolean, M.; Borin, I.A.; Simão, A.M.S.; Bottini, M.; Bagatolli, L.A.; Hoylaerts, M.F.; Millán, J.L.; Ciancaglini, P. Topographic Analysis by Atomic Force Microscopy of Proteoliposomes Matrix Vesicle Mimetics Harboring TNAP and AnxA5. Biochim. Biophys. Acta—Biomembr. 2017, 1859, 1911–1920. [Google Scholar] [CrossRef]
  10. Bottini, M.; Mebarek, S.; Anderson, K.L.; Strzelecka-Kiliszek, A.; Bozycki, L.; Simão, A.M.S.; Bolean, M.; Ciancaglini, P.; Pikula, J.B.; Pikula, S.; et al. Matrix Vesicles from Chondrocytes and Osteoblasts: Their Biogenesis, Properties, Functions and Biomimetic Models. Biochim. Biophys. Acta—Gen. Subj. 2018, 1862, 532–546. [Google Scholar] [CrossRef]
  11. Cruz, M.A.E.; Ferreira, C.R.; Tovani, C.B.; de Oliveira, F.A.; Bolean, M.; Caseli, L.; Mebarek, S.; Luis Millán, J.; Buchet, R.; Bottini, M.; et al. Phosphatidylserine Controls Calcium Phosphate Nucleation and Growth on Lipid Monolayers: A Physicochemical Understanding of Matrix Vesicle-Driven Biomineralization. J. Struct. Biol. 2020, 212, 107607. [Google Scholar] [CrossRef]
  12. Anderson, H.C. Electron Microscopic Studies of Induced Cartilage Development and Calcification. J. Cell Biol. 1967, 35, 81–101. [Google Scholar] [CrossRef]
  13. Bonucci, E. Fine Structure of Early Cartilage Calcification. J. Ultrasruct. Res. 1967, 20, 33–50. [Google Scholar] [CrossRef]
  14. Anderson, H.C. Vesicles Associated with Calcification in the Matrix of Epiphyseal Cartilage. J. Cell Biol. 1969, 41, 59–72. [Google Scholar] [CrossRef] [Green Version]
  15. Bonucci, E. Fine Structure and Histochemistry of “Calcifying Globules” in Epiphyseal Cartilage. Z. Zellforsch. Mikrosk. Anat. 1970, 103, 192–217. [Google Scholar] [CrossRef]
  16. Rabinovitch, A.L.; Anderson, H.C. Biogenesis of Matrix Vesicles in Cartilage Growth Plates. Fed. Proc. 1976, 35, 112–116. [Google Scholar]
  17. Borg, T.K.; Runyan, R.; Wuthier, R.E. A Freeze-fracture Study of Avian Epiphyseal Cartilage Differentiation. Anat. Rec. 1981, 199, 449–457. [Google Scholar] [CrossRef]
  18. Akisaka, T.; Shigenaga, Y. Ultrastructure of Growing Epiphyseal Cartilage Processed by Rapid Freezing and Freeze-Substitution. Microscopy 1983, 32, 305–320. [Google Scholar] [CrossRef]
  19. Akisaka, T.; Kawaguchi, H.; Subita, G.P.; Shigenaga, Y.; Gay, C.V. Ultrastructure of Matrix Vesicles in Chick Growth Plate as Revealed by Quick Freezing and Freeze Substitution. Calcif. Tissue Int. 1988, 42, 383–393. [Google Scholar] [CrossRef]
  20. Wu, L.N.Y.; Genge, B.R.; Lloyd, G.C.; Wuthier, R.E. Collagen-Binding Proteins in Collagenase-Released Matrix Vesicles from Cartilage. Interaction between Matrix Vesicles Proteins and Different Types of Collagen. J. Biol. Chem. 1991, 266, 1195–1203. [Google Scholar] [CrossRef]
  21. Kirsch, T.; Pfäffle, M. Selective Binding of Anchorin CII (Annexin V) to Type II and X Collagen and to Chondrocalcin (C-Propeptide of Type II Collagen) Implications for Anchoring Function between Matrix Vesicles and Matrix Proteins. FEBS Lett. 1992, 310, 143–147. [Google Scholar] [CrossRef]
  22. Hale, J.E.; Wuthier, R.E. The Mechanism of Matrix Vesicle Formation. Studies on the Composition of Chondrocyte Microvilli and on the Effects of Microfilament-Perturbing Agents on Cellular Vesiculation. J. Biol. Chem. 1987, 262, 1916–1925. [Google Scholar] [CrossRef]
  23. Warner, G.P.; Lee Hubbard, H.; Lloyd, G.C.; Wuthier, R.E. 32Pi- And45Ca-Metabolism by Matrix Vesicle-Enriched Microsomes Prepared from Chicken Epiphyseal Cartilage by Isosmotic Percoll Density-Gradient Fractionation. Calcif. Tissue Int. 1983, 35, 327–338. [Google Scholar] [CrossRef]
  24. Solomon, D.H.; Browning, J.A.; Wilkins, R.J. Inorganic Phosphate Transport in Matrix Vesicles from Bovine Articular Cartilage. Acta Physiol. 2007, 190, 119–125. [Google Scholar] [CrossRef]
  25. Valhmu, W.B.; Wu, L.N.Y.; Wuthier, R.E. Effects of Ca Pi Ratio, Ca2+ × Pi Ion Product, and PH of Incubation Fluid on Accumulation of 45Ca2+ by Matrix Vesicles in Vitro. Bone Miner. 1990, 8, 195–209. [Google Scholar] [CrossRef]
  26. Montessuit, C.; Caverzasio, J.; Bonjour, J.P. Characterization of a P(i) Transport System in Cartilage Matrix Vesicles: Potential Role in the Calcification Process. J. Biol. Chem. 1991, 266, 17791–17797. [Google Scholar] [CrossRef]
  27. Wu, L.N.Y.; Sauer, G.R.; Genge, B.R.; Valhmu, W.B.; Wuthier, R.E. Effects of Analogues of Inorganic Phosphate and Sodium Ion on Mineralization of Matrix Vesicles Isolated from Growth Plate Cartilage of Normal Rapidly Growing Chickens. J. Inorg. Biochem. 2003, 94, 221–235. [Google Scholar] [CrossRef] [PubMed]
  28. Yadav, M.C.; Bottini, M.; Cory, E.; Bhattacharya, K.; Kuss, P.; Narisawa, S.; Sah, R.L.D.; Beck, L.; Fadeel, B.; Farquharson, C.; et al. Skeletal Mineralization Deficits and Impaired Biogenesis and Function of Chondrocyte-Derived Matrix Vesicles in Phospho1−/− and Phospho1/Pit1 Double Knockout Mice HHS Public Access. J Bone Min. Res. 2016, 31, 1275–1286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Plaut, J.S.; Strzelecka-Kiliszek, A.; Bozycki, L.; Pikula, S.; Buchet, R.; Mebarek, S.; Chadli, M.; Bolean, M.; Simao, A.M.S.; Ciancaglini, P.; et al. Quantitative Atomic Force Microscopy Provides New Insight into Matrix Vesicle Mineralization. Arch. Biochem. Biophys. 2019, 667, 14–21. [Google Scholar] [CrossRef] [PubMed]
  30. Wu, L.N.Y.; Genge, B.R.; Sauer, G.R.; Wuthier, R.E. Characterization and Reconstitution of the Nucleational Complex Responsible for Mineral Formation by Growth Plate Cartilage Matrix Vesicles. Connect. Tissue Res. 1996, 35, 309–315. [Google Scholar] [CrossRef] [PubMed]
  31. Wu, L.N.Y.; Genge, B.R.; Dunkelberger, D.G.; Legeros, R.Z.; Concannon, B.; Wuthier, R.E. Physicochemical Characterization of the Nucleational Core of Matrix Vesicles. J. Biol. Chem. 1997, 272, 4404–4411. [Google Scholar] [CrossRef]
  32. Wu, L.N.Y.; Genge, B.R.; Wuthier, R.E. Analysis and Molecular Modeling of the Formation, Structure, and Activity of the Phosphatidylserine-Calcium-Phosphate Complex Associated with Biomineralization. J. Biol. Chem. 2008, 283, 3827–3838. [Google Scholar] [CrossRef] [Green Version]
  33. Wu, L.N.Y.; Yoshimori, T.; Genge, B.R.; Sauer, G.R.; Kirsch, T.; Ishikawa, Y.; Wuthier, R.E. Characterization of the Nucleational Core Complex Responsible for Mineral Induction by Growth Plate Cartilage Matrix Vesicles. J. Biol. Chem. 1993, 268, 25084–25094. [Google Scholar] [CrossRef]
  34. Peress, N.S.; Anderson, H.C.; Sajdera, S.W. The Lipids of Matrix Vesicles from Bovine Fetal Epiphyseal Cartilage. Calcif. Tissue Res. 1974, 14, 275–281. [Google Scholar] [CrossRef]
  35. Wuthier, R.E. Lipids of Matrix Vesicles. Fed. Proc. 1976, 35, 117–121. [Google Scholar]
  36. Wuthier, R.E. Lipid Composition of Isolated Epiphyseal Cartilage Cells, Membranes and Matrix Vesicles. Biochim. Biophys. Acta (BBA)/Lipids Lipid Metab. 1975, 409, 128–143. [Google Scholar] [CrossRef]
  37. Wu, L.N.Y.; Genge, B.R.; Kang, M.W.; Arsenault, A.L.; Wuthier, R.E. Changes in Phospholipid Extractability and Composition Accompany Mineralization of Chicken Growth Plate Cartilage Matrix Vesicles. J. Biol. Chem. 2002, 277, 5126–5133. [Google Scholar] [CrossRef] [Green Version]
  38. Genge, B.R.; Wu, L.N.Y.; Wuthier, R.E. Separation and Quantification of Chicken and Bovine Growth Plate Cartilage Matrix Vesicle Lipids by High-Performance Liquid Chromatography Using Evaporative Light Scattering Detection. Anal. Biochem. 2003, 322, 104–115. [Google Scholar] [CrossRef]
  39. Terkeltaub, R.A. Invited Review Inorganic Pyrophosphate Generation and Disposition in Pathophysiology. Am. J. Physiol. Cell Physiol. 2001, 281, C1–C11. [Google Scholar] [CrossRef] [Green Version]
  40. Huang, R.; Rosenbach, M.; Vaughn, R.; Prowedini, D.; Rebbe, N.; Hickman, S.; Goding, J.; Terkeltaub, R.; Diego, S. Expression of the Murine Plasma Cell Nucleotide Pyrophosphohydrolase PC-1 Is Shared by Human Liver, Bone, and Cartilage Cells. Regulation of PC-1 Expression in Osteosarcoma Cells by Transforming Growth Factor-β. J. Clin. Investig. 1994, 94, 560–567. [Google Scholar] [CrossRef] [Green Version]
  41. Andrilli, L.H.S.; Sebinelli, H.G.; Favarin, B.Z.; Cruz, M.A.E.; Ramos, A.P.; Bolean, M.; Millán, J.L.; Bottini, M.; Ciancaglini, P. NPP1 and TNAP Hydrolyze ATP Synergistically during Biomineralization. Purinergic Signal. 2022. [Google Scholar] [CrossRef] [PubMed]
  42. Hashimoto, S.; Ochs, R.L.; Rosen, F.; Quach, J.; Mccabe, G.; Solan, J.; Seegmiller, J.E.; Terkeltaub, R.; Lotz, M. Chondrocyte-Derived Apoptotic Bodies and Calcification of Articular Cartilage. Proc. Natl. Acad. Sci. USA 1998, 95, 3094–3099. [Google Scholar] [CrossRef] [PubMed]
  43. Johnson, K.; Hashimoto, S.; Lotz, M.; Pritzker, K.; Goding, J.; Terkeltaub, R. Up-regulated Expression of the Phosphodiesterase Nucleotide Pyrophosphatase Family Member PC-1 Is a Marker and Pathogenic Factor for Knee Meniscal Cartilage Matrix Calcification. Arthritis Rheum. 2001, 44, 1071–1081. [Google Scholar] [CrossRef] [PubMed]
  44. Terkeltaub, R.; Rosenbach, M.; Fong, F.; Goding, J. Causal Link between Nucleotide Pyrophosphohydrolase Overactivity and Increased Intracellular Inorganic Pyrophosphate Generation Demonstrated by Transfection of Cultured Fibroblasts and Osteoblasts with Plasma Cell Membrane Glycoprotein–1. Arthritis Rheum. 1994, 37, 934–941. [Google Scholar] [CrossRef] [PubMed]
  45. Jones, A.C.; Chuck, A.J.; Arie, E.A.; Green, D.J.; Doherty, M. Diseases Associated with Calcium Pyrophosphate Deposition Disease. Semin. Arthritis Rheum. 1992, 22, 188–202. [Google Scholar] [CrossRef]
  46. Ciancaglini, P.; Yadav, M.C.; Simão, A.M.S.; Narisawa, S.; Pizauro, J.M.; Farquharson, C.; Hoylaerts, M.F.; Millán, J.L. Kinetic Analysis of Substrate Utilization by Native and TNAP-, NPP1-, or PHOSPHO1-Deficient Matrix Vesicles. J. Bone Miner. Res. 2010, 25, 716–723. [Google Scholar] [CrossRef]
  47. Simão, A.M.S.; Yadav, M.C.; Ciancaglini, P.; Millán, J.L. Proteoliposomes as Matrix Vesicles’ Biomimetics to Study the Initiation of Skeletal Mineralization. Braz. J. Med. Biol. Res. 2010, 43, 234–241. [Google Scholar] [CrossRef] [Green Version]
  48. Simão, A.M.S.; Bolean, M.; Hoylaerts, M.F.; Millán, J.L.; Ciancaglini, P. Effects of PH on the Production of Phosphate and Pyrophosphate by Matrix Vesicles’ Biomimetics. Calcif. Tissue Int. 2013, 93, 222–232. [Google Scholar] [CrossRef] [Green Version]
  49. Bolean, M.; Simão, A.M.S.; Barioni, M.B.; Favarin, B.Z.; Sebinelli, H.G.; Veschi, E.A.; Janku, T.A.B.; Bottini, M.; Hoylaerts, M.F.; Itri, R.; et al. Biophysical Aspects of Biomineralization. Biophys. Rev. 2017, 9, 747–760. [Google Scholar] [CrossRef] [Green Version]
  50. Latini, S.; Pedata, F. Adenosine in the Central Nervous System: Release Mechanisms and Extracellular Concentrations. J. Neurochem. 2001, 79, 463–484. [Google Scholar] [CrossRef] [Green Version]
  51. Stewart, A.J.; Roberts, S.J.; Seawright, E.; Davey, M.G.; Fleming, R.H.; Farquharson, C. The Presence of PHOSPHO1 in Matrix Vesicles and Its Developmental Expression Prior to Skeletal Mineralization. Bone 2006, 39, 1000–1007. [Google Scholar] [CrossRef] [Green Version]
  52. Yadav, M.C.; Simão, A.M.S.; Narisawa, S.; Huesa, C.; McKee, M.D.; Farquharson, C.; Millán, J.L. Loss of Skeletal Mineralization by the Simultaneous Ablation of PHOSPHO1 and Alkaline Phosphatase Function: A Unified Model of the Mechanisms of Initiation of Skeletal Calcification. J. Bone Miner. Res. 2011, 26, 286–297. [Google Scholar] [CrossRef]
  53. Hsu, H.H.T.; Clarke Anderson, H. A Role for ATPase in the Mechanisms of ATP-Dependent Ca and Phosphate Deposition by Isolated Rachitic Matrix Vesicles. Int. J. Biochem. Cell Biol. 1995, 27, 1349–1356. [Google Scholar] [CrossRef]
  54. Anderson, H.C.; Sipe, J.B.; Hessle, L.; Dhamyamraju, R.; Atti, E.; Camacho, N.P.; Millán, J.L. Impaired Calcification Around Matrix Vesicles of Growth Plate and Bone in Alkaline Phosphatase-Deficient Mice. Am. J. Pathol. 2004, 164, 841–847. [Google Scholar] [CrossRef] [Green Version]
  55. Morth, J.P.; Pedersen, B.P.; Buch-Pedersen, M.J.; Andersen, J.P.; Vilsen, B.; Palmgren, M.G.; Nissen, P. A Structural Overview of the Plasma Membrane Na+,K +-ATPase and H+-ATPase Ion Pumps. Nat. Rev. Mol. Cell Biol. 2011, 12, 60–70. [Google Scholar] [CrossRef]
  56. Lingrel, J. The Physiological Significance of the Cardiotonic Steroid/Ouabain-Binding Site of the Na,K-ATPase. Annu. Rev. Physiol. 2010, 17, 395–4112. [Google Scholar] [CrossRef] [Green Version]
  57. Kaplan, J.H. Biochemistry of Na,K-ATPase. Annu. Rev. Biochem. 2002, 71, 511–535. [Google Scholar] [CrossRef]
  58. Sheen, C.R.; Kuss, P.; Narisawa, S.; Yadav, M.C.; Nigro, J.; Wang, W.; Chhea, T.N.; Sergienko, E.A.; Kapoor, K.; Jackson, M.R.; et al. Pathophysiological Role of Vascular Smooth Muscle Alkaline Phosphatase in Medial Artery Calcification. J. Bone Miner. Res. 2015, 30, 824–836. [Google Scholar] [CrossRef] [Green Version]
  59. Pinkerton, A.B.; Sergienko, E.A.; Bravo, Y.; Dahl, R.; Ma, C.-T.; Sun, Q.; Jackson, M.R.; Cosford, N.D.P.; Millán, J.L. Discovery of 5-((5-Chloro-2- Methoxyphenyl)Sulfonamido)Nicotinamide (SBI-425), a Potent and Orally Bioavailable Tissue-Nonspecific Alkaline Phosphatase (TNAP) Inhibitor. Bioorg. Med. Chem. Lett. 2018, 28, 31–34. [Google Scholar] [CrossRef]
  60. Ciancaglini, P.; Simão, A.M.S.; Bolean, M.; Millán, J.L.; Rigos, C.F.; Yoneda, J.S.; Colhone, M.C.; Stabeli, R.G. Proteoliposomes in Nanobiotechnology. Biophys. Rev. 2012, 4, 67–81. [Google Scholar] [CrossRef] [Green Version]
  61. Cornelius, F.; Habeck, M.; Kanai, R.; Toyoshima, C.; Karlish, S.J.D. General and Specific Lipid-Protein Interactions in Na,K-ATPase. Biochim. Biophys. Acta—Biomembr. 2015, 1848, 1729–1743. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Santos, H.L.; Lamas, R.P.; Ciancaglini, P. Solubilization of Na,K-ATPase from Rabbit Kidney Outer Medulla Using Only C12E8. Braz. J. Med. Biol. Res. 2002, 35, 277–288. [Google Scholar] [CrossRef] [PubMed]
  63. De Lima Santos, H.; Lopes, M.L.; Maggio, B.; Ciancaglini, P. Na,K-ATPase Reconstituted in Liposomes: Effects of Lipid Composition on Hydrolytic Activity and Enzyme Orientation. Colloids Surf. B Biointerfaces 2005, 41, 239–248. [Google Scholar] [CrossRef] [PubMed]
  64. Yoneda, J.S.; Scanavachi, G.; Sebinelli, H.G.; Borges, J.C.; Barbosa, L.R.S.; Ciancaglini, P.; Itri, R. Multimeric Species in Equilibrium in Detergent-Solubilized Na,K-ATPase. Int. J. Biol. Macromol. 2016, 89, 238–245. [Google Scholar] [CrossRef] [PubMed]
  65. Yoneda, J.S.; Sebinelli, H.G.; Itri, R.; Ciancaglini, P. Overview on Solubilization and Lipid Reconstitution of Na,K-ATPase: Enzyme Kinetic and Biophysical Characterization. Biophys. Rev. 2020, 12, 49–64. [Google Scholar] [CrossRef] [PubMed]
  66. Sebinelli, H.G.; Borin, I.A.; Ciancaglini, P.; Bolean, M. Topographical and Mechanical Properties of Liposome Surfaces Harboring Na,K-ATPase by Means of Atomic Force Microscopy. Soft Matter 2019, 15, 2737–2745. [Google Scholar] [CrossRef]
  67. Majeska, R.J.; Wuthier, R.E. Studies on Matrix Vesicles Isolated from Chick Epiphyseal Cartilage Association of Pyrophosphatase and ATPase Activities with Alkaline Phosphatase. BBA—Enzymol. 1975, 391, 51–60. [Google Scholar] [CrossRef]
  68. Van Belle, H. Alkaline Phosphatase. I. Kinetics and Inhibition by Levamisole of Purified Isoenzymes from Humans. Clin. Chem. 1976, 22, 972–976. [Google Scholar] [CrossRef]
  69. Heinonen, J.K.; Lahti, R.J. A New and Convenient Colorimetric Determination of Inorganic Orthophosphate and Its Application to the Assay of Inorganic Pyrophosphatase. Anal. Biochem. 1981, 113, 313–317. [Google Scholar] [CrossRef]
  70. Lévesque, S.A.; Lavoie, É.G.; Lecka, J.; Bigonnesse, F.; Sévigny, J. Specificity of the Ecto-ATPase Inhibitor ARL 67156 on Human and Mouse Ectonucleotidases. Br. J. Pharmacol. 2007, 152, 141–150. [Google Scholar] [CrossRef]
  71. Luo, Y.; Wu, W.; Gu, J.; Zhang, X.; Dang, J.; Wang, J.; Zheng, Y.; Huang, F.; Yuan, J.; Xue, Y.; et al. Human Gingival Tissue-Derived MSC Suppress Osteoclastogenesis and Bone Erosion via CD39-Adenosine Signal Pathway in Autoimmune Arthritis. EBioMedicine 2019, 43, 620–631. [Google Scholar] [CrossRef] [Green Version]
  72. Wu, W.; Xiao, Z.; Chen, Y.; Deng, Y.; Zeng, D.; Liu, Y.; Huang, F.; Wang, J.; Liu, Y.; Bellanti, J.A.; et al. CD39 Produced from Human GMSCs Regulates the Balance of Osteoclasts and Osteoblasts through the Wnt/β-Catenin Pathway in Osteoporosis. Mol. Ther. 2020, 28, 1518–1532. [Google Scholar] [CrossRef]
  73. Camolezi, F.L.; Daghastanli, K.R.P.; Magalhães, P.P.; Pizauro, J.M.; Ciancaglini, P. Construction of an Alkaline Phosphatase-Liposome System: A Tool for Biomineralization Study. Int. J. Biochem. Cell Biol. 2002, 34, 1091–1101. [Google Scholar] [CrossRef]
  74. Zhang, L.; Balcerzak, M.; Radisson, J.; Thouverey, C.; Pikula, S.; Rard Azzar, G.; Buchet, R. Phosphodiesterase Activity of Alkaline Phosphatase in ATP-Initiated Ca(2+) and Phosphate Deposition in Isolated Chicken Matrix Vesicles. J. Biol. Chem. 2005, 280, 37289–37296. [Google Scholar] [CrossRef] [Green Version]
  75. Buchet, R.; Pikula, S.; Magne, D.; Mebarek, S. Isolation and Characteristics of Matrix Vesicles. In Phosphatase Modulators; Millán, J.L., Ed.; Methods in Molecular Biology; Humana Press: Totowa, NJ, USA, 2013; Volume 1053, pp. 115–124. ISBN 978-1-62703-561-3. [Google Scholar]
  76. Bradford, M.M. A Rapid and Sensitive Method for the Quantitation Microgram Quantities of Protein Utilizing the Principle of Protein-Dye Binding. Anal. Biochem. 1976, 72, 248–254. [Google Scholar] [CrossRef]
  77. Kozlenkov, A.; Le Du, M.H.; Cuniasse, P.; Ny, T.; Hoylaerts, M.F.; Millán, J.L. Residues Determining the Binding Specificity of Uncompetitive Inhibitors to Tissue-Nonspecific Alkaline Phosphatase. J. Bone Miner. Res. 2004, 19, 1862–1872. [Google Scholar] [CrossRef]
  78. Meyer, J.L. Can Biological Calcification Occur in the Presence of Pyrophosphate? Arch. Biochem. Biophys. 1984, 231, 1–8. [Google Scholar] [CrossRef]
  79. Johnson, K.; Moffa, A.; Chen, Y.; Pritzker, K.; Goding, J.; Terkeltaub, R. Matrix Vesicle Plasma Cell Membrane Glycoprotein-1 Regulates Mineralization by Murine Osteoblastic MC3T3 Cells. J. Bone Miner. Res. 1999, 14, 883–892. [Google Scholar] [CrossRef]
  80. Hessle, L.; Johnson, K.A.; Anderson, H.C.; Narisawa, S.; Sali, A.; Goding, J.W.; Terkeltaub, R.; Millán, J.L. Tissue-Nonspecific Alkaline Phosphatase and Plasma Cell Membrane Glycoprotein-1 Are Central Antagonistic Regulators of Bone Mineralization. Proc. Natl. Acad. Sci. USA 2002, 99, 9445–9449. [Google Scholar] [CrossRef] [Green Version]
  81. Shinoda, T.; Ogawa, H.; Cornelius, F.; Toyoshima, C. Crystal Structure of the Sodium-Potassium Pump at 2.4 Resolution. Nature 2009, 459, 446–450. [Google Scholar] [CrossRef]
  82. Simão, A.M.S.; Bolean, M.; Favarin, B.Z.; Veschi, E.A.; Tovani, C.B.; Ramos, A.P.; Bottini, M.; Buchet, R.; Millán, J.L.; Ciancaglini, P. Lipid Microenvironment Affects the Ability of Proteoliposomes Harboring TNAP to Induce Mineralization without Nucleators. J. Bone Miner. Metab. 2019, 37, 607–613. [Google Scholar] [CrossRef] [PubMed]
  83. Tejera-Garcia, R.; Ranjan, S.; Zamotin, V.; Sood, R.; Kinnunen, P.K.J. Making Unilamellar Liposomes Using Focused Ultrasound. Langmuir 2011, 27, 10088–10097. [Google Scholar] [CrossRef] [PubMed]
  84. De Lima Santos, H.; Ciancaglini, P. Kinetic Characterization of Na,K-ATPase from Rabbit Outer Renal Medulla: Properties of the (Aβ)2 Dimer. Comp. Biochem. Physiol.—B Biochem. Mol. Biol. 2003, 135, 539–549. [Google Scholar] [CrossRef] [PubMed]
  85. Rigos, C.F.; Santos, H.D.L.; Thedei, G.; Ward, R.J.; Ciancaglini, P. Influence of Enzyme Conformational Changes on Catalytic Activity Investigated by Circular Dichroism Spectroscopy. Biochem. Mol. Biol. Educ. 2003, 31, 329–332. [Google Scholar] [CrossRef]
  86. Santos, H.D.L.; Rigos, C.F.; Tedesco, A.C.; Ciancaglini, P. Rose Bengal Located within Liposome Do Not Affect the Activity of Inside-out Oriented Na,K-ATPase. Biochim. Biophys. Acta—Biomembr. 2005, 1715, 96–103. [Google Scholar] [CrossRef] [Green Version]
  87. Rigos, C.F.; Santos, H.D.L.; Ward, R.J.; Ciancaglini, P. Lipid Bilayer Stabilization of the Na,K-ATPase Reconstituted in DPPC/DPPE Liposomes. Cell Biochem. Biophys. 2006, 44, 438–445. [Google Scholar] [CrossRef]
  88. Harmey, D.; Hessle, L.; Narisawa, S.; Johnson, K.A.; Terkeltaub, R.; Millán, J.L. Concerted Regulation of Inorganic Pyrophosphate and Osteopontin by Akp2, Enpp1, and Ank: An Integrated Model of the Pathogenesis of Mineralization Disorders. Am. J. Pathol. 2004, 164, 1199–1209. [Google Scholar] [CrossRef]
  89. Li, L.; Buchet, R.; Wu, Y. Dimethyl Sulfoxide-Induced Hydroxyapatite Formation: A Biological Model of Matrix Vesicle Nucleation to Screen Inhibitors of Mineralization. Anal. Biochem. 2008, 381, 123–128. [Google Scholar] [CrossRef]
  90. Genge, B.R.; Wu, L.N.Y.; Wuthier, R.E. In Vitro Modeling of Matrix Vesicle Nucleation: Synergistic Stimulation of Mineral Formation by Annexin A5 and Phosphatidylserine. J. Biol. Chem. 2007, 282, 26035–26045. [Google Scholar] [CrossRef] [Green Version]
  91. Veschi, E.A.; Bolean, M.; da Silva Andrilli, L.H.; Sebinelli, H.G.; Strzelecka-Kiliszek, A.; Bandorowicz-Pikula, J.; Pikula, S.; Granjon, T.; Mebarek, S.; Magne, D.; et al. Mineralization Profile of Annexin A6-Harbouring Proteoliposomes: Shedding Light on the Role of Annexin A6 on Matrix Vesicle-Mediated Mineralization. Int. J. Mol. Sci. 2022, 23, 8945. [Google Scholar] [CrossRef]
  92. Favarin, B.F.; Andrade, M.A.R.; Bolean, M.; Simão, A.M.S.; Ramos, A.P.; Hoylaerts, M.F.; Millán, J.L.; Ciancaglini, P. Effect of the Presence of Cholesterol in the Interfacial Microenvironment on the Modulation of the Alkaline Phosphatase Activity during in Vitro Mineralization. Colloids Surf. B Biointerfaces 2017, 155, 466–476. [Google Scholar] [CrossRef]
  93. Bolean, M.; Simão, M.S.; Kiffer-Moreira, T.; Hoylaerts, M.F.; Luis Millán, J.; Itri, R.; Ciancaglini, P. Proteoliposomes with the Ability to Transport Ca2+ into the Vesicles and Hydrolyze Phosphosubstrates on Their Surface Graphical Abstract HHS Public Access. Arch. Biochem. Biophys. 2015, 584, 79–89. [Google Scholar] [CrossRef]
Figure 1. Pi homeostasis during MV-mediated mineralization, adapted from Reference [10]. MVs provide an environment that allows the initial nucleation of apatitic mineral inside the vesicles as well as to provide nucleation sites that can help propagate the apatitic crystals onto the extracellular matrix through interaction with the collagen fibers. Orphan phosphatase 1 (PHOSPHO1) produces Pi from the hydrolysis of phosphocholine (PC), which itself is derived from sphingomyelin (SM) by the action of sphingomyelin phosphodiesterase 3 (SMPD3) located in the inner surface of the MV membrane. In addition, phosphate transporter 1 (PiT-1, in blue box) and possibly other unidentified transporter(s) help supply Pi ions generated perivesicularly to help nucleation of apatite (HA) inside the MVs. TNAP and NPP1 both control the extracellular PPi/Pi ratio. An Na+ gradient could be generated by NKA (red circle) after ATP hydrolysis inside MVs, helping increase the Pi pool generated by PHOSPHO1.
Figure 1. Pi homeostasis during MV-mediated mineralization, adapted from Reference [10]. MVs provide an environment that allows the initial nucleation of apatitic mineral inside the vesicles as well as to provide nucleation sites that can help propagate the apatitic crystals onto the extracellular matrix through interaction with the collagen fibers. Orphan phosphatase 1 (PHOSPHO1) produces Pi from the hydrolysis of phosphocholine (PC), which itself is derived from sphingomyelin (SM) by the action of sphingomyelin phosphodiesterase 3 (SMPD3) located in the inner surface of the MV membrane. In addition, phosphate transporter 1 (PiT-1, in blue box) and possibly other unidentified transporter(s) help supply Pi ions generated perivesicularly to help nucleation of apatite (HA) inside the MVs. TNAP and NPP1 both control the extracellular PPi/Pi ratio. An Na+ gradient could be generated by NKA (red circle) after ATP hydrolysis inside MVs, helping increase the Pi pool generated by PHOSPHO1.
Ijms 23 15072 g001
Figure 2. Tapping mode AFM images of isolated MV prepared in SCL buffer without phosphate and placed in fresh cleaved mica. (A) 3D topographic image showing the vesicle’s topology. (B) Phase image showing short range of phase shift in the vesicle’s topology, revealing a high protein complexity. Scanning was performed in air at 25 °C by using standard 4-sided pyramidal silicon probes with a resonance frequency ranging from 324 to 369 kHz (Nanosensors™, Switzerland). The scan rate was set at 0.2–0.3 Hz to prevent tip-induced vesicle deformations and/or damages. The vesicles sizes were determined by SPM Offline software, from Shimadzu. For each sample, N = 100 vesicles were analyzed.
Figure 2. Tapping mode AFM images of isolated MV prepared in SCL buffer without phosphate and placed in fresh cleaved mica. (A) 3D topographic image showing the vesicle’s topology. (B) Phase image showing short range of phase shift in the vesicle’s topology, revealing a high protein complexity. Scanning was performed in air at 25 °C by using standard 4-sided pyramidal silicon probes with a resonance frequency ranging from 324 to 369 kHz (Nanosensors™, Switzerland). The scan rate was set at 0.2–0.3 Hz to prevent tip-induced vesicle deformations and/or damages. The vesicles sizes were determined by SPM Offline software, from Shimadzu. For each sample, N = 100 vesicles were analyzed.
Ijms 23 15072 g002
Figure 3. MV (A) and cMV (B) were incubated in 96 wells plates with SCL for 120 h at 37 °C in an incubation hydrated chamber. The SCL medium with 2.0 mM of CaCl2 was supplemented with the inhibitors Levamisole 10.0 mM, SBI-425 10 μM and Ouabain 3.0 mM either alone or in combination as indicated. Samples were pre-incubated for 30 min before the addition of 3.41 mM ATP. The absorbance at 340 nm was measured within 24 h intervals and normalized by protein concentration generating Abs340nm/mg data and resulting in the kinetic parameters presented on Table 3, as described in Materials and Methods. Control (absence of inhibitor) (■ black); Levamisole, phosphatase inhibitor (red ); SBI-425, TNAP specific inhibitor (cyan ); Ouabain, NKA specific inhibitor (green ); Levamisole + Ouabain ( blue); SBI-425 + Ouabain ( pink).
Figure 3. MV (A) and cMV (B) were incubated in 96 wells plates with SCL for 120 h at 37 °C in an incubation hydrated chamber. The SCL medium with 2.0 mM of CaCl2 was supplemented with the inhibitors Levamisole 10.0 mM, SBI-425 10 μM and Ouabain 3.0 mM either alone or in combination as indicated. Samples were pre-incubated for 30 min before the addition of 3.41 mM ATP. The absorbance at 340 nm was measured within 24 h intervals and normalized by protein concentration generating Abs340nm/mg data and resulting in the kinetic parameters presented on Table 3, as described in Materials and Methods. Control (absence of inhibitor) (■ black); Levamisole, phosphatase inhibitor (red ); SBI-425, TNAP specific inhibitor (cyan ); Ouabain, NKA specific inhibitor (green ); Levamisole + Ouabain ( blue); SBI-425 + Ouabain ( pink).
Ijms 23 15072 g003
Figure 4. ATR-FTIR absorbance spectra of dehydrated minerals obtained from DPPC-NKA and DPPC:DPPE-NKA liposomes samples after 48 h’ incubation with ATP, 0.2% C12E8 detergent and 3.0 mM Ouabain (NKA inhibitor) when described. Bands of HPO32− (987 cm−1) and PO43− (1040 cm−1) are indicated in blue and orange dash lines, respectively. (A1A3) DPPC-NKA liposomes. (A1) Effect of C12E8. (A2) Effect of Ouabain. (A3) Effect of combination of 0.2% C12E8 and 3 mM Ouabain. (B1B3) DPPC-DPPE-NKA liposomes. (B1) Effect of 0.2% C12E8. (B2) Effect of Ouabain. (B3) Effect of combination of 0.2% C12E8 and 3 mM Ouabain. In the figures, Black line was used for the control spectra with only NKA-liposome, A1 to A3 DPPC-NKA and B1 to B3 DPPC:DPPE-NKA (▬) and red line represent the spectra with 3 mM Ouabain or 0.2% C12E8 ().
Figure 4. ATR-FTIR absorbance spectra of dehydrated minerals obtained from DPPC-NKA and DPPC:DPPE-NKA liposomes samples after 48 h’ incubation with ATP, 0.2% C12E8 detergent and 3.0 mM Ouabain (NKA inhibitor) when described. Bands of HPO32− (987 cm−1) and PO43− (1040 cm−1) are indicated in blue and orange dash lines, respectively. (A1A3) DPPC-NKA liposomes. (A1) Effect of C12E8. (A2) Effect of Ouabain. (A3) Effect of combination of 0.2% C12E8 and 3 mM Ouabain. (B1B3) DPPC-DPPE-NKA liposomes. (B1) Effect of 0.2% C12E8. (B2) Effect of Ouabain. (B3) Effect of combination of 0.2% C12E8 and 3 mM Ouabain. In the figures, Black line was used for the control spectra with only NKA-liposome, A1 to A3 DPPC-NKA and B1 to B3 DPPC:DPPE-NKA (▬) and red line represent the spectra with 3 mM Ouabain or 0.2% C12E8 ().
Ijms 23 15072 g004
Table 1. Comparison of ATP hydrolysis activity determined by colorimetric and 31P NMR method. The activities were determined in % considering 100% activity in the absence of inhibitor under optimal conditions for the determination of NKA activity, as described in materials and Methods. Ouabain (NKA specific inhibitor); Levamisole and SBI-425 (TNAP inhibitors); ARL-67156 (CD39 specific inhibitor). ATPase activity was colorimetrically determined discontinuously, at 37 °C and 31P NMR activity was determined by measuring inorganic phosphate released, in 50.0 mM HEPES reaction medium, at 25 °C using medium with pH 7.5, containing 3.0 mM ATP, 10.0 mM KCl, 5.0 mM MgCl2 and 50.0 mM NaCl and 10% 2H2O for 31P NMR assay [62,63,69].
Table 1. Comparison of ATP hydrolysis activity determined by colorimetric and 31P NMR method. The activities were determined in % considering 100% activity in the absence of inhibitor under optimal conditions for the determination of NKA activity, as described in materials and Methods. Ouabain (NKA specific inhibitor); Levamisole and SBI-425 (TNAP inhibitors); ARL-67156 (CD39 specific inhibitor). ATPase activity was colorimetrically determined discontinuously, at 37 °C and 31P NMR activity was determined by measuring inorganic phosphate released, in 50.0 mM HEPES reaction medium, at 25 °C using medium with pH 7.5, containing 3.0 mM ATP, 10.0 mM KCl, 5.0 mM MgCl2 and 50.0 mM NaCl and 10% 2H2O for 31P NMR assay [62,63,69].
ATPase Activity MethodOuabain
(3 mM)
Levamisole
(5 mM)
ARL
(0.1 mM)
SBI-425
(5 μM)
SBI-425
(10 μM)
Colorimetric93.2 ± 3.627.892.0 ± 5.678.564.6
31P NMR97.6 ± 2.931.4 ± 3.986.3 ± 4.2 NDND
ND (Not determined).
Table 3. Kinetic parameters of nucleation process obtained from the mineralization curves for MV and cMV, after 120 h assay in SCL buffer supplemented with 2.0 mM of CaCl2, and inhibitors. Data are reported as the mean ± SD of triplicate measurements of three different MV preparations.
Table 3. Kinetic parameters of nucleation process obtained from the mineralization curves for MV and cMV, after 120 h assay in SCL buffer supplemented with 2.0 mM of CaCl2, and inhibitors. Data are reported as the mean ± SD of triplicate measurements of three different MV preparations.
Inhibitor (Concentration)ti (h)tf (h)tmax rate (h)Umax (Abs340nm·mg−1) PMP (h−1)
MVWithout20.0 ± 0.4120.0 ± 0.657.9 ± 0.850.1 ± 0.80.86 ± 0.04
Levamisole (5 mM)NDNDNDNDND
Ouabain (3 mM)24.3 ± 0.7160.5 ± 7.674.4 ± 0.754.6 ± 3.30.73 ± 0.03
Levamisole (5mM) + Ouabain (3 mM)60.0 ± 0.3119.0 ± 1.691.7 ± 0.521.4 ± 0.80.23 ± 0.02
SBI-425 (10 μM)15.0 ± 0.4120.0 ± 2.458.9 ± 0.348.8 ± 1.10.83 ± 0.05
SBI-425 (10 μM) + Ouabain (3 mM)14.6 ± 0.2140.7 ± 1.562.3 ± 5.671.6 ± 2.31.15 ± 0.07
cMVWithout34.1 ± 1.3140.9 ± 5.279.7 ± 3.458.6 ± 3.70.74 ± 0.04
Levamisole (5 mM)NDNDNDNDND
Ouabain (3 mM)35.9 ± 1.6125.6 ± 4.135.2 ± 1.069.8 ± 1.21.98 ± 0.09
Levamisole (5 mM) + Ouabain (3 mM)NDNDNDNDND
SBI-425 (10 μM)48.9 ± 1.8128.2 ± 3.937.4 ± 2.383.3 ± 2.12.22 ± 0.11
SBI-425 (10 μM) + Ouabain (3 mM)37.6 ± 3.894.4 ± 2.415.6 ± 2.255.9 ± 4.83.58 ± 0.14
Mineralization-related parameters obtained were: the initial mineralization time (ti) is characterized by a rapid increase in U; the final mineralization time (tf) is characterized by a decrease in U; the time in which the maximum rate of mineral formation is reached (tmax rate) corresponds to the maximum of the dU/dt curve; Umax is the maximum of turbidity (Abs340nm/mg); Umax/tmax rate is the potential of mineral propagation (PMP) that is a measure of the tendency to form mineral [32]. ND, not determined.
Table 4. Biochemical and biophysical characteristics of NKA solubilized and NKA-liposomes prepared in DPPC and DPPC:DPPE. Protein quantification was assayed by Cornelius et al. [65]. ATPase activity was assayed discontinuously at 37 °C by quantifying phosphate release using standard conditions: 50.0 mM HEPES buffer, pH 7.5, containing 3.0 mM ATP, 10.0 mM KCl, 5.0 mM MgCl2 and 50.0 mM NaCl as described in [62,63,69]. The ability of mineralization was tested in SCL buffer with 2.0 mM of CaCl2, in the presence of PS-CPLX nucleator and 48 h of incubation at 37 °C with 3.41 mM ATP as Pi source. Absorbance was measured at 340 nm and normalized by protein concentration in the assay resulting in Abs340nm·mg−1.
Table 4. Biochemical and biophysical characteristics of NKA solubilized and NKA-liposomes prepared in DPPC and DPPC:DPPE. Protein quantification was assayed by Cornelius et al. [65]. ATPase activity was assayed discontinuously at 37 °C by quantifying phosphate release using standard conditions: 50.0 mM HEPES buffer, pH 7.5, containing 3.0 mM ATP, 10.0 mM KCl, 5.0 mM MgCl2 and 50.0 mM NaCl as described in [62,63,69]. The ability of mineralization was tested in SCL buffer with 2.0 mM of CaCl2, in the presence of PS-CPLX nucleator and 48 h of incubation at 37 °C with 3.41 mM ATP as Pi source. Absorbance was measured at 340 nm and normalized by protein concentration in the assay resulting in Abs340nm·mg−1.
SampleNKA (mg·mL−1)ATPase
(U·mg−1)
Diameter (nm)IPAbs340nm·mg−1
NKA solubilized0.190.33619.1 ± 1 *--51.9 ± 5
DPPC------365 ± 350.7---
DPPC:DPPE------504 ± 501.3---
NKA-DPPC0.61 0.141634 ± 601.438.4 ± 4
NKA-DPPC:DPPE0.530.290899 ± 901.451.4 ± 5
* Values from reference [65].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sebinelli, H.G.; Andrilli, L.H.S.; Favarin, B.Z.; Cruz, M.A.E.; Bolean, M.; Fiore, M.; Chieffo, C.; Magne, D.; Magrini, A.; Ramos, A.P.; et al. Shedding Light on the Role of Na,K-ATPase as a Phosphatase during Matrix-Vesicle-Mediated Mineralization. Int. J. Mol. Sci. 2022, 23, 15072. https://doi.org/10.3390/ijms232315072

AMA Style

Sebinelli HG, Andrilli LHS, Favarin BZ, Cruz MAE, Bolean M, Fiore M, Chieffo C, Magne D, Magrini A, Ramos AP, et al. Shedding Light on the Role of Na,K-ATPase as a Phosphatase during Matrix-Vesicle-Mediated Mineralization. International Journal of Molecular Sciences. 2022; 23(23):15072. https://doi.org/10.3390/ijms232315072

Chicago/Turabian Style

Sebinelli, Heitor Gobbi, Luiz Henrique Silva Andrilli, Bruno Zoccaratto Favarin, Marcos Aantonio Eufrasio Cruz, Maytê Bolean, Michele Fiore, Carolina Chieffo, David Magne, Andrea Magrini, Ana Paula Ramos, and et al. 2022. "Shedding Light on the Role of Na,K-ATPase as a Phosphatase during Matrix-Vesicle-Mediated Mineralization" International Journal of Molecular Sciences 23, no. 23: 15072. https://doi.org/10.3390/ijms232315072

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop