Next Article in Journal
Flame-Synthesis of Carbon Nanotube Forests on Metal Mesh Structure: Dependence, Morphology, and Application
Next Article in Special Issue
Novel Nanocomposites Based on Functionalized Magnetic Nanoparticles and Polyacrylamide: Preparation and Complex Characterization
Previous Article in Journal
Robust Polymer Nanocomposite Membranes Incorporating Discrete TiO2 Nanotubes for Water Treatment
Previous Article in Special Issue
Stretchable and Low-Haze Ag-Nanowire-Network 2-D Films Embedded into a Cross-linked Polydimethylsiloxane Elastomer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photoluminescent Hydroxylapatite: Eu3+ Doping Effect on Biological Behaviour

by
Ecaterina Andronescu
1,2,3,
Daniela Predoi
4,
Ionela Andreea Neacsu
1,2,
Andrei Viorel Paduraru
1,
Adina Magdalena Musuc
1,5,
Roxana Trusca
2,3,
Ovidiu Oprea
1,2,
Eugenia Tanasa
2,3,
Otilia Ruxandra Vasile
2,3,
Adrian Ionut Nicoara
1,2,
Adrian Vasile Surdu
1,2,
Florin Iordache
6,
Alexandra Catalina Birca
1,2,
Simona Liliana Iconaru
4 and
Bogdan Stefan Vasile
1,2,3,*
1
Faculty of Applied Chemistry and Materials Science, Department of Science and Engineering of Oxide Materials and Nanomaterials, University Politehnica of Bucharest, 060042 Bucharest, Romania
2
National Centre for Micro and Nanomaterials, University Politehnica of Bucharest, 060042 Bucharest, Romania
3
National Research Center for Food Safety, University Politehnica of Bucharest, 060042 Bucharest, Romania
4
Multifunctional Materials and Structures Laboratory, National Institute of Materials Physics, 077125 Magurele, Romania
5
Ilie Murgulescu Institute of Physical Chemistry, 060021 Bucharest, Romania
6
Faculty of Veterinary Medicine, Department of Biochemistry, University of Agronomic Science and Veterinary Medicine, 011464 Bucharest, Romania
*
Author to whom correspondence should be addressed.
Nanomaterials 2019, 9(9), 1187; https://doi.org/10.3390/nano9091187
Submission received: 20 July 2019 / Revised: 10 August 2019 / Accepted: 18 August 2019 / Published: 22 August 2019

Abstract

:
Luminescent europium-doped hydroxylapatite (EuXHAp) nanomaterials were successfully obtained by co-precipitation method at low temperature. The morphological, structural and optical properties were investigated by scanning electron microscopy (SEM), transmission electron microscopy (TEM), X-ray diffraction (XRD), Fourier Transform Infrared (FT-IR), UV-Vis and photoluminescence (PL) spectroscopy. The cytotoxicity and biocompatibility of EuXHAp were also evaluated using MTT (3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide)) assay, oxidative stress assessment and fluorescent microscopy. The results reveal that the Eu3+ has successfully doped the hexagonal lattice of hydroxylapatite. By enhancing the optical features, these EuXHAp materials demonstrated superior efficiency to become fluorescent labelling materials for bioimaging applications.

1. Introduction

Bioceramics can be defined as the category of ceramics used in repairing and replacing processes of damaged and diseased parts of skeletal system [1,2]. Their biocompatibility varies from inert ceramic oxides to bioresorbable materials. One of the most used bioresorbable ceramics for biomedical applications are calcium orthophosphates (CaP’s) [3,4]. The CaP’s give hardness and stability to the tissues and can be found in teeth, bones, and tendons. Starting with Ca/P molar ratio equal to 0.5 and finishing with 2.0, there are 11 known non-ion substituted calcium orthophosphates, among which the most used one is hydroxylapatite (HAp) [5,6].
HAp synthesis, with its diverse morphologies, structures and textures, has attracted much interest in academic and industrial research for many heterogeneous catalysis applications [7,8,9]. Numerous synthetic routes for obtaining hydroxylapatite were developed over time, and can be divided in four main categories: (1) wet methods [10,11,12], (2) dry methods [13], (3) microwave-assisted methods [14,15,16], ball-milling [17,18,19] or ultrasound methods [20,21], and (4) miscellaneous methods [22]. Depending on the reagents and conditions, each category offers several variations [23,24].
Stoichiometric hydroxylapatite (Ca10(PO4)6(OH)2) is the most similar material to the mineral component of human hard tissues and therefore it is considered the ideal substance for bone defects restorations [25,26,27]. Keeping the same geometry while accepting a big variety of anions and cations is one of the most important structural characteristics of hydroxylapatite [28,29]. Synthetic HAp can also be doped with several metal ions in order to improve its properties, like bioactivity, degradation rate, antibacterial characteristics, luminescence and magnetic properties [30,31,32,33,34].
A photoluminescent material is the most promising candidate for clinical applications and implantation. The biocompatibility is not the only important feature, a longer lifetime of luminescence being also an significant benefit in practical applications [35,36].
Photoluminescence is a very important and useful mechanism in in situ investigations for tissue engineering, surgery, tissue restoration, etc. Using of organic fluorescent molecules for labelling it was a popular practice in clinical trials for many years. Recent studies use inorganic components, even in form of nanoparticles, to replace photoluminescent organic compounds. Because of the toxicity and the nano-size of this particles, the usage of these materials represents a challenge yet [37,38].
Due to high values of the Stokes shift and long lifetime of the excited state, lanthanide coordination compounds are the perfect materials for bioimaging. Europium complexes possess this unique luminescent properties and beside this Europium combine high photoluminescence quantum yields (PLQYs) with the emission in the long wavelength range, which can easily penetrate through the tissues [39,40,41].
The luminescence of the europium (III) and terbium (III) complexes is especially sensitive to changes in the structure and coordination environment of ions and depends substantially on the interaction with the analyte. The intensive luminescence and characteristic Stark structure of Eu3+ luminescence spectra allow registering fine changes in the structure of the coordination sphere of a rare-earth ion upon surrounding impact. Luminescent lanthanide-containing complex compounds can be applied as optical chemosensors in detection of anions, cations, gases, etc. [42].
The new generation of biomaterials with multifunctional europium (III)-doped HAp scaffolds has shown remarkable development. The luminescent multifunctional biomaterials show potential for use in various biomedical applications such as smart drug delivery, bioimaging, and photothermal therapy. In this generation, the Eu3+ ion has been widely used as traceable fluorescence probe due to well-known dopant narrow emission spectral lines by visible-light excitation caused by shielding by the 5 s and 5 p orbitals. The spectral shapes depend on the local ion symmetry and forbidden f–f transitions. Thus, Eu3+ is a sensitive optical probe for the dopant site environment because of its characteristic luminescence properties [43].

2. Experimental

2.1. Materials

All the reagents for synthesis, including calcium nitrate tetrahydrate (Ca(NO3)2·4H2O, 99.0%, Sigma-Aldrich, St. Louis, MI, USA), europium-(III) nitrate hexahydrate (Eu(NO3)3·6H2O, 99.9%, Alfa Aesar, Haverhill, MA, USA), ammonium phosphate dibasic (NH4)2HPO4, 99.0%, Alfa Aesar, Haverhill, MA, USA), sodium hydroxide (NaOH, 25% solution, Alfa Aesar, Haverhill, MA, USA) were used as received, without further purification. Deionised water was used for the experiment.

2.2. Synthesis

The biocompatible photoluminescent europium-doped hydroxylapatite (EuXHAp) nanomaterials have been synthesized by co-precipitation method. In order to obtain EuXHAp powders, appropriate amounts of calcium nitrate tetrahydrate and europium-(III) nitrate hexahydrate were dissolved in deionized water, under vigorous stirring at room temperature, thus obtaining solution A. Meanwhile, a solution B was prepared by dissolving an appropriate amount of ammonium phosphate dibasic in deionized water, under vigorous stirring at room temperature. The atomic ratio Ca/P and [Ca+Eu]/P was 1.67, while the atomic ratio Eu/(Eu+Ca) was varied between 0 and 50%. Solution B was added dropwise to solution A, at 80 °C, under vigorous stirring, while adjusting and maintaining the pH of the resulting suspension at 10, by adding NH4OH (25%) solution, for 2 h. Pure HAp was synthesized following the same methodology, except the Eu3+ precursor addition. The resulting suspensions were matured for 24 h. The precipitate was then filtered and washed several times with deionized water, until the pH values were close to 7. Finally, the resulting precipitates were dried at 80 °C for 50 h in an air oven.

2.3. Samples Notation

EuXHAp represents Eu–doped hydroxylapatite, where X is equal to the used europium content (XEu = 0.05, 0.1, 0.15, 0.2, 1.0, 5.0). The correspondent Eu-free hydroxylapatite is noted HAp (XEu = 0).

2.4. Morphological and Structural Characterization

X-ray diffraction (XRD) studies were carried out using a PANalytical Empyrean diffractometer at room temperature, with a characteristic Cu X-ray tube (λ Cu Kα1 = 1.541874 Ǻ) with in-line focusing, programmable divergent slit on the incident side and a programmable anti-scatter slit mounted on the PIXcel3D detector on the diffracted side. The samples were scanned in a Bragg - Brentano geometry with a scan step increment of 0.02° and a counting time of 255 s/step. The XRD patterns were recorded in the 2θ angle range of 20°–80°. Lattice parameters were refined by the Rietveld method, using the HighScore Plus 3.0 e software. The morphology of the samples was analyzed using a Quanta Inspect F50 FEG (field emission gun) scanning electron microscope with 1.2 nm resolution, equipped with an energy-dispersive X-ray (EDX) analyzer (resolution of 133 eV at MnKα, Thermo Fisher, Waltham, MA, USA) on sample covered with a thin gold layer. The high-resolution TEM images of the samples were obtained on finely powdered samples using a Tecnai G2 F30 S-Twin high-resolution transmission electron microscope from Thermo Fisher (former FEI) (Waltham, MA, USA).The microscope operated in transmission mode at 300 kV acceleration voltage with a TEM resolution of 1.0 Å. FTIR spectra were recorded with a Nicolet iS50R spectrometer (Thermo Fisher Waltham, MA, USA), at room temperature, in the measurement range 4000–400 cm−1. Spectral collection was carried out in ATR mode at 4 cm−1 resolution. For each spectrum, 32 scans were co-added and converted to absorbance using OmincPicta software (Thermo Scientific, Waltham, MA, USA). Raman spectra were recorded at room temperature on a Horiba Jobin-Yvon LabRam HR spectrometer equipped with nitrogen cooled detector. The near–infrared (NIR) line of a 785 nm laser was employed for excitation and the spectral range went from 500 to 1200 cm−1. UV-Vis diffused reflectance spectra were obtained using an Able Jasco V-560 spectrophotometer (PW de Meern, Netherlands,) with a scan speed of 200 nm/s, between 200 and 850 nm. The fluorescence spectra were measured by using a Perkin Elmer LS 55 fluorescence spectrophotometer (Arkon, OH, USA). Spectra were recorded with a scan speed of 200 nm/s between 350 and 800 nm, and with excitation and emission slits widths of 7 and 5 nm, respectively. An excitation wavelength of 320 nm was used.

2.5. Cellular Viability Assays

2.5.1. Quantitative In-Vitro Evaluation of Biocompatibility—MTT Assay

MTT assay is a quantitative colorimetric method, which allows evaluation of cell viability and proliferation, and cytotoxicity of different compounds. The method is based on reduction of MTT tetrazolium salt (3-(4,5-dimethylthiazolyl)-2,5-diphenyltetrazolium bromide) to dark blue formazan. Reduction by mitochondrial enzymes (especially succinate dehydrogenase) is an indication of cell/mitochondrial integrity. Formazan, insoluble in water, can be solubilized with isopropanol, dimethylsulfoxide or other organic solvent. The optical density (DO) of solubilized formazan is evaluated spectrophotometrically, resulting in a color-absorbent-color-counting function of the number of metabolic active cells in the culture.
The human mesenchymal amniotic fluid stem cells (AFSC) were used to evaluate the biocompatibility of EuXHAp nanoparticles. The cells were cultured in DMEM medium (Sigma-Aldrich, Saint Luis, MI, USA) supplemented with 10% fetal bovine serum, 1% penicillin and 1% streptomycin antibiotics (Sigma-Aldrich, Saint Luis, MI, USA). To maintain optimal culture conditions, medium was changed twice a week. The biocompatibility was assessed using MTT assay (Vybrant® MTT Cell Proliferation Assay Kit, Thermo Fischer Scientific, Waltham, MA, USA). Briefly, the AFSC were grown in 96-well plates, with a seeding density of 3000 cells/well in the presence of EuXHAp for 72 h. Then 15 mL Solution I (12 mM MTT) was added and incubated at 37 °C for 4 h. Solution II (1 mg Sodium Dodecyl Sulphate + 10 ml HCl, 0,01M) was added and pipettes vigorously to solubilize formazan crystals. After 1 h the absorbance was read using spectrophotometer at 570 nm (TECAN Infinite M200, Männedorf, Switzerland).

2.5.2. Oxidative Stress Assessment—GSH-Glo Glutathione Assay

The GSH-Glo Assay is a luminescent-based assay for the detection and quantification of glutathione (GSH) in cells or in various biological samples. A change in GSH levels is important in the assessment of toxicological responses and is an indicator of oxidative stress, potentially leading to apoptosis or cell death. The assay is based on the conversion of a luciferin derivative into luciferin in the presence of GSH. The reaction is catalyzed by a glutathione S-transferase (GST) enzyme supplied in the kit. The luciferin formed is detected in a coupled reaction using Ultra-Glo Recombinant Luciferase that generates a glow type luminescence that is proportional to the amount of glutathione present in cells. The assay provides a simple, fast and sensitive alternative to colorimetric and fluorescent methods and can be adapted easily to high-throughput applications.
AFSC were seeded at a density of 3000 cells in 300 µL of Dulbecco’s Modified Eagle’s medium (DMEM) supplemented with 10% fetal bovine serum and 1% antibiotics (penicillin, streptomycin/neomycin) in 96-well plates. Twenty-four hours after seeding, cells are treated with EuXHAp and incubated for 72 h.
The working protocol consisted of adding 100 µL 1X GSH-Glo Reagent and incubating at 37 °C for 30 min. Then, 100 µL Luciferin DeectionReagent was added and incubated at 37 °C for an additional 15 min. At the end of the time, the wells were well homogenized and then the plate was read on the luminometer (MicroplateLuminometerCentro LB 960, Berthold, Germany).

2.5.3. Qualitative In-Vitro Evaluation of Biocompatibility—Fluorescent Microscopy

The biocompatibility of the EuXHAp was also evaluated by fluorescent microscopy, using RED CMTPX fluorophore (Thermo Fischer Scientific, Waltham, MA, USA), a cell tracker for long-term tracing of living cells. The CMTPX tracker was added in cell culture treated with EuXHAp nanomaterials and the viability and morphology of the AFSC was evaluated after 5 days. The CMTPX fluorophore was added in the culture medium at a final concentration of 5 µM, incubated for 30 min in order to allow the dye penetration into the cells. Next, the AFSC were washed with PBS and visualized by fluorescent microscopy. The photomicrographs were taken with Olympus CKX 41 digital camera driven by CellSense Entry software (Olympus, Tokyo, Japan).

3. Results and Discussions

3.1. X-ray Diffraction

The X-ray diffraction patterns of Eu-doped hydroxylapatite, with different concentration of europium and with pure HAp are shown in Figure 1.
Pristine HAp (Figure 1) is shown as a single phase calcium phosphate material (ICDD PDF4+ no.04-021-1904 [44]), with a crystallinity of 33.9% and an average crystallite size of 10.84 nm (Table 1). The XRD patterns of all EuXHAp samples indicate only the pure crystalline hexagonal HAp phase (according to ICDD PDF4+ no.04-021-1904 [44] of the space group P63/m, in consistence with literature [45] up to a substitution degree of 10%. For 50% substitution, Eu(OH)3 secondary phase in a mass ratio of 27.6% has occurred (according to ICDD PDF4+ no. 01-083-2305 [46]). This means that for 50% substitution, the limit of solubility of Eu in HAp lattice was exceeded. The intensities of X-ray peaks decrease when the Eu-doping level increases up to 2%, indicating an interference of Eu3+ with HAp crystal structure. Also, the peak position is influenced by Eu for Ca substitution as they shift to higher angle values which suggest decreasing of unit cell parameters. Table 2 and Figure 2 illustrate the values of unit cell parameters a, c and V and the agreement indices of the Rietveld analysis (Rexp, Rp, Rwp and χ2), which indicate the quality of the fit.
Hydroxylapatite crystallizes in a hexagonal symmetry with lattice parameters a = bc and space group P63/m. In this structure, PO4 tetrahedral form basic structural units, while the coordination around the distinct Ca sites defines the Ca1O13O23 metaprism and the distorted Ca2O1O2O34(OH) polyhedron. Thus, the formula per unit cell may be expressed as Ca14Ca26(PO1O2O32)6(OHH)2 [47]. Taking into consideration that the ionic radius of Eu3+ (0.947 Å) is smaller than that of Ca2+ (1 Å) and that small ions are preferentially substituted at Ca1, in this case Eu3+ most probably substitutes Ca2+ from site 2 [48,49]. While a low concentration does not induce significant changes (Figure 2), when looking at higher substitution degree (10% and 50%) it may be concluded that there is an increase of a-axis and a decrease of c-axis parameters which is consistent with the results obtained for Al3+ substitution by Fahami et al. [49]. Moreover, the decrease of unit cell volume proves the incorporation of Eu3+ in hydroxylapatite lattice.
The estimation of crystallite size, lattice microstrain and degree of crystallinity are shown in Table 2 and Figure 3.
Up to 1.5%, the introduction of Eu3+ in HAp lattice induces an increase of crystallite size from 9.62 nm to 17.3 nm and a decrease of lattice microstrain from 0.95% to 0.65%. For a substitution degree of more than 1.5%, there may be observed a decrease of the crystallite size to 3.55 nm and an increase of lattice microstrain to 2.66% in the case of maximum substitution degree. In what concerns the crystallinity degree (Table 2), as it may be appreciated qualitatively from the profile of the peaks, the crystallinity decreases when the substitution degree increases.

3.2. SEM Analysis

Figure 4 shows the SEM images (column A) and EDX spectra (column B) of pure HAp (Figure 4a) and EuXHAp samples (Figure 4b).
It can be seen that the doping with Eu3+ has little influence on the morphology of substituted HAp compared to the pure HAp. SEM images (Figure 4, column A) reveal quasi-spherical (at low dopant concentration) and acicular (mostly after 10% Eu) particles with width in the range of 6–16 nm. Due to high surface area, agglomerates are present in all samples. The EDX spectra of studied samples (Figure 4, column B) confirm the presence of all elements specific to Eu-doped HAp powders: calcium (Ca), phosphor (P), oxygen (O) and europium (Eu).

3.3. TEM Analysis

Figure 5 shows the bright field TEM, HRTEM images, SAED patterns and particle size distribution of pure HAp and EuXHAp at different Eu3+ concentrations.
From the bright field TEM images presented in Figure 5a,d,g,j,m,p,s it can be observed that the doping level does not affect the the sample morphologies from 0 to 10% Eu3+ doping, the only visible modification if for the 50% Eu3+ doping and it is related to the width (smaller) and length (longer) of the acicular particles. The SEM micrographs together with TEM images confirmed the tendency to form agglomerate due to probably the nanometric dimensions of particles. The size distribution presented in Figure 5c,f,i,l,o,r,u is depended on the dopant concentration. The HRTEM images of the Eu-doped HAp powders are also presented in Figure 5 which allows us to see that the particles are well crystalize and the measured distance of the Miller indices correspond to the hexagonal hydroxylapatite.

3.4. FTIR Spectra

The FTIR spectra of EuXHAp samples with various europium concentrations are shown in Figure 6.
The broad band in the region 3200–3400 cm−1 corresponds to [OH]- bands of adsorbed water. The strong band at 632 cm-1 (presented in all FTIR spectra) corresponds to [OH]- arising from stretching vibrational mode [50,51]. The typical bands attributed to [PO4]3− can be also found. The bands at around 1090 cm−1 and about 1040 cm−1 may be assigned to the antisymetric stretching ν3[PO4]3− of P-O bond, while the band at 962 can be due to the symmetric stretching ν1[PO4]3−. The 602 cm−1 and 564 cm−1 bands appear from ν4[PO4]3− of P-O bond. The band at 475 cm−1 can be attributed to the ν2[PO4]3− [52]. In all spectra of EuXHAp a band at 875 cm−1 was detected, and is due to [HPO4]2− ions [53]. The intensity of the phosphate bands decreases with increase of europium concentration. The bands at 475 and 962 cm−1 progressively reduced their intensities with the increase of europium concentration, and disappear in Eu50HAp sample, as already suggested also by XRD results. The different possible mechanisms of Ca substitution with Eu are not fully comprehended and need further studies

3.5. Raman Spectroscopy

The Raman spectra of EuXHAp samples are shown in Figure 7.
As it can be seen from Figure 7 all analyzed powders present bands attributed to PO43− group [54]. Raman spectrum of pure HAp shows a very intense band at 959 cm−1 attributed to symmetric stretching mode of PO43−, which is the characteristic peak of HAp. The asymmetric stretching and symmetric bending modes of PO43− are also observed at 580 cm−1 and 1044 cm−1 [53]. Raman spectra of EuXHAp powders show the internal modes of the frequency ν1 PO43− tetrahedral, which appears at 962 cm−1 and it corresponds to the symmetric stretching of P–O bonds [55]. There are observed several other bands at 580 cm−1 and 610 cm−1 attributed to ν4 PO43−, and 1048 cm−1 and 1080 cm−1, respectively, attributed to ν3 PO43−. The intensity of the characteristic peak of HAp, compared to the other secondary peaks, decreases with increasing of the europium content.

3.6. UV-Vis and PL Spectra

The electronic spectra of EuXHAp samples (Figure 8) contain several bands, which increase in their intensity with increasing of europium content.
The pure HAp has an absorption peak in UV at 218 nm. The 242 nm (41322 cm−1) absorption peak is absent in HAp, but is increasing in intensity from a weak shoulder in Eu0.5HAp to a strong, broad peak in Eu50HAp. This is a charge-transfer band characteristic to Eu3+ in oxides [56]. The band which appears at 395 nm, close to the visible light domain, corresponds to the 7F05L6 transition, and is the most intense transition of europium (III) in UV-Vis absorption spectra [57]. Another three bands appear at 465, 526 and 535 nm, assignable to 7F05D2; 7F05D1; and 7F15D1, transitions respectively [56]. The last peak, at 535 nm, is belonging to a group called “hot” bands because it can be observed only at room temperatures or higher since it requires the thermal population of 7F1 level (at room temperature ~35% of ions are populating this level, rest being on 7F0 ground state) [56].
Figure 8B presents the photoluminescence emission spectra of EuXHAp samples excited with 320 nm wavelength. Usually, HAp sample shows three strong emission peaks at 458 nm with two shoulders (433 nm and 446 nm), 403 nm and 482 nm, respectively, associated with various oxygen defects. The intensity of luminescence for HAp is enhanced by the presence of small quantities of Eu3+ ions. The effect is stronger for small numbers of dopant ions and is decreasing as the quantity of Eu3+ ions is increasing.
Over the fluorescence maxima of HAp, the emission spectra of europium are superposed. The peak at 389 nm is due to the 7F05L6 electronic transition, the peak at 512 nm is due to the 7F05D1 electronic transition and the peak at 526 nm is due to the 7F15D1 electronic transition. The characteristic emission peaks at 589 nm and 613 nm are due to the 5D07F1 and 5D07F2 electronic transition, indicating a red fluorescence [58]. For higher Eu3+ concentration even the 648 nm emission peak of 5D07F3 becomes visible. As expected, when the europium doping concentration increased, the characteristic emission intensity at 589 and 613 nm are also increased.
Beside the two dominant peaks, the weak 5D07F0 transition from 579 nm (which can be observed in emission spectra of Eu50HAp) is related with Eu3+ ions distributed on Ca2+ sites of the apatitic structure [59]. The dominant emission peaks present no shift due to modification of Eu3+ concentration, the most intense being the hypersensitive 5D07F2 transition from 613 nm. This feature shows the potential application of the EuXHAp compounds to be tracked or monitored by the characteristic luminescence.

4. Biocompatibility and Cytotoxicity of EuXHAp Photoluminescent Ceramic Materials

Cytotoxic effect of EuXHAp was evaluated by measuring the metabolic activity of AFSC using MTT assay. EuXHAp biomaterials did not have cytotoxic effect, the absorption values being near or higher compared to the control sample. Furthermore, the EuXHAp increases cellular metabolism, suggesting that it stimulates cell proliferation. The proliferation increases between 12%–58%, depending on the sample, highest increase being observed for Eu2HAp (58%) (Figure 9). The results are similar to other research groups that showed the viability of HGF-1 fibroblast was decreased in the Eu10Hap after 48 h. Naderi et al. 2012 [60] tested different concentrations of nanohydroxyapatite from 2 to 0.002  mg/mL on gingiva-derived fibroblast cell line (HGF-2) at 24, 48, and 72 h, and concluded that after 24 h high doses of nanohydroxyapatite have cytotoxic effect on gingival-derived fibroblasts suggesting that cytotoxicity is dependent on the cell line [60,61].
Glutathione, an oxidative stress marker, is capable of preventing cellular damage caused by reactive oxygen species, such as free radicals, peroxides, lipid peroxides and heavy metals. In the presence of EuXHAp biomaterials, AFSC responded similarly to control cells, indicating that the analysed materials did not induce cellular stress (Figure 10). Furthermore, the morphology of AFSC was investigated by fluorescence microscopy using CMTPX cell tracker for long-term tracing of living cells. Cellular metabolism is active, as shown in microscopy images, cells absorbing CMTPX fluorophore in the cytoplasm, suggesting that they are viable.
After 5 days in the presence of EuXHAp, AFSCs presents a normal morphology with fibroblastic-like characteristic appearance (Figure 11). Fluorescent images show that AFSC cells are viable, no dead cells or cell fragments are observed, more the cells spread filopodia to move and establish contacts with neighboring cells, suggesting that AFSC have an active phenotype.

5. Conclusions

In recent years, a special attention has been created regarding to multimodal imaging. This technique represents various imaging modalities in the manner of photoluminescence and magnetic resonance imaging that are generally connected in an individual diagnostic step. If we talk about multimodal contrast agents that can be used or marker for biomedical imaging with luminescence, literature data provides information on the use of doped hydroxyapatite with rare earth elements like europium of nanometric dimensions.
The aim of the present study was to obtain europium-doped nano hydroxylapatite by using a simple method of synthesis (co-precipitation method) and, thereafter characterize by physico-chemical techniques and also by biological point of view because of the fact that the obtained material has medical application.
The results after the XRD analysis shows that the obtained material is represented by pure hexagonal phase of hydroxylapatite, when using up to 10% dopant Eu3+ ions concentration, but there is a correlation between HAp and Eu3+ by the interference of Eu3+ ions with HAp crystal structure. About the morphology of HAp doped with Eu3+ it can be concluded that there is a limited modification after the addition of Eu3+ and also by the EDS spectra is confirmed the presence of Eu3+ in all obtained materials. Typically, the HAp sample shows three strong emission peaks, but the photoluminescence intensities is enhanced by the characteristic Eu3+ peaks with the presence of Eu3+ ions. This feature is required and necessary to have a good action in medical imaging.
From the biological point of view, it can be seen by the MTT assay results that the obtained material supports the proliferation process of the amniotic fluid stem cells. Also, the best results about the viability characteristic is offered by the Eu2HAp material which can be a standard for the doped HAp, because after the addition of more Eu3+ ions there is a small evidence about the cellular viability that can decrease. As a complete evaluation, the qualitative information by the fluorescence microscopy gives characteristic about the biocompatibility of the material in light of the fact that after 5 days of incubation with amniotic fluid stem cells, no dead cells are observed.
Therefore, there is a fine line about the using doped HAp with Eu3+ as fluorescent or multimodal contrast agent, but the results can be a promising start due to its characteristics but there is a need for the further investigation.

Author Contributions

The authors have participated to the paper as follows; conceptualization, B.S.V., D.P.; methodology, E.A., B.S.V., D.P., S.L.I.; validation, E.A., B.S.V.; formal analysis, I.A.N., A.P., R.T., O.O., E.T., O.R.V., A.I.N., A.V.S., F.I., A.C.B., S.L.I.; investigation, I.A.N., R.T., O.O., E.T., O.R.V., A.I.N., A.V.S., F.I., A.C.B.; data curation, A.V.P., A.M.M.; writing—original draft preparation, A.M.M, A.V.P.; writing—review and editing, I.A.N., B.S.V., A.M.M.; visualization, E.A.; supervision, E.A., B.S.V.; project administration, E.A.

Funding

This research was funded by project “Innovative biomaterials for treatment and diagnosis” BIONANOINOV—P3 grant number PN-IIIP1-1.2-PCCD-I2017-0629, and the support of the EU-funding project POSCCE-A2-O2.2.1-2013-1/Priority Axe 2, Project No. 638/12.03.2014, ID 1970, SMIS-CSNR code 48652 is gratefully acknowledged for the equipment’s purchased from this project.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhou, Y.; Wu, C.; Chang, J. Bioceramics to regulate stem cells and their microenvironment for tissue regeneration. Mater. Today 2019, 24, 41–56. [Google Scholar] [CrossRef]
  2. Ben-nissan, B.; Cazalbou, S.; Choi, A. Bioceramics. Biomater. Sci. Eng. 2019, 1, 16–33. [Google Scholar]
  3. Dorozhkin, S.V. Calcium orthophosphates (CaPO4): Occurrence and properties (Les orthophosphates de calcium (CaPO4): Occurrence et Proprieties). Morphologie 2017, 101, 125–142. [Google Scholar] [CrossRef] [PubMed]
  4. Dorozhkin, S.V. A history of calcium orthophosphates (CaPO4) and their biomedical applications. Morphologie 2017, 101, 143–153. [Google Scholar] [CrossRef] [PubMed]
  5. Pelin, I.M.; Maier, V.; Suflet, D.M.; Popescu, I.; Darie-Nita, R.N.; Aflori, M.; Butnaru, M. Formation and characterization of calcium orthophosphates in the presence of two different acidic macromolecules. J. Cryst. Growth 2017, 475, 266–273. [Google Scholar] [CrossRef]
  6. Sathiskumar, S.; Vanaraj, S.; Sabarinathan, D.; Bharath, S. Green synthesis of biocompatible nanostructured hydroxyapatite from Cirrhinus mrigala fish scale—A biowaste to biomaterial. Ceram. Int. 2019, 45, 7804–7810. [Google Scholar] [CrossRef]
  7. Beau, S.; Rouillon, T.; Millet, P.; Le, J.; Weiss, P.; Chopart, J.; Daltin, A. Synthesis of calcium-deficient hydroxyapatite nanowires and nanotubes performed by template-assisted electrodeposition. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 98, 333–346. [Google Scholar]
  8. Xia, X.; Chen, J.; Shen, J.; Huang, D.; Duan, P.; Zou, G. Synthesis of hollow structural hydroxyapatite with different morphologies using calcium carbonate as hard template. Adv. Powder Technol. 2018, 29, 1562–1570. [Google Scholar] [CrossRef]
  9. Hashemi, S.; Javadpour, J.; Khavandi, A.; Erfan, M. Morphological evolution on the surface of hydrothermally synthesized hydroxyapatite microspheres in the presence of EDTMP. Ceram. Int. 2018, 44, 19743–19750. [Google Scholar]
  10. Turan, İ.; Kaygili, O.; Tatar, C.; Bulut, N.; Koytepe, S.; Ates, T. The effects of Ni-addition on the crystal structure, thermal properties and morphology of Mg-based hydroxyapatites synthesized by a wet chemical method. Ceram. Int. 2018, 44, 14036–14043. [Google Scholar]
  11. Wolff, J.; Hofmann, D.; Amelung, W.; Lewandowski, H.; Kaiser, K.; Bol, R. Applied Geochemistry Rapid wet chemical synthesis for 33 P-labelled hydroxyapatite—An approach for environmental research. Appl. Geochem. 2018, 97, 181–186. [Google Scholar] [CrossRef]
  12. Yelten-yilmaz, A.; Yilmaz, S. Wet chemical precipitation synthesis of hydroxyapatite (HA) powders. Ceram. Int. 2018, 44, 9703–9710. [Google Scholar] [CrossRef]
  13. Jiao, Y.; Lu, Y.; Xiao, G.; Xu, W.; Zhu, R. Preparation and characterization of hollow hydroxyapatite microspheres by the centrifugal spray drying method. Powder Technol. 2012, 217, 581–584. [Google Scholar] [CrossRef]
  14. Sabu, U.; Logesh, G.; Rashad, M.; Joy, A.; Balasubramanian, M. Microwave assisted synthesis of biomorphic hydroxyapatite. Ceram. Int. 2019, 45, 6718–6722. [Google Scholar] [CrossRef]
  15. Ghorbani, F.; Zamanian, A.; Behnamghader, A.; Daliri-joupari, M. Bone-like hydroxyapatite mineralization on the bio-inspired PDA nanoparticles using microwave irradiation. Surf. Interfaces 2019, 15, 38–42. [Google Scholar] [CrossRef]
  16. Anwar, A.; Akbar, S. Novel continuous microwave assisted flow synthesis of nanosized manganese substituted hydroxyapatite. Ceram. Int. 2018, 44, 10878–10882. [Google Scholar] [CrossRef]
  17. Wu, S.; Hsu, H.; Hsu, S.; Chang, Y.; Ho, W. Synthesis of hydroxyapatite from eggshell powders through ball milling and heat treatment. J. Asian Ceram. Soc. 2016, 4, 85–90. [Google Scholar] [CrossRef]
  18. Hannora, A.E.; Ataya, S. Structure and compression strength of hydroxyapatite/titania nanocomposites formed by high energy ball milling. J. Alloy. Compd. 2016, 658, 222–233. [Google Scholar] [CrossRef]
  19. Ferro, A.C.; Guedes, M. Mechanochemical synthesis of hydroxyapatite using cuttle fish bone and chicken eggshell as calcium precursors. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 97, 124–140. [Google Scholar] [CrossRef]
  20. Bouyarmane, H.; Gouza, A.; Masse, S.; Saoiabi, S.; Saoiabi, A.; Coradin, T.; Laghzizil, A. Nanoscale conversion of chlorapatite into hydroxyapatite using ultrasound irradiation. Colloids Surf. A Physicochem. Eng. Asp. 2016, 495, 187–192. [Google Scholar] [CrossRef] [Green Version]
  21. Nikolaev, A.L.; Gopin, A.V.; Severin, A.V.; Rudin, V.N.; Mironov, M.A.; Dezhkunov, N.V. Ultrasonic synthesis of hydroxyapatite in non-cavitation and cavitation modes. Ultrason. Sonochem. 2018, 44, 390–397. [Google Scholar] [CrossRef] [PubMed]
  22. Valle, L.J.; Katsarava, R. Other Miscellaneous Materials and Their Nanocomposites; Elsevier Inc.: Amsterdam, The Netherlands, 2019; ISBN 9780128146156. [Google Scholar]
  23. Fihri, A.; Len, C.; Varma, R.S.; Solhy, A. Hydroxyapatite: A review of syntheses, structure and applications in heterogeneous catalysis. Coord. Chem. Rev. 2017, 347, 48–76. [Google Scholar] [CrossRef]
  24. Sadat-Shojai, M.; Khorasani, M.T.; Dinpanah-Khoshdargi, E.; Jamshidi, A. Synthesis methods for nanosized hydroxyapatite with diverse structures. Acta Biomater. 2013, 9, 7591–7621. [Google Scholar] [CrossRef]
  25. Kubasiewicz-Ross, P.; Hadzik, J.; Seeliger, J.; Kozak, K.; Jurczyszyn, K.; Gerber, H.; Dominiak, M.; Kunert-Keil, C. New nano-hydroxyapatite in bone defect regeneration: A histological study in rats. Ann. Anat. 2017, 213, 83–90. [Google Scholar] [CrossRef] [PubMed]
  26. Munhoz, M.A.S.; Hirata, H.H.; Plepis, A.M.G.; Martins, V.C.A.; Cunha, M.R. Use of collagen/chitosan sponges mineralized with hydroxyapatite for the repair of cranial defects in rats. Injury 2018, 49, 2154–2160. [Google Scholar] [CrossRef] [PubMed]
  27. Solonenko, A.P.; Blesman, A.I.; Polonyankin, D.A. Preparation and in vitro apatite-forming ability of hydroxyapatite and β-wollastonite composite materials. Ceram. Int. 2018, 44, 17824–17834. [Google Scholar] [CrossRef]
  28. Kaur, K.; Singh, K.J.; Anand, V.; Islam, N.; Bhatia, G.; Kalia, N.; Singh, J. Lanthanide (=Ce, Pr, Nd and Tb) ions substitution at calcium sites of hydroxyl apatite nanoparticles as fluorescent bio probes: Experimental and density functional theory study. Ceram. Int. 2017, 43, 10097–10108. [Google Scholar] [CrossRef]
  29. Safarzadeh, M.; Ramesh, S.; Tan, C.Y.; Chandran, H.; Fauzi, A.; Noor, M.; Krishnasamy, S.; Alengaram, U.J.; Ramesh, S. Effect of multi-ions doping on the properties of carbonated hydroxyapatite bioceramic. Ceram. Int. 2019, 45, 3473–3477. [Google Scholar] [CrossRef]
  30. Tesch, A.; Wenisch, C.; Herrmann, K.H.; Reichenbach, J.R.; Warncke, P.; Fischer, D.; Müller, F.A. Luminomagnetic Eu3+- and Dy3+-doped hydroxyapatite for multimodal imaging. Mater. Sci. Eng. C Mater. Biol. Appl. 2017, 81, 422–431. [Google Scholar] [CrossRef]
  31. Zilm, M.E.; Yu, L.; Hines, W.A.; Wei, M. Magnetic properties and cytocompatibility of transition-metal-incorporated hydroxyapatite. Mater. Sci. Eng. C Mater. Biol. Appl. 2018, 87, 112–119. [Google Scholar] [CrossRef]
  32. Li, H.; Sun, X.; Li, Y.; Li, B.; Liang, C.; Wang, H. Preparation and properties of carbon nanotube (Fe)/hydroxyapatite composite as magnetic targeted drug delivery carrier. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 97, 222–229. [Google Scholar] [CrossRef] [PubMed]
  33. Riaz, M.; Zia, R.; Ijaz, A.; Hussain, T.; Mohsin, M.; Malik, A. Synthesis of monophasic Ag doped hydroxyapatite and evaluation of antibacterial activity. Mater. Sci. Eng. C Mater. Biol. Appl. 2018, 90, 308–313. [Google Scholar] [CrossRef] [PubMed]
  34. Kim, H.; Mondal, S.; Jang, B.; Manivasagan, P. Biomimetic synthesis of metal—Hydroxyapatite (Au-HAp, Ag-HAp, Au-Ag- HAp): Structural analysis, spectroscopic characterization and biomedical application. Ceram. Int. 2018, 44, 20490–20500. [Google Scholar] [CrossRef]
  35. Gonzalez, G.; Costa-vera, C.; Borrero, L.J.; Soto, D.; Lozada, L.; Chango, J.I.; Diaz, J.C.; Lascano, L. Effect of carbonates on hydroxyapatite self-activated photoluminescence response. J. Lumin. 2018, 195, 385–395. [Google Scholar] [CrossRef]
  36. Neacsu, I.A.; Stoica, A.E.; Vasile, B.S. Luminescent Hydroxyapatite Doped with Rare Earth Elements for Biomedical Applications. Nanomaterials 2019, 9, 239. [Google Scholar] [CrossRef] [PubMed]
  37. Liu, J.; Lécuyer, T.; Seguin, J.; Mignet, N.; Scherman, D.; Viana, B.; Richard, C. Imaging and therapeutic applications of persistent luminescence nanomaterials. Adv. Drug Deliv. Rev. 2019, 138, 193–210. [Google Scholar] [CrossRef] [PubMed]
  38. Xing, Q.; Zhang, X.; Wu, D.; Han, Y.; Wickramaratne, M.N.; Dai, H.; Wang, X. Ultrasound-Assisted Synthesis and Characterization of Heparin-Coated Eu3+ Doped Hydroxyapatite Luminescent Nanoparticles. Colloid Interface Sci. Commun. 2019, 29, 17–25. [Google Scholar] [CrossRef]
  39. Kudryashova, V.A. Europium and terbium pyrrole-2-carboxylates: Structures, luminescence, and energy transfer. Inorg. Chim. Acta 2019, 492, 1–7. [Google Scholar]
  40. Corbin, B.A.; Hovey, J.L.; Thapa, B.; Schlegel, H.B.; Allen, M.J. Luminescence differences between two complexes of divalent europium. J. Organomet. Chem. 2018, 857, 88–93. [Google Scholar] [CrossRef]
  41. Utochnikova, V.V.; Koshelev, D.S.; Medvedko, A.V.; Kalyakina, A.S.; Bushmarinov, I.S.; Grishko, A.Y.; Schepers, U.; Bräse, S.; Vatsadze, S.Z. Europium 2-benzofuranoate: Synthesis and use for bioimaging. Opt. Mater. 2017, 74, 191–196. [Google Scholar] [CrossRef]
  42. Petrochenkova, N.V.; Mirochnik, A.G.; Emelina, T.B.; Sergeev, A.A.; Leonov, A.A.; Voznesenskii, S.S. Luminescent amine sensor based on europium(III) chelate. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2018, 200, 70–75. [Google Scholar] [CrossRef] [PubMed]
  43. Lima, T.A.R.M.; Valerio, M.E.G. X-ray absorption fine structure spectroscopy and photoluminescence study of multifunctional europium (III)-doped hydroxyapatite in the presence of cationic surfactant medium. J. Lumin. 2018, 201, 70–76. [Google Scholar] [CrossRef]
  44. Mugnaioli, E.; Reyes-gasga, J.; Kolb, U.; Hemmerlø, J. Evidence of Noncentrosymmetry of Human Tooth Hydroxyapatite Crystals. Chem. A Eur. J. 2014, 20, 6849–6852. [Google Scholar] [CrossRef] [PubMed]
  45. Chen, M.; Tan, J.; Lian, Y.; Liu, D. Preparation of Gelatin coated hydroxyapatite nanorods and the stability of its aqueous colloidal. Appl. Surf. Sci. 2008, 254, 2730–2735. [Google Scholar] [CrossRef]
  46. Mullica, D.F.; Milligan, W.O.; Beall, G.W. Crystal structures of Pr(OH)3, Eu(OH)3 and Tm(OH)3. J. Inorg. Nucl. Chem. 1979, 41, 525–532. [Google Scholar] [CrossRef]
  47. Terra, J.; Dourado, E.R.; Eon, J.G.; Ellis, D.E.; Gonzalez, G.; Rossi, A.M. The structure of strontium-doped hydroxyapatite: An experimental and theoretical study. Phys. Chem. Chem. Phys. 2009, 11, 568–577. [Google Scholar] [CrossRef]
  48. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  49. Fahami, A.; Nasiri-Tabrizi, B.; Beall, G.W.; Basirun, W.J. Structural insights of mechanically induced aluminum-doped hydroxyapatite nanoparticles by Rietveld refinement. Chin. J. Chem. Eng. 2017, 25, 238–247. [Google Scholar] [CrossRef] [Green Version]
  50. Fowler, B. Infrared Studies of Apatites. I. Vibrational Assignments for Calcium, Strontium, and Barium Hydroxyapatites Utilizing Isotopic Substitution. Inorg. Chem. 1974, 13, 194–207. [Google Scholar] [CrossRef]
  51. Iconaru, S.; Motelica-Heino, M.; Predoi, D. Study on Europium-Doped Hydroxyapatite Nanoparticles by Fourier Transform Infrared Spectroscopy and Their Antimicrobial Properties. J. Spectrosc. 2013, 2013, 284285. [Google Scholar] [CrossRef]
  52. Sitarz, M.; Rokita, M.; Bułat, K. Infrared spectroscopy of different phosphates structures. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2011, 79, 722–727. [Google Scholar]
  53. Markovic, M.; Fowler, O.B. Preparation and Comprehensive Characterization of a Calcium Hydroxyapatite Reference Material Volume. J. Res. Natl. Inst. Stand. Technol. 2004, 109, 553–568. [Google Scholar] [CrossRef] [PubMed]
  54. Cuscó, R.; Guitián, F.; de Aza, S.; Arttis, L. Differentiation between Hydroxyapatite and/I-Tricalcium Phosphate by Means of p-Raman Spectroscopy. J. Eur. Ceram. Soc. 1998, 2219, 1301–1305. [Google Scholar] [CrossRef]
  55. Ciobanu, C.S.; Massuyeau, F.; Andronescu, E.; Stan, M.S.; Predoi, D. Biocompatibility study of europium doped crystalline hydroxyapatite bioceramics. Dig. J. Nanomater. Biostructures 2011, 6, 1639–1647. [Google Scholar]
  56. Binnemans, K. Interpretation of europium (III) spectra. Coord. Chem. Rev. 2015, 295, 1–45. [Google Scholar] [CrossRef]
  57. Al-kattan, A.; Dufour, P.; Dexpert-ghys, J.; Drouet, C. Preparation and Physicochemical Characteristics of Luminescent Apatite-Based Colloids. J. Phys. Chem. C 2010, 114, 2918–2924. [Google Scholar] [CrossRef]
  58. Yang, C.; Yang, P.; Wang, W.; Gai, S.; Wang, J.; Zhang, M.; Lin, J. Synthesis and characterization of Eu-doped hydroxyapatite through a microwave assisted microemulsion process. Solid State Sci. 2009, 11, 1923–1928. [Google Scholar] [CrossRef]
  59. Ciobanu, C.S.; Iconaru, S.L.; Massuyeau, F.; Constantin, L.V.; Costescu, A.; Predoi, D. Synthesis, Structure, and Luminescent Properties of Europium-Doped Hydroxyapatite Nanocrystalline Powders. J. Nanomater. 2012, 2012, 942801. [Google Scholar] [CrossRef]
  60. Noushin Jalayer Naderi, R.Y.; Jamali, D.; Mohammad Rezvani, B. Cytotoxicity of nano-hydroxyapatite on gingiva-derived fibroblast cell line (HGF2): An in vitro study. Daneshvar Med. 2012, 19, 79–86. [Google Scholar]
  61. Paulina-Guadalupe, M.-M.; Gabriel-Alejandro, M.-C.; Nereyda, N.-M.; Nuria, P.-M.; Miguel-Ángel, C.-S.; Brenda-Erendida, C.-S.; Facundo, R. Facile Synthesis, Characterization, and Cytotoxic Activity of Europium-Doped Nanohydroxyapatite. Bioinorg. Chem. Appl. 2016, 2016, 1057260. [Google Scholar]
Figure 1. X-ray diffraction patterns of hydroxylapatite (HAp) and EuXHAp.
Figure 1. X-ray diffraction patterns of hydroxylapatite (HAp) and EuXHAp.
Nanomaterials 09 01187 g001
Figure 2. Unit cell parameters versus substitution degree for HAp and EuXHAp.
Figure 2. Unit cell parameters versus substitution degree for HAp and EuXHAp.
Nanomaterials 09 01187 g002
Figure 3. The estimated crystallite size, lattice microstrain and degree of crystallinity.
Figure 3. The estimated crystallite size, lattice microstrain and degree of crystallinity.
Nanomaterials 09 01187 g003
Figure 4. The SEM images (column A) and EDX spectra (column B) of pure HAp and EuXHAp samples (a) HAp, (b) Eu0.5HAp, (c) Eu1HAp, (d) Eu1.5HAp, (e) Eu2HAp, (f) Eu10HAp, (g) Eu50HAp.
Figure 4. The SEM images (column A) and EDX spectra (column B) of pure HAp and EuXHAp samples (a) HAp, (b) Eu0.5HAp, (c) Eu1HAp, (d) Eu1.5HAp, (e) Eu2HAp, (f) Eu10HAp, (g) Eu50HAp.
Nanomaterials 09 01187 g004aNanomaterials 09 01187 g004bNanomaterials 09 01187 g004c
Figure 5. The TEM and HRTEM images, SAED patterns and particle size distribution of pure HAp and EuXHAp samples (ac) HAp, (df) Eu0.5HAp, (gi) Eu1HAp, (jl) Eu1.5HAp, (mo) Eu2HAp, (pr) Eu10HAp, (su) Eu50HAp.
Figure 5. The TEM and HRTEM images, SAED patterns and particle size distribution of pure HAp and EuXHAp samples (ac) HAp, (df) Eu0.5HAp, (gi) Eu1HAp, (jl) Eu1.5HAp, (mo) Eu2HAp, (pr) Eu10HAp, (su) Eu50HAp.
Nanomaterials 09 01187 g005aNanomaterials 09 01187 g005bNanomaterials 09 01187 g005cNanomaterials 09 01187 g005dNanomaterials 09 01187 g005e
Figure 6. FTIR spectra of EuXHAp samples.
Figure 6. FTIR spectra of EuXHAp samples.
Nanomaterials 09 01187 g006
Figure 7. Raman spectra of EuXHAp samples.
Figure 7. Raman spectra of EuXHAp samples.
Nanomaterials 09 01187 g007
Figure 8. (A) UV-Vis absorption spectra of EuXHAp at different Eu-concentrations (B) Room-temperature photoluminescence spectra of EuXHAp at different Eu-concentrations.
Figure 8. (A) UV-Vis absorption spectra of EuXHAp at different Eu-concentrations (B) Room-temperature photoluminescence spectra of EuXHAp at different Eu-concentrations.
Nanomaterials 09 01187 g008
Figure 9. MTT assay showing the viability of AFSC in the presence of the EuXHAp ceramic materials: HAp, Eu05HAp, Eu1HAp, Eu1.5HAp, Eu2HAp, Eu10HAp, Eu50HAp, and control (cell only).
Figure 9. MTT assay showing the viability of AFSC in the presence of the EuXHAp ceramic materials: HAp, Eu05HAp, Eu1HAp, Eu1.5HAp, Eu2HAp, Eu10HAp, Eu50HAp, and control (cell only).
Nanomaterials 09 01187 g009
Figure 10. GSH assay showing the oxidative stress of AFSC in the presence of the EuXHAp ceramic materials: HAp, Eu05HAp, Eu1HAp, Eu1.5HAp, Eu2HAp, Eu10HAp, Eu50HAp and control (cell only).
Figure 10. GSH assay showing the oxidative stress of AFSC in the presence of the EuXHAp ceramic materials: HAp, Eu05HAp, Eu1HAp, Eu1.5HAp, Eu2HAp, Eu10HAp, Eu50HAp and control (cell only).
Nanomaterials 09 01187 g010
Figure 11. Fluorescence images of EuXHAp samples coloured with CMTPX fluorophore (a) Control sample, (b) Eu0.5HAp, (c) Eu1HAp, (d) Eu1.5HAp, (e) Eu2HAp, (f) Eu10HAp, (g) Eu50HAp.
Figure 11. Fluorescence images of EuXHAp samples coloured with CMTPX fluorophore (a) Control sample, (b) Eu0.5HAp, (c) Eu1HAp, (d) Eu1.5HAp, (e) Eu2HAp, (f) Eu10HAp, (g) Eu50HAp.
Nanomaterials 09 01187 g011aNanomaterials 09 01187 g011b
Table 1. Calculated crystallite size (D) values and degree of crystallinity (Xc) of pure HAp and europium doped hydroxylapatite with various amount of Eu.
Table 1. Calculated crystallite size (D) values and degree of crystallinity (Xc) of pure HAp and europium doped hydroxylapatite with various amount of Eu.
No.SamplesD/nmS/%Xc/%
1HAp9.62 ± 1.460.95 ± 0.2933.93
2Eu0.5HAp14.88 ± 2.960.74 ± 0.6626.25
3Eu1HAp12.82 ± 2.720.84 ± 0.6624.11
4Eu1.5HAp17.3 ± 4.920.65 ± 0.5522.97
5Eu2HAp14.44 ± 3.560.77 ± 0.6322.63
6Eu10HAp10.78 ± 3.041.04 ± 0.8523.03
7Eu50HAp3.55 ± 0.342.66 ± 1.5621.87
Table 2. Unit cell parameters a, c, V and agreement indices for hydroxylapatite HAp and Eu-doped HAp (with concentration of Eu3+ of 0.5, 1, 1.5, 2, 10 and 50%).
Table 2. Unit cell parameters a, c, V and agreement indices for hydroxylapatite HAp and Eu-doped HAp (with concentration of Eu3+ of 0.5, 1, 1.5, 2, 10 and 50%).
Samplea [Å]c [Å]V [Å3]RexpRpRwpχ2
HAp9.434979 ± 0.0014016.8803397 ± 0.001042530.65883.43233.996945.192772.28891
Eu0.5HAp9.440372 ± 0.0020676.878142 ± 0.001529530.86013.966014.823496.5542.7309
Eu1HAp9.445147 ± 0.0024446.879144 ± 0.001839531.47453.928944.956026.613782.83367
Eu1.5HAp9.439496 ± 0.0018746.879717 ± 0.001413530.8833.79954.386385.760832.29889
Eu2HAp9.440372 ± 0.0020676.878142 ± 0.001529530.86013.966014.823496.5542.7309
Eu10HAp9.443276 ± 0.0022276.879882 ± 0.001665531.3213.884574.436625.826122.24943
Eu50HAp9.544463 ± 0.0219186.315845 ± 0.023992498.27043.023853.116884.087831.82753

Share and Cite

MDPI and ACS Style

Andronescu, E.; Predoi, D.; Neacsu, I.A.; Paduraru, A.V.; Musuc, A.M.; Trusca, R.; Oprea, O.; Tanasa, E.; Vasile, O.R.; Nicoara, A.I.; et al. Photoluminescent Hydroxylapatite: Eu3+ Doping Effect on Biological Behaviour. Nanomaterials 2019, 9, 1187. https://doi.org/10.3390/nano9091187

AMA Style

Andronescu E, Predoi D, Neacsu IA, Paduraru AV, Musuc AM, Trusca R, Oprea O, Tanasa E, Vasile OR, Nicoara AI, et al. Photoluminescent Hydroxylapatite: Eu3+ Doping Effect on Biological Behaviour. Nanomaterials. 2019; 9(9):1187. https://doi.org/10.3390/nano9091187

Chicago/Turabian Style

Andronescu, Ecaterina, Daniela Predoi, Ionela Andreea Neacsu, Andrei Viorel Paduraru, Adina Magdalena Musuc, Roxana Trusca, Ovidiu Oprea, Eugenia Tanasa, Otilia Ruxandra Vasile, Adrian Ionut Nicoara, and et al. 2019. "Photoluminescent Hydroxylapatite: Eu3+ Doping Effect on Biological Behaviour" Nanomaterials 9, no. 9: 1187. https://doi.org/10.3390/nano9091187

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