Next Article in Journal
Surface Properties of 1DTiO2 Microrods Modified with Copper (Cu) and Nanocavities
Previous Article in Journal
New Insights into the Cell Death Signaling Pathways Triggered by Long-Term Exposure to Silicon-Based Quantum Dots in Human Lung Fibroblasts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Crystallization, Luminescence and Cytocompatibility of Hexagonal Calcium Doped Terbium Phosphate Hydrate Nanoparticles

by
Jaime Gómez-Morales
1,*,
Raquel Fernández-Penas
1,
Ismael Romero-Castillo
1,
Cristóbal Verdugo-Escamilla
1,
Duane Choquesillo-Lazarte
1,
Annarita D’Urso
2,
Maria Prat
2,3,4,5,* and
Jorge Fernando Fernández-Sánchez
6
1
Laboratorio de Estudios Cristalográficos, IACT, CSIC-UGR, Avda. Las Palmeras, nº 4, 18100 Granada, Spain
2
Dipartimento di Scienze della Salute, Università del Piemonte Orientale, Via Solaroli, 17, 28100 Novara, Italy
3
Centro di Biotecnologie per la Ricerca Medica Applicata (BRMA), Via Solaroli 17, 28100 Novara, Italy
4
Consorzio Interuniversitario per Biotecnologie (CIB), Località Padriciano 99, 34149 Area di Ricerca, Italy
5
Consorzio Interuniversitario Nazionale per la Scienza e Tecnologia dei Materiali (INSTM), 28100 Novara, Italy
6
Department of Analytical Chemistry, Faculty of Sciences, University of Granada, Avda. Fuentenueva s/n, 18071 Granada, Spain
*
Authors to whom correspondence should be addressed.
Nanomaterials 2021, 11(2), 322; https://doi.org/10.3390/nano11020322
Submission received: 24 December 2020 / Revised: 19 January 2021 / Accepted: 21 January 2021 / Published: 27 January 2021

Abstract

:
Luminescent lanthanide-containing biocompatible nanosystems represent promising candidates as nanoplatforms for bioimaging applications. Herein, citrate-functionalized calcium-doped terbium phosphate hydrate nanophosphors of the rhabdophane type were prepared at different synthesis times and different Ca2+/Tb3+ ratios by a bioinspired crystallization method consisting of thermal decomplexing of Ca2+/Tb3+/citrate/phosphate/carbonate solutions. Nanoparticles were characterized by XRD, TEM, SEM, HR-TEM, FTIR, Raman, Thermogravimetry, inductively coupled plasma spectroscopy, thermoanalysis, dynamic light scattering, electrophoretic mobility, and fluorescence spectroscopy. They displayed ill-defined isometric morphologies with sizes ≤50 nm, hydration number n ~ 0.9, tailored Ca2+ content (0.42–8.11 wt%), and long luminescent lifetimes (800–2600 µs). Their relative luminescence intensities in solid state are neither affected by Ca2+, citrate content, nor by maturation time for Ca2+ doping concentration in solution below 0.07 M Ca2+. Only at this doping concentration does the maturation time strongly affect this property, decreasing it. In aqueous suspensions, neither pH nor ionic strength nor temperature affect their luminescence properties. All the nanoparticles displayed high cytocompatibility on two human carcinoma cell lines and cell viability correlated positively with the amount of doping Ca2+. Thus, these nanocrystals represent promising new luminescent nanoprobes for potential biomedical applications and, if coupled with targeting and therapeutic moieties, they could be effective tools for theranostics.

Graphical Abstract

1. Introduction

Luminescent nanoparticles are excellent optical probes for uses in biological imaging since they provide the essential fluorescent contrast to analyze and study cells and tissues [1]. Different luminescent labeling agents, mainly organic dyes, and nanomaterials including semiconductor quantum dots, nanodiamonds, gold nanoparticles [2,3,4], several nanostructures labeled with organic dyes [5], and more recently calcium phosphate apatite nanoparticles (nAp) both labeled with organic dyes [6] or doped with luminescent lanthanides (Ln3+) ions [7,8], have been proposed for these applications. The main limitation of organic dyes is photobleaching [5], while the use of quantum dots is controversial due to their cytotoxicity [9]. In the last years lanthanide orthophosphates (LnPO4, Ln3+ = Eu3+, Tb3+, Ce3+, Y3+…) started to attract much attention because they combine the very low solubility and high thermal conductivity of the metal phosphates as host matrices with the favorable features of the luminescent lanthanide ions, such as sharp emission bands, Stokes shift of hundreds of nanometers, and luminescence lifetimes of the order of milliseconds. For all the above reasons they find applications as phosphors, catalysts, sensors and heat-resistant materials [10]. In addition, they exhibit low cytotoxicity, high photostability, and resistance to photobleaching, thus being good optical nanoprobes for bioimaging [11,12,13]. LnPO4·nH2O (n = 0–2) presents five polymorphic modifications, among which the low temperature hydrated phase with hexagonal structure is known as raphdophane [14,15]. In respect to Ln3+-doped nAp, the LnPO4·nH2O system presents the main advantage that at low temperatures (<100 °C) raphdophane is usually the only precipitated phase from aqueous solutions [15]. In calcium phosphate solutions, in contrast, several polymorphs can eventually appear including amorphous calcium phosphate (ACP), dicalcium phosphate dihydrate (DCPD, brushite), octacalcium phosphate (OCP), and Ap [16,17,18]. The synthetic methods to prepare LnPO4 nanocrystals include hydrothermal crystallization [19,20,21], microwave heating [22], ultrasounds [23], oil bath [24], sol-gel combined with electrospinning [25], slow crystallization using surfactants [26], or the layer-by layer adsorption followed by the reaction and crystallization at room temperature [27].
Among lanthanides, Eu3+ and Tb3+ started to be used in various host structures using different excitation wavelength in the UV region, with main emission bands in the Visible spectral region. Concerning LnPO4, the TbPO4·nH2O raphdophane nanorods were excited in the UV region at either 270, 312, 350 or 370 nm, and in the visible region at 482 nm, resulting in similar emission spectra, and yielding a characteristic green emission, due to 5D47F5 transition at 546 nm emission of Tb3+ [28]. Doping with Eu3+ allows the preparation of Eu3+: TbPO4 of hexagonal form with improved energy transfer efficiency between Tb3+ and Eu3+ and non-cytotoxicity on HeLa cells [13]. Recently, our group used citrate (cit) to prepare new luminescent cit-coated Ca2+:EuPO4·nH2O (with n~1) by the method of thermal decomplexing of M(2+/3+)/Ca2+/cit/phosphate/carbonate solutions (M(2+/3+), metallic ions) [29]. The method was initially developed to precipitate monodisperse biomimetic apatites [30] and carbonated apatite nanocrystals, both undoped and doped with transition metals [31,32], and Eu3+ [8]. In this method citrate ions played an active role in the nucleation and growth steps, and remained adsorbed on the nanoparticles surface [33], a feature presented in bone apatite nanocrystals [34]. The new luminescent cit-Ca2+:EuPO4·H2O nanomaterials proved to be fully cytocompatible against GTL-16 human carcinoma cells and showed an improved cytocompatibility as the Ca2+ content increased when contacted with the more sensitive m17.ASC murine mesenchymal stem cells, thus proving to be suitable for bioimaging and cell labeling.
The aim of the present work is to employ the same bioinspired route to prepare cit- Ca2+:TbPO4·nH2O nanocrystals with tuned luminescence and high cytocompatibility, in the assumption that the presence of Ca2+ in the structure and citrate adsorbed on these luminescent solids must be beneficial from the point of view of the cytocompatibility required in bioimaging.

2. Materials and Methods

2.1. Materials and Precipitation Method

For the experiments the following reagents were purchased from Sigma-Aldrich (St Louis, MO, USA): Terbium (III) chloride anhydrous (TbCl3, 99.9% pure, trace metals), sodium citrate tribasic dihydrate (Na3(cit)·2H2O, where cit = citrate = C6H5O7, ACS reagent, ≥99.0% pure), calcium chloride dihydrate (CaCl2·2H2O, Bioxtra, ≥99.0% pure), and sodium phosphate dibasic (Na2HPO4, ACS reagent, ≥99.0% pure). Sodium carbonate monohydrate (Na2CO3·H2O, ACS reagent, 99.5% pure) and hydrochloric acid (HCl, ACS reagent, 37 wt % in H2O) were provided by Merck (Darmstadt, Germany) and Panreac (Barcelona, Spain), respectively. Solutions were prepared with ultrapure deionized water (0.22 µS, 25 °C, Milli-Q, Millipore, Burlington, MA, USA).
The synthesis of nanocrystals was carried out by the already established method of thermal decomplexing of M(2+/3+)/Ca2+/cit/phosphate/carbonate solutions [29], using Tb3+ as lanthanide ion (y = 0.09, 0.07, 0.05 and 0.03 M) and increasing concentrations of doping Ca2+ in the solution (x = 0.01 M, 0.03, 0.05, and 0.07 M) to set x + y = 0.1 M. The experiments lasted 4 h, 24 h, and 96 h. Some of them were matured for 7 days. The precipitates were subjected to washing by centrifugation with ultrapure water (6 cycles, 9000 rpm, 9 min each) and freeze-dried overnight at −50 °C under vacuum (3 mbar).

2.2. Physico-Chemical Characterization of Solid Nanoparticles

As-prepared powders were analyzed by different techniques, as it is described below:
X-ray powder diffraction (XRD) patterns were recorded using a diffractometer PANAlytical MPD (PANAlytical, Almelo, The Netherlands) with a Bragg-Brentano parafocusing geometry and Cu Kα radiation. Data processing of most matured samples was carried out with software TOPAS 6.0 (Coelho Software, Brisbane, Australia) [35]. The contribution of the isotropic peak broadening due to domain size was modeled using the ‘‘CS_L” TOPAS macro based on the Scherrer approximation, and considering the instrumental contribution from a measurement of LaB6 standard (NIST SRM 660c).
TEM observations were performed with a TEM Libra 120 Plus (EELS) instrument at 80 kV (Carl Zeiss, Jena, Germany). Prior to observation samples were dispersed in absolute ethanol (≥99.8% v/v) and deposited on copper microgrids coated with FORMVAR carbon film. HRTEM analysis were done with a TITAN G2 60-300 FEI Instrument (FEI, Hillsboro, OR, USA) operating at 300 kV, equipped with EDX Super X detector to perform microanalysis, and STEM type HAADF.
Fourier transform infrared spectra (FTIR) were recorded in transmittance mode within the wavelength range from 4000 cm−1 to 400 cm−1 using a Perkin-Elmer Spectrum One FTIR spectrometer (Perkin Elmer, Shelton, WA, USA). Pellets with ~1 wt% sample in anhydrous KBr were prepared and pressed with a hydraulic pump at 10 tons. Pure KBr pellets were used to record the background. Raman spectra were recorded with a LabRAM HR spectrometer (Jobin–Yvon, Horiba, Tokyo, Japan). The excitation line was provided by a diode laser emitting at a wavelength of 532 nm while a Peltier cooled charge–couple device (CCD) (1064 × 256 pixels) was used as a detector.
Crystal size distribution (CSD) and electrophoretic mobility (ζ-potential) were analyzed with a Zetasizer Nano ZS analyzer (Malvern Instruments Ltd., Malvern, UK) in aqueous suspensions (~0.5 mg/mL, 25 °C) contained in disposable polystyrene cuvettes. For ζ-potential versus pH measurements the MPT-2 autotitrator (Malvern Instruments Ltd., Malvern, UK) was employed to adjust the pH of the suspensions. Diluted HCl and NaOH solutions (0.25 and 0.1 M, respectively) were used as titration agents without any additional electrolyte.
Elemental analysis of Tb was carried out by inductively coupled plasma mass spectroscopy (ICP MS) using a Perkin Elmer NexION 300D ICP Mass spectrometer (Perkin Elmer). Ca and P were analyzed with Perkin Elmer ICP-OES OPTIMA 8300 spectrometer (Perkin Elmer); C and H were determined by thermoanalysis using Thermo Scientific™ FLASH 2000 CHNS/O Analyzer of Thermo Fisher Scientific (Waltham, MA, USA). TGA analyses were performed with a thermogravimetric analyzer TGA-50H model (Shimadzu, Tokyo, Japan). Samples were weighted in a platinum crucible and heated from room temperature (~28 °C) to 950 °C under 50 mL/min air flow with a heating rate of 20 °C/min.

2.3. Luminescence Spectroscopy

Excitation and emission spectra of both powder and aqueous suspension mg/mL samples, the latter being at ~0.5 mg/mL, were recorded using a Cary Eclipse Varian Fluorescence Spectrophotometer (Varian Australia, Mulgrave, Australia) using λexc = 375 nm, λem = 545 nm, 0.120 µs delay time (td) and 5 ms gate time (tg); photomultiplier voltage of 470 V and slit widthexc/em 5/5 nm for powder samples and photomultiplier voltage of 800 V and slit widthexc/em 10/10 nm for aqueous suspension particles. The excitation and emission spectra were recorded within the range 250–500 nm and 500–750 nm, respectively. Lifetime (τ) measurements were also recorded by a Cary Eclipse Varian Fluorescence Spectrophotometer (Varian Australia, Mulgrave, Australia) using λexc/em = 375/475 nm, 100 µs delay time (td), 0.010 ms gate time (tg), photomultiplier voltage of 600 V, slit widthexc/em 10/10 nm and 100 cycles.
For solid sample analysis, a Cary Eclipse Solid Sample Holder (https://www.agilent.com/en/product/molecular-spectroscopy/fluorescence-spectroscopy/fluorescence-accessories/cary-eclipse-solid-sample-holder) was used to allow the measuring at the optimum angle and to minimize the effect of the thickness.
Concerning the dispersity of the measurements, they are not shown in the figures due to their lower value. In all the case the RSD (relative standard deviation) were lower than 2% (n = 3).

2.4. Cytocompatibility Tests

Nanocrystal cytocompatibility was tested on two cell lines: GTL-16 (a human gastric carcinoma cell line) and A549 (a human lung adenocarcinoma cell line). Then, 24 h after seeding cells (12,000 GTL-16 and 5000 A549 cells/well in 96-well plates), different concentrations of the differentially doped (x = 0.01, 0.03, 0.05, 0.07 M Ca2+) nanoparticles, ranging from 0.1 to 100 mg/mL, were added in 100 mL of fresh medium. Hydrogen peroxide (1 mM) was used as control of toxicity. After 72 h incubation, cell viability was evaluated by the 3-(4,5-Dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide) (MTT, Sigma) colorimetric assay. Briefly, 20 mL of MTT solution (5 mg/mL in a PBS solution) were added to each well and the plate was incubated at 37 °C for 3 h. After the removal of the solution, 125 mL of isopropanol, 0.2 M HCl was added to dissolve formazan crystals. One hundred mL were then removed carefully, and the optical density was measured in a multiwell reader (2030 Multilabel Reader Victor TM X4, Perkin Elmer) at 570 nm. Viability of parallel cultures of untreated cells was taken as 100% viability, and values obtained from cells undergoing the different treatments were referred to this value. Experiments were performed 3 times using 3 replicates for each sample. Data were statistically analyzed and are expressed as mean ± standard deviation. Statistical analyses were performed using a one-way ANOVA with Bonferroni’s post-hoc test for grouped analyses using GraphPad Prism version 4.03 for Windows, GraphPad Software (GraphPad Prism, San Diego, CA, USA). Differences at p < 0.05 were considered to be statistically significant.

3. Results

3.1. Structural, Physicochemical, Morphological Characteristics, and Colloidal Stability of Solid Nanoparticles

The XRD patterns of the solids precipitated for increasing concentration of Ca2+ (x, from 0.01 M to 0.07 M) and increasing maturation times (from 4 to 96 h) are reported in Figure 1. In particular, the patterns showing the distinguishing reflections of the hexagonal (raphdophane) phase identified as TbPO4·nH2O, space group P3121 (PDF 20 1244) are displayed in panels a–d. At x = 0.01 M Ca2+ this is the only crystalline phase identified. The main reflections of this phase are located at 2θ 14.85° (100), 20.43° (101), 25.91° (110), 30.02° (200), 32.10° (102), 38.72° (112), 42.70° (003), 48.10° (301), 49.63° (212), 53.2° (203), and 54.5° (302). Other minor reflections (non-indexed) appear beyond 2θ 55°. These latter peaks are not shown in the PDF 20-1244 card, but are similar to those appearing in the PDF 20-1044 card corresponding to the raphdophane structure in the EuPO4·H2O. In addition, small differences in 2θ positions are found between the two patterns because some peak assignments in PDF file 20 1244 are affected with errors ≥ 0.2°. When x ≥ 0.03 M Ca2+ (Figure 1b–d), however, an additional phase started to crystallize. The reflections of this phase at ~27° and 32.8° are assigned to CaCO3 vaterite (PDF 33-0268), possibly doped with terbium. The presence of vaterite is more evident at x = 0.05 and 0.07 M Ca2+. Thus, this compound becomes stable in calcium-rich medium due to the presence of residual CO32− in the crystallizing solutions. Nevertheless, its presence does not represent a problem concerning biomedical uses since this polymorph of CaCO3 is biocompatible [36]. At 0.07 M Ca2+ we found a slightly different phase evolution with time. At 4 h, the bulging of the baseline is characteristic of an amorphous phase, probably of terbium phosphate, whereas at 24 h the reflections of the raphdophane phase started to emerge from the bulging baseline. In none of the experiments have we detected the presence of any calcium phosphate phase.
The microstructural study of 96 h samples was performed by analysis of the full XRD pattern, excluding the reflections of vaterite (Figure 2). Results show isometric crystalline domains whose average Scherrer diameters were 37.6 ± 0.2 nm for the less Ca2+-doped sample (Figure 2a), 28.6 ± 0.2 nm for x = 0.03 M Ca2+ (Figure 2b), 24.9 ± 0.2 for x = 0.05 (Figure 2c), and 30.2 ± 0.1 for x = 0.07 M Ca2+ (Figure 2d), respectively.
TEM observations of these precipitates show nanocrystals with ill-defined shapes and aspect ratios close to 1 (Figure 3a–c). In contrast to what has been reported for cit-Ca2+:EuPO4·H2O nanoparticles, no appreciable dependence of their morphology with Ca2+ doping was found [29].
The sizes deduced on the basis of either TEM (29 ± 5, 28 ± 3, 33 ± 6, Figure 3a–c) or SEM (x = 0.07 M Ca2+, L = 55 ± 20 nm, not shown) images are equivalent to those measured from XRD data. The indexed SAED patterns (insets) show the crystallographic planes of the hexagonal phase, i.e., (100), (110) in Figure 3a, or (100), (102), (200) in Figure 3c. The sample obtained for x = 0.01 Ca2+ was studied in more details, as shown in Figure 3d–i. Figure 3d displays the image in high angle annular dark field (HAADF) in Scanning Transmission Electron Microscopy (STEM). The elemental mappings composition of the nanoparticles (Figure 3e,f,h,i) show a homogeneous distribution of Tb, Ca, P, and O. HRTEM images of this sample (Figure 3g) reveal lattice fringes, basically monodomains, whose d-spacing of 2.78 Å corresponds to plane (102), also shown in the fast Fourier transform FFT images of this sample (Figure 3d, inset).
FTIR and Raman spectroscopy were employed to reveal the spectroscopic features of the Ca2+-doped samples (see Figure 4 for the stacking of spectra of precipitates prepared at different Ca2+ doping concentrations, and Supplementary Figures S1 and S2 for the stackings as a function of maturation time).
FTIR spectra of Ca2+-doped samples (Figure 4a) show basically the same spectroscopic features upon calcium doping, especially for x = 0.01, 0.03, and 0.05 M Ca2+. These are a broadband between 3600–3700 cm−1 and 2600 cm−1 corresponding to the O–H stretching of associated water, a main broad band at 1000–1100 cm−1, which corresponds to the antisymmetric stretching mode of PO43− groups (ν3PO4) and two intense bands at ~620 and 535 cm−1 ascribed to the bending mode of the same PO43− groups (ν4PO4). The broader signals around 1590–1620 cm−1 and 1405–1430 cm−1 are assignable to the antisymmetric and symmetric stretching frequencies of carboxylate (−COO) groups of the adsorbed citrate [29]. No appreciable differences in peak assignments in the FTIR profiles were found after 4 h of maturation at any of the Ca2+ doping concentrations used (Supplementary Figure S1a–d).
In parallel, Raman characterization (Figure 4b) show the most intense band at 996 cm−11), which can be assigned to the P–O symmetric stretching mode of PO4 group in raphdophane-type orthophosphates, while that at ~1100 cm−1 is attributed to the antisymmetric stretching (ν3PO4). The intensity of both bands decreases upon Ca2+ doping. The region comprised between 300 and 700 cm−1 corresponds to the deformation modes of the PO4 tetrahedron. The most intense band in this region is at 485 cm−1 and is assigned to the symmetric vibrations, while those at ~570 and 620 cm−1 can be ascribed to the asymmetric vibrations (ν4) [38,39]. The presence of vaterite in the spectra for x = 0.05 and 0.07 M Ca2+, with bands at 1080–1090 cm−1 ν1CO3 and 740–750 cm−1 ν4CO3 [40] is hidden by the most intense bands of the phosphate groups. Clear differences were detected when plotting Raman spectra at increasing maturation times (Supplementary Figure S2). First, an increase in the intensity of the antisymmetric stretching mode ν3PO4 at ~1100 cm−1, and particularly in the less Ca2+-doped sample (Supplementary Figure S2a) is observed. Second, the highest Ca2+-doped sample shows the ν1PO4 mode shifted from 970 cm−1 to 990 cm−1 (Supplementary Figure S2d), and the appearance of bands around 745 cm−1 and 1080 cm−1 (related to the vaterite phase). Also, the peaks at 845 cm−1 and 1460 cm−1 in samples at 4 and 24 h, in vibrational zones of citrate groups [41].
Thermogravimetric analyses (Supplementary Figure S3) revealed a behavior very close to the ones reported for lanthanide phosphate materials of raphdophane structure [29,42,43,44,45] but with some peculiarities reflecting the presence of citrates and carbonates in samples prepared at x = 0.05 and 0.07 M Ca2+. The first weight loss, recorded between Tamb and 120–140 °C of about 2 wt%, is attributed to the residual adsorbed water, while the second one between 140 °C and about 600 °C, of 5–5.5 wt%, corresponds to the release of structural water, and this is the main weight loss. The third weight loss above 600 °C is associated with the decomposition of citrate molecules in samples prepared with x = 0.01 and 0.03 M Ca2+, and with decomposition of citrate and of CaCO3 in CO2 and CaO, when x = 0.05 and 0.07 M Ca2+. Specific losses are observed in these two samples. For x = 0.05 M Ca2+ the loss of CO2 is around 0.6 wt% and for x = 0.07 M Ca2+ is of 2.2 wt%.
Quantitative analyses of Ca, P, Tb, C, and H of as-prepared samples are reported in Table 1. This table also shows the percent weights of structural H2O determined by TGA and thus, the hydration number (n) of the raphdophane phase. Percent weights of CaCO3 and citrate were determined by combining TGA data with mass balances. The data reveal increased percentages of dopant Ca from 0.42 wt% to 8.11 wt% in the structure of TbPO4.nH2O as the percentage of Tb decreases, according to the increase of Ca2+/Tb3+ ratio in the precursor solutions.
In these samples, the percentages of adsorbed citrate were around 1.1–1.2 wt% but in the highest Ca2+-doped sample that percentage rose to 4 wt%. This higher amount of citrate is likely due to the higher percentage of CaCO3, which can be also coated with citrate. Table 1 also shows that structural H2O ranges between 5.0 and 6.3 wt% rendering hydration numbers around 0.9.
The analysis of CSD and ζ-potential versus pH of aqueous suspensions of the nanoparticles is also relevant to assess their potential as luminescent probes in nanomedicine applications. The tendency of the colloid to disperse or aggregate in simulated physiological conditions in the blood (pH around 7.4) or in the tumor microenvironment (pH around 5–6) are related to the size and surface charge of the nanoparticles [16,46]. In addition, these features have an impact on the formation of the protein corona around the nanoparticles [47]. In this study, the CSD of cit-Ca2+:TbPO4·nH2O nanoparticles was plotted as cumulative volume-based distribution because they visually show the percentiles of the distribution D10, D50 and D90 (Figure 5 and Supplementary Figure S4), widely employed to characterize crystal populations in pharmaceutical industry. These percentiles characterize the percentage of cumulative volume undersize distribution (percentage of the population smaller than the indicated size). Thus, D10 is a percentile closer to the size of the individual particles while D50, the median of the population, is here somewhat influenced by particle aggregation. D90 is entirely influenced by the aggregation of nanocrystals. The percentiles for samples prepared with x = 0.01, 0.03, 0.05, and 0.07 M Ca2+ for 96 h were D10 21, 51, 57, and 45 nm, respectively, and D50 37, 531, 122, and 94 nm, respectively (Figure 5b).
In these samples, the evolution of cumulative CSD versus time (at 4 h, 24 h and 96 h, Figure S4a–d) does not follow a general trend, thus reflecting a difference in Ca composition as well as the presence of a secondary phase (vaterite) at high doping concentrations and its influence in the aggregation.
These differences are also reflected in the profile of the ζ -potential versus pH curves (Figure 5c–f). While samples obtained with x = 0.01 and 0.03 M Ca2+ displayed similar curves, with ζ -potential values −17.3 and −17.4 mV at pH 7, and −14.0 and −10.5 mV at pH 5 (Figure 5c,d), the curves of sample obtained with x = 0.05 M Ca2+ (Figure 5e) and x = 0.07 M Ca2+ (Figure 5f) show a different profile. The minimum ζ -potential values were found at pH 6.0 (−11.5 mV) and pH 8.0 (−15.9 mV) respectively. ζ -potentials were 0 beyond pH 6.0 in the first case, and below pH 8.0 in the second one.

3.2. Luminescence Properties of Cit-Ca2+:TbPO4·nH2O Nanoparticles

3.2.1. Luminescence in Solid-State

It is well-known that some lanthanides, especially europium (III) and terbium (III), form highly fluorescent chelates with many different organic ligands. The sensitized fluorescence results from the ligand absorbing light, the energy of which is then transferred to the chelated metal ion. In fact, Tb(III) emits the energy as narrow-banded, line-type fluorescence with a long Stokes shift (over 250 nm) and an exceptionally long fluorescence decay time (up to 1 ms) [48]. Because of the long fluorescence decay time (over 10 times longer than the average background fluorescence) of Tb(III), a delay time (td) and a gate time (tg) can be used during the measuring, remarkably reducing the background fluorescence.
The luminescence properties of solid cit-Ca2+:TbPO4·nH2O samples are shown in Supplementary Materials (Figures S5 and S6). They are the same as those depicted in Figure 6 which correspond to x = 0.01 M Ca2+; neither the maturation time nor the Ca2+ doping concentrations affect the excitation and emission wavelengths. This finding is expected because the electronic transitions of f orbitals are not affected by crystal’s field.
The observed excitation wavelengths for the powder were 230, 284, 300, 320, 340, 350, 368 and 375 nm. The broad bands between 200 and 300 nm, centered at 230 nm, correspond to charge transfer (called charge transfer band, CTB), which occurs by electron delocalization from the filled 2p shell of O2− to the partially filled 4f shell of Tb3+. Also, this band can partially be attributed to the charge transfer transition X5+–O2− [49,50]. The rest of the less intensive excitation wavelengths correspond to the 7F65I8 and 5F4,55H4, 7F65H5,6, 7F65H7, 7F65L7,8 and 7F65L7,8, 7F65L9,5D2,5G5, 7F65L10, and 7F65G6, 5D3 transitions [51].
SM (see Supplementary Figure S5) shows the emission spectra of the cit-Ca2+:TbPO4·nH2O with x = 0.01 M and 96 h of maturation time using 230 and 375 nm excitation wavelengths. It is possible to conclude that the emission spectra are the same; only the emission intensity is affected. In order to increase the biological applicability of the system, 375 nm (which is closer to the visible range) was selected as excitation wavelength.
Concerning the emission wavelengths, they are centered at 490, 545, 586 and 620 nm which correspond to the Tb3+ 5D47F6, 5D47F5, 5D47F4 and 5D47F3 transitions, respectively [52]. The emission wavelength corresponding to the hypersensitive transition without inversion centre (5D47F5, 545 nm for Tb3+) produces the highest relative luminescence intensity (R.L.I.). Therefore, the optimum excitation and emission wavelengths of solid cit-Ca2+:TbPO4·nH2O samples were 375 nm and 545 nm, respectively.
Supplementary Figures S7 and S8 show the effect of maturation time (t) at different doping Ca2+ concentrations (x) on the relative luminescence intensity (R.L.I.).
The R.L.I. of samples prepared with x = 0.01, 0.03 and 0.05 M Ca2+ is not affected by the maturation time (Supplementary Figure S7). It can be due to the slowing down of the crystal growth caused by the adsorption of citrate up to 2 h maturation [29]. In contrast, for x = 0.07 M Ca2+ doping concentration, the R.L.I is highly affected; at shorter times (4 and 24 h) the signal is high, but it decreases drastically in the sample prepared at 96 h. As it was previously commented for this concentration, at shorter times (4 and 24 h) the material is amorphous without an ordered structure and with a high amount of Tb3+ adsorbed on its surface, showing high luminescence. However, at 96 h it is crystalline, exhibiting an improved internal ordering in which Tb3+ is buried in the structure of the nanoparticles, resulting in a decrease on the luminescence. It was previously reported that amorphous materials provide much more luminescence emission than crystalline ones [29]. On the other hand, this sample contains a higher proportion of citrate (~4 wt%) providing a less emissive material than at other Ca2+-doped concentrations.
Analyzing the variation of the R.L.I at a fixed maturation time versus x it is also possible to observe that the main differences are found for x = 0.07 M (Supplementary Figure S8). It can be explained for the same reasons commented before that, at shorter maturation times, the sample has an amorphous nature and at the higher one it is crystalline, and the amount of citrate is higher, providing a less luminescent material.
Concerning the luminescence lifetime (τ), Supplementary Figures S9–S11 show the luminescence decay curves and the variation of the luminescence lifetime versus both maturation time and Ca2+ doped concentrations, respectively. For each case, the decay profile was analyzed as a single exponential component (R.L.I. = A·e^(−t⁄τ) + C). It is possible to conclude that the maturation time at a given Ca2+ doped concentration does not affect the lifetime (Supplementary Figure S10). However, the luminesce lifetime for a given maturation time is increased by increasing the x. It might be due to the presence of progressive amounts of vaterite.

3.2.2. Luminescence of the Nanoparticles in Aqueous Suspension

The luminescence properties of cit-Ca2+:TbPO4·nH2O dispersed in aqueous media are similar to those depicted in Figure 7, which corresponds to x = 0.01 M Ca2+, 96 h maturation time, pH 7.4, and 25 °C. It is also possible to observe that the excitation and emission wavelength of the material dispersed in water is practically the same than those in powder as well as the luminescence lifetime is not affected by dispersing the sample in an aqueous media. On the other hand, it is possible to conclude that the material has the same luminescence properties (R.L.I. and lifetime) at any physiological pH, so variation of the pH in real biological samples does not affect the luminescence properties of the particles.
The effect of ionic strength was evaluated by suspending the particles in 0, 25, 50, 75 and 100 mM NaCl solutions (Figure S12). It is possible to deduce that neither the R.L.I. nor the luminescence lifetime is significantly affected by the ionic strength, which is also important for the final applications of these nanoparticles.
The temperature may also theoretically affect the luminescence by quenching of the excited states, i.e., for increasing T the molecular motion and collisions increase, and hence the luminescence emission decreases by increasing encounters probabilities. [53] Figure S13 shows the experimental results. To sum up, increasing from 25 to 40 °C does not affect considerably the luminescence emission of all the tested materials in suspension. The change in fluorescence intensity is normally 1% per degree Celsius [53] and the decreases for these materials were 0.2, 0.4, 2.6, and 0.6% per degree Celsius, for 0.01 M, 0.03 M, 0.05 M and 0.07 M Ca2+ doping concentration, respectively. This is very important in view of medical and biomedical imaging (~37.4 °C) while the rest of the experiments were performed at room temperature (25 °C).

3.3. Cytocompatibility of Cit-Ca2+:TbPO4·nH2O Nanoparticles

The cytocompatibility of cit-TbPO4·nH2O samples doped with different concentrations of Ca2+ was tested in a MTT assay on the GTL-16 human carcinoma cells and the A549 human lung adenocarcinoma cells, after incubation at concentrations ranging from 0.1 to 100 µg/mL. GTL-16 cells were chosen because this is the reference standard cell line used in the lab for this type of experiments. In particular, the fact that these tumor cells express a high level of receptors for a growth factor has made this cell line a good model for studying tumor cell ligand-mediated targeting by nanoparticles, possibly loaded by a drug [6,16]. A549 cells were chosen since they are easily available cells in many laboratories.
No toxicity was observed on both cell lines at any nanoparticle concentration (Figure 8), since in all cases a cell viability higher than 85% was observed, largely above the cut-off of 70% indicated by ISO 10993–5:2009 [54]. The presence of the doping Ca2+ appears to increase the biocompatibility of the nanocrystals in a dose-dependent manner. On the other side, both cell types were sensitive to the toxic activity of hydrogen peroxide, which reduced their viability to less than 50%. This in vitro assay thus shows the good cytocompatibility of the cit-Ca2+:TbPO4·nH2O nanoparticles.

4. Discussion

The above results confirm the herein reported bioinspired crystallization route succeeded in preparing cit-Ca2+:TbPO4·nH2O nanocrystals of the hexagonal rhabdophane phase, with n~0.9 and tailored Ca2+content between 0.42 wt% and 8.11 wt%. The nanocrystals displayed isometric ill-defined morphologies, and adsorbed citrate on their surfaces, a chemical feature exhibited by bone nAps [34]. As Ca2+ increases the risk of appearance of a secondary phase of CaCO3 (vaterite) must be considered. However, this trouble can be properly assumed since vaterite is biocompatible. In addition, the percent weight of citrate increases, particularly in the sample prepared with x = 0.07 M. As citrate molecules were not removed after different cycles of washing, they should be adsorbed not only to Ca2+:TbPO4·H2O but also to the CaCO3 surface. As hypothesized, these features have an impact on the cytocompatibility of the cit-Ca2+:TbPO4·H2O. In fact, all samples displayed a high cytocompatibility when tested on two human carcinoma cell lines and cell viability appears to positively correlate with the amount of doping Ca2+ in the nanocrystals, in line with what was reported for cit-Ca2+:EuPO4·nH2O nanocrystals [29].
Also important are the average sizes of the nanocrystals. The average sizes measured in TEM or SEM micrographs were equivalent to those of their crystal domains, indicating they are basically monodomains, with no appreciable dependence of their morphology (isometric) with Ca2+ doping. This finding contrasts with what was reported for cit-Ca2+:EuPO4·H2O nanoparticles [29]. Those nanoparticles showed elongated (anisometric) morphologies and their aspect ratio decreased with Ca2+ doping. It is worth highlighting that percentiles D10 of the CSD are close to these average sizes deduced from electron microscopy images. This finding indicates that aggregation is very low and D10 could define the nanoparticle size in the context of the whole population, particularly in the less Ca2+ doped sample. D50, instead, doubles the D10 percentile or is even 10 times higher in sample x = 0.03M Ca2+, suggesting a higher contribution of particle aggregation below the median size.
Concerning the ζ-potential versus pH of nanoparticle suspensions, those samples prepared at x = 0.01 and 0.03 M Ca2+ show a high and negative ζ-potential above pH 4, therefore they display great stability both at pH~7.4 simulating physiological pH, and at pHs~5–6, simulating the environment of cancer cells. The CSD of the sample prepared at x = 0.01 M Ca2+ correlates well with the ζ-potential versus pH measurements of this sample, and reflects the role of surface charge on the stability of the nanoparticles and their aggregation behavior. However, there is a lack of correlation between the CSD and the ζ-potential values for the 0.03 M Ca2+sample, indicating that other effects besides the nanoparticles surface charge can be involved in particle aggregation. Also, the other samples present a negative ζ-potential, but these are less negative, indicating poorer stability of the suspensions at the pHs of interest for biological applications. In the latter samples, the curves reflect the presence of vaterite, and also the increasing amount of citrate, which correlate with a relatively higher cytocompatibility. Thus, when x = 0.05 M Ca2+ the ζ -potential value close to zero indicates a high amount of surface Ca2+ that compensate the negative charge due to citrate at basic pH. When x = 0.07 M Ca2+ (Figure 5f) the ζ-potential values at basic pHs are again negative. This finding is due to the higher percentage of citrate negatively charged respect to positively charged surface Ca2+ coating both Ca2+:TbPO4·nH2O and CaCO3 particles. It was reported that the mode of adsorption of citrate on a substrate depends on the pH of the solution. Citrate molecules absorb on monodentate or bidentate configuration (i.e., on nanoapatites [55]) displaying the third carboxylate group upward toward the solution. This arrangement gives rise to a negative net surface charge that is reflected in a negative ζ -potential. While preventing aggregation, the negative ζ -potentials of nanocrystal could interfere with their interaction with the cell surface, which is also negatively charged, and their eventual uptake by cell endocytosis. This behavior, however, is not the only possibility, since some negatively charged nanoparticles, such as Fluorescein-5-isothiocyanate-labelled hydroxypatite nanoparticles [6] and magnetic nanoparticles [56], were reported to be internalized by cells.
To discuss how the nanocrystals herein described were formed, one must consider both the influence of their tridimensional crystal structure and the presence of additives in the precursor solution. Indeed, the growth morphology of a nanocrystal results from the interplay between the growth rates of its outermost crystallographic planes and the growth-inhibiting/promoting effect of the additives. The raphdophane structure of the LnPO4·nH2O is described as chains formed by lanthanide with eight-fold coordination alternating with tetrahedral orthophosphate ions. These chains are extended along with the c-axis, each one linked to four neighboring chains, thus forming open channels which are filled with between 0.5 to 2 H2O molecules per formula unit, stabilizing the structure [15,20,57]. Consequently, LnPO4·nH2O show the general trend to grow along [001] direction forming nanowires.
The second influence we must consider is the effect of citrate, which is a known growth inhibitor. Citrate adsorbs on specific outermost planes of incipiently formed nanocrystals slowing down its growth along the [001] direction, and leading to the formation of nanocrystals with lower aspect ratios than those formed in additive-free solutions. As we obtain basically the same crystal morphology irrespective of the amount of Ca2+ doping the cit-Ca+:TbPO4·nH2O structure, we conclude that Ca2+has limited inhibiting/promoting effect on the growth of these nanocrystals, against that of citrate, in contrast to what found in cit-Ca2+:EuPO4·nH2O [29]. On the other side, as already discussed above, the addition of this metal ion had a positive effect on the cytocompatibility of the nanoparticles, in line with the previous report relative to cit-Ca2+:EuPO4·nH2O [29]. Moreover, at the highest doping dose, Ca2+ enhances the adsorption of citrate, an ion contributing to biomimetic properties of apatites [34].
Finally, the luminescence study of these nanoparticles in solid phase indicate that R.L.I. is affected by neither the maturation time nor the Ca2+ content when x is lower than 0.07 M Ca2+, as occurred on cit- Ca2+:EuPO4·nH2O, while the R.L.I. at x = 0.07 M Ca2+ is high at short maturation time (due to the amorphous nature of the samples) and low at high maturation time (due to the crystalline nature of the sample and also due the higher citrate amount coating the raphdophane and vaterite nanoparticles). Concerning the luminescence properties in aqueous dispersion, neither pH nor ionic strength nor temperature affect their luminescence properties at physiological range.
According to all the above considerations the optimal conditions to synthesize cytocompatible cit-Ca2+:TbPO4·nH2O nanocrystals free of vaterite or with negligible amount of this phase, with enough and exploitable luminescent properties (R.L.I and luminescent lifetimes) and higher stability of their aqueous suspensions at the pH range of biological interest (5–7.4), are when using Ca2+ doping concentration x = 0.01 M and x = 0.03 M, at any of the maturation times from 4 to 96 h.
Summing up, all the above properties are positive relative to the applications of these nanoparticles as luminescent labeling agents and highlight the potential of the thermal decomplexing synthetic method to prepare this kind of nanophosphors.

5. Conclusions

Cit-Ca2+:TbPO4·nH2O nanocrystals of the hexagonal rhabdophane phase, with n ~ 0.9 and tailored Ca2+ content between 0.42 and 8.11 wt% were successfully obtained by thermal decomplexing of Ca2+/Tb3+/citrate/phosphate/carbonate solutions. These nanocrystals of about 29–37 nm are mainly monodomains and display basically isometric ill-defined morphologies, with Ca2+ having limited inhibiting/promoting effect on the growth of these nanoparticles, but enhancing the amount of citrate adsorbed on the surface in dose-depending manner. All the nanoparticles displayed high cytocompatibility on two human carcinoma cell lines and cell viability correlated positively with the amount of doping Ca2+.
Luminescence properties of the nanocrystals reveal that luminescent lifetimes increase (between ~800 and ~2600 µs) by increasing the Ca2+doping concentration. Their relative luminescence intensities in solid state (around 200 units) are affected by neither Ca2+, nor citrate content, nor maturation time for Ca2+ doping concentrations below 0.07 M Ca2+. Only at this doping concentration the maturation time strongly affects R.L.I, decreasing it. At low maturation times (4 h, 24 h) it is high (around 650 units). However, at 96 h, it strongly decreases up to ~40. This finding was related to the amorphous nature of the precipitates obtained at 4 and 24 h, whereas at 96 h, the precipitates were crystalline. In addition, this sample also contains vaterite and a higher percent weight of citrate than the samples obtained with lower doping concentrations of Ca2+, and therefore it presents a lower percent weight of emissive material. In aqueous suspensions, neither pH nor ionic strength nor the temperature affect their luminescence properties.
According to all the above considerations, the optimal conditions to synthesize cytocompatible cit-Ca2+:TbPO4·nH2O nanocrystals free of vaterite or with negligible amount of this phase, with enough and exploitable luminescent properties (R.L.I and luminescent lifetimes) and higher stability of their aqueous suspensions at the pH range of biological interest (5–7.4), are when using Ca2+ doping concentration x = 0.01 M and x = 0.03 M, at any of the maturation times from 4 to 96 h.
Summing up, the above properties are encouraging relative to the potential bioimaging applications of the nanocrystals. They can thus represent promising new luminescent nanoprobes for such applications and, if coupled with targeting and therapeutic moieties, they could be effective tools for theranostics.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/11/2/322/s1, Figure S1: Evolution of FTIR spectra with time of samples prepared with different Ca2+ doping concentrations x = 0.01 to 0.07 M, Figure S2: Evolution of Raman spectra with time of samples prepared with different Ca2+ doping concentrations x = 0.01 to 0.07 M, Figure S3: TGA analyses of precipitates obtained at 96 h from solutions with Ca2+ doping concentrations ranging from (a) x = 0.01 to (d) x= 0.07 M, Figure S4: Cumulative volume oversize distribution of the cit-Ca2+:TbPO4.nH2O nanocrystals prepared with Ca2+ doping concentrations x = 0.01, 0.03, 0.05 and 0.07 M at 4, 24 and 96 h, Figure S5: Excitation (dashed lines) and emission (solid lines) uncorrected spectra of solid cit-Ca2+:TbPO4·nH2O samples prepared with x = 0.01 M Ca2+ at maturation times of 96 h using td = 120 µs, td = 5 ms and (a) λexc/em = 230/545 nm, slit width exc/em = 2.5/2.5 nm, detector voltage 545 V; b) λexc/em = 375/545 nm, slit width exc/em = 5/5 nm, detector voltage 470 v, Figure S6: Excitation (dashed lines) and emission (solid lines) uncorrected spectra of solid cit-Ca2+:TbPO4·nH2O samples prepared with different Ca2+ doped concentration at maturation times of 4 h, 24 h, 96 h and 7 days, Figure S7: Variation of the R.L.I. of the solid cit-Ca2+:TbPO4·nH2O samples at the maximum excitation and emission wavelengths at several Ca2+ concentrations when the maturation time is changed. Figure S8: Variation of the R.L.I. of the solid cit-Ca2+:TbPO4·nH2O samples at the maximum excitation and emission wavelengths at several maturation time when the Ca2+ concentration is changed. Figure S9: Luminescence decay curve of different solid cit-Ca2+:TbPO4·nH2O samples at maturation times of 96 h, td = 100 µs, tg = 0.01 ms, λexc/em = 375/545 nm, slit-widthsexc/em = 10/10 nm, and detector voltage = 600 V. Circles correspond to experimental data (100 cycles) and lines to the fitting equation. Figure S10: Variation of the luminescence lifetime of the solid cit-Ca2+:TbPO4·nH2O nanoparticles prepared at several Ca2+ concentrations when the maturation time is changed. Figure S11: Variation of the luminescence lifetime of the solid cit-Ca2+:TbPO4·nH2O samples at several maturation times when the Ca2+ concentration is changed. Figure S12: Effect of the ionic strength over the (a) R.L.I. and (b) luminescence lifetime of the cit-Ca2+:TbPO4·nH2O samples at 96 h maturation time dispersed in aqueous media at several Ca2+ concentrations. Figure S13. Effect of the temperature over the (a) R.L.I. and (b) luminescence lifetime of the cit-Ca2+:TbPO4·nH2O samples at 96h maturation time dispersed in aqueous media at several Ca2+ concentrations.

Author Contributions

Conceptualization, J.G.-M., D.C.-L., and M.P.; Investigation, J.G.-M., J.F.F.-S., M.P., I.R.-C., R.F.-P., A.D., and C.V.-E.; methodology, R.F.-P., A.D., J.F.F.-S., and C.V.-E; Supervision, J.G.-M.; writing—original draft preparation, J.G.-M. and M.P.; writing—review and editing, J.G.-M., M.P., and J.F.F.-S.; funding acquisition, J.G.-M., D.C.-L. and M.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Spanish Agencia Estatal de Investigación of the Ministerio de Ciencia, Innovación y Universidades (MICIU) and co-funded with FEDER, UE, Project No. PGC2018-102047-B-I00 (MCIU/AEI/FEDER, UE). J.F.F.-S. acknowledges the Spanish Ministerio de Economía y Competitividad Project No. CTQ2017-88079-P for support. M.P. acknowledges the Progetto di Ricerca Fondi di Ateneo per la Ricerca—FAR 2017 “Development of innovative biological materials for the functional regeneration of cardiac tissue models”.

Acknowledgments

Authors acknowledge technical support of technician of the CIC of the University of Granada.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sreenivasan, V.K.A.; Zvyagin, A.V.; Goldys, E.M. Luminescent nanoparticles and their applications in the life sciences. J. Phys. Condens. Matter 2013, 25, 194101. [Google Scholar] [CrossRef] [PubMed]
  2. Li, J.; Zhu, J.-J. Quantum dots for fluorescent biosensing and bio-imaging applications. Analyst 2013, 138, 2506. [Google Scholar] [CrossRef] [PubMed]
  3. Fu, C.-C.; Lee, H.-Y.; Chen, K.; Lim, T.-S.; Wu, H.-Y.; Lin, P.-K.; Wei, P.K.; Tsao, P.H.; Chang, H.-C.; Fann, W. Characterization and application of single fluorescent nanodiamonds as cellular biomarkers. Proc. Natl. Acad. Sci. USA 2007, 104, 727–732. [Google Scholar] [CrossRef] [Green Version]
  4. Yeh, Y.-C.; Creran, B.; Rotello, V.M. Gold nanoparticles: Preparation, properties, and applications in bionanotechnology. Nanoscale 2012, 4, 1871–1880. [Google Scholar] [CrossRef]
  5. Wang, F.; Tan, W.B.; Zhang, Y.; Fan, X.P.; Wang, M.Q. Luminescent nanomaterials for biological labelling. Nanotechnology 2006, 17, R1–R13. [Google Scholar] [CrossRef]
  6. Oltolina, F.; Gregoletto, L.; Colangelo, D.; Gómez-Morales, J.; Delgado-López, J.M.; Prat, M. Monoclonal antibody-targeted fluorescein-5-isothiocyanate-labeled biomimetic nanoapatites: A promising fluorescent probe for imaging applications. Langmuir 2015, 31, 1766–1775. [Google Scholar] [CrossRef] [PubMed]
  7. Neacsu, I.A.; Stoica, A.E.; Vasile, B.S.; Andronescu, E. Luminescent hydroxyapatite doped with rare earth elements for biomedical applications. Nanomaterials 2019, 9, 239. [Google Scholar] [CrossRef] [Green Version]
  8. Gómez-Morales, J.; Verdugo-Escamilla, C.; Fernández-Penas, R.; Parra-Milla, C.M.; Drouet, C.; Maube-Bosc, F.; Oltolina, F.; Prat, M.; Fernández-Sánchez, J.F. Luminescent biomimetic citrate-coated europium-doped carbonated apatite nanoparticles for use in bioimaging: Physico-chemistry and cytocompatibility. RSC Adv. 2018, 8, 2385–2397. [Google Scholar] [CrossRef] [Green Version]
  9. Derfus, A.M.; Chan, W.C.W.; Bhatia, S.N. Probing the cytotoxicity of semiconductor quantum dots. Nano Lett. 2004, 4, 11–18. [Google Scholar] [CrossRef] [Green Version]
  10. Fang, Y.-P.; Xu, A.-W.; Song, R.-Q.; Zhang, H.-X.; You, L.-P.; Yu, J.C.; Liu, H.-Q. Systematic synthesis and characterization of single-crystal lanthanide orthophosphate nanowires. JACS 2003, 125, 16025–16034. [Google Scholar] [CrossRef]
  11. Luo, Q.; Shen, S.; Lu, G.; Xiao, X.; Mao, D.; Wang, Y. Synthesis of cubic ordered mesoporous YPO4:Ln3+ and their photoluminescence properties. J. Mater. Chem. 2009, 19, 8079–8085. [Google Scholar] [CrossRef]
  12. Zhang, F.; Wong, S.S. Ambient large-scale template-mediated synthesis of high-aspect ratio single-crystalline, chemically doped rare-earth phosphate nanowires for bioimaging. ACS Nano 2010, 4, 99–112. [Google Scholar] [CrossRef] [PubMed]
  13. Di, W.; Li, J.; Shirahata, N.; Sakka, Y.; Willingere, M.-G.; Pinna, N. Photoluminescence, cytotoxicity and in vitro imaging of hexagonal terbium phosphate nanoparticles doped with europium. Nanoscale 2011, 3, 1263–1269. [Google Scholar] [CrossRef]
  14. Ito, H.; Fujishiro, Y.; Sato, T.; Okuwaki, A. Preparation of lanthanide orthophosphates by homogeneous precipitation under hydrothermal conditions using lanthanide-EDTA chelates. Br. Ceram. Trans. 1995, 94, 146–150. [Google Scholar]
  15. Zollfrank, C.; Scheel, H.; Brungs, S.; Greil, P. Europium(III) orthophosphates: Synthesis, characterization, and optical properties. Cryst. Growth Des. 2008, 8, 766–770. [Google Scholar] [CrossRef]
  16. Iafisco, M.; Delgado-López, J.M.; Varoni, E.M.; Tampieri, A.; Rimondini, L.; Gómez-Morales, J.; Prat, M. Cell surface receptor targeted biomimetic apatite nanocrystals for cancer therapy. Small 2013, 9, 3834–3844. [Google Scholar] [CrossRef]
  17. Iafisco, M.; Marchetti, M.; Gómez-Morales, J.; Hernández-Hernández, M.A.; García-Ruiz, J.M.; Roveri, N. Silica gel template for calcium phosphates crystallization. Cryst. Growth Des. 2009, 9, 4912–4921. [Google Scholar] [CrossRef]
  18. Ramírez-Rodríguez, G.B.; Delgado López, J.M.; Gómez-Morales, J. Evolution of calcium phosphate precipitation in hanging drop vapor diffusion by in situ Raman microspectroscopy. CrystEngComm 2013, 15, 2206–2212. [Google Scholar] [CrossRef]
  19. Bu, W.; Zhang, L.; Hua, Z.; Chen, H.; Shi, J. Synthesis and characterization of uniform spindle-shaped microarchitectures self-assembled from aligned single-crystalline nanowires of lanthanum phosphates. Cryst. Growth Des. 2007, 7, 2305–2309. [Google Scholar] [CrossRef]
  20. Fang, Y.-P.; Xu, A.-W.; Dong, W.-F. Highly improved green photoluminescence from CePO4:Tb/LaPO4 core/shell nanowires. Small 2005, 1, 967–971. [Google Scholar] [CrossRef]
  21. Nuñez, N.O.; Liviano, S.R.; Ocaña, M.J. Citrate mediated synthesis of uniform monazite LnPO4 (Ln = La, Ce) and Ln:LaPO4 (Ln = Eu, Ce, Ce + Tb) spheres and their photoluminescence. Colloid Interface Sci. 2010, 349, 484–491. [Google Scholar] [CrossRef]
  22. Patra, C.R.; Bhattacharya, R.; Patra, S.; Basu, S.; Mukherjee, P.; Mukhopadhyay, D. Inorganic phosphate nanorods are a novel fluorescent label in cell biology. J. Nanobiotechnol. 2006, 4, 11. [Google Scholar] [CrossRef] [Green Version]
  23. Yu, C.; Yu, M.; Li, C.; Liu, X.; Yang, J.; Yang, P.; Lin, J. Facile sonochemical synthesis and photoluminescent properties of lanthanide orthophosphate nanoparticles. J. Solid State Chem. 2009, 182, 339–342. [Google Scholar] [CrossRef]
  24. Wang, X.; Gao, M. A facile route for preparing rhabdophane rare earth phosphate nanorods. J. Mater. Chem. 2006, 16, 1360–1365. [Google Scholar] [CrossRef]
  25. Hou, Z.; Wang, L.; Lian, H.; Chai, R.; Zhang, C.; Cheng, Z.; Lin, J. Preparation and luminescence properties of Ce3+ and/or Tb3+ doped LaPO4 nanofibers and microbelts by electrospinning. J. Solid State Chem. 2009, 182, 698–708. [Google Scholar] [CrossRef]
  26. Xing, Y.; Li, M.; Davis, S.A.; Mann, S. Synthesis and characterization of cerium phosphate nanowires in microemulsion reaction media. J. Phys. Chem. B 2006, 110, 1111–1113. [Google Scholar] [CrossRef]
  27. Liu, X.; Wang, Q.; Gao, Z.; Sun, J.; Shen, J. Fabrication of lanthanide phosphate nanocrystals with well-controlled morphologies by layer-by-layer adsorption and reaction method at room temperature. Cryst. Growth Des. 2009, 9, 3707–3713. [Google Scholar] [CrossRef]
  28. Huong Tran, T.; Anh Tran, K.; Khuyen Hoang, T.; Hien Pham, T.; Minh Le, Q. Fabrication and properties of terbium phosphate nanorods. Adv. Nat. Sci. Nanosci. Nanotechnol. 2012, 3, 015010. [Google Scholar] [CrossRef]
  29. Gómez-Morales, J.; Verdugo-Escamilla, C.; Fernández-Penas, R.; Parra-Milla, C.M.; Drouet, C.; Iafisco, M.; Oltolina, F.; Prat, M.; Fernández-Sánchez, J.F. Bioinspired crystallization, sensitized luminescence and cytocompatibility of citrate-functionalized Ca-substituted europium phosphate monohydrate nanophosphors. J. Colloid Interface Sci. 2019, 538, 174–186. [Google Scholar] [CrossRef] [Green Version]
  30. López-Macipe, A.; Gómez-Morales, J.; Rodríguez-Clemente, R. Nanosized hydroxyapatite precipitation from homogeneous calcium/citrate/phosphate solutions using microwave and conventional heating. Adv. Mater. 1998, 10, 49–53. [Google Scholar] [CrossRef]
  31. Delgado-López, J.M.; Iafisco, M.; Rodríguez, I.; Prat, M.; Gómez-Morales, J.; Tampieri, A. Crystallization of bioinspired citrate-functionalized nanoapatites with tailored carbonate content. Acta Biomater. 2012, 8, 3491–3499. [Google Scholar] [CrossRef]
  32. Martínez-Casado, F.J.; Iafisco, M.; Delgado-López, J.M.; Martínez-Benito, C.; Ruiz-Pérez, C.; Colangelo, D.; Oltolina, F.; Prat, M.; Gómez-Morales, J. Bioinspired citrate–apatite nanocrystals doped with divalent transition metal ions. Cryst. Growth Des. 2016, 1, 145–153. [Google Scholar] [CrossRef]
  33. Delgado-López, J.M.; Frison, R.; Cervellino, A.; Gómez-Morales, J.; Guagliardi, A.; Masciocchi, N. Crystal size, morphology, and growth mechanism in bio-inspired apatite nanocrystals. Adv. Funct. Mater. 2014, 24, 1090–1099. [Google Scholar] [CrossRef]
  34. Hu, Y.Y.; Rawal, A.; Schmidt-Rohr, K. Strongly bound citrate stabilizes the apatite nanocrystals in bone. Proc. Natl. Acad. Sci. USA 2010, 107, 22425–22429. [Google Scholar] [CrossRef] [Green Version]
  35. Coelho, A.A. TOPAS and TOPAS-Academic: An optimization program integrating computer algebra and crystallographic objects written in C++. J. Appl. Cryst. 2018, 51, 210–218. [Google Scholar] [CrossRef] [Green Version]
  36. Combes, C.; Bareille, R.; Rey, C. Calcium carbonate–calcium phosphate mixed cement compositions for bone reconstruction. J. Biomed. Mater. Res. Part A 2006, 79, 318–328. [Google Scholar] [CrossRef] [Green Version]
  37. Young, R.A. Introduction to the Rietveld method. In The Rietveld Method; Young, R.A., Ed.; Oxford University Press: Oxford, UK, 1993; pp. 1–38. [Google Scholar]
  38. Clavier, N.; del Mesbah, A.; Szenknect, S.; Dacheux, N. Monazite, rhabdophane, xenotime & churchite: Vibrational spectroscopy of gadolinium phosphate polymorphs. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2018, 205, 85–94. [Google Scholar] [CrossRef]
  39. Assaaoudi, H.; Ennaciri, A.; Rulmont, A. Vibrational spectra of hydrated rare earth orthophosphates. Vibrat. Spectr. 2001, 2, 81–90. [Google Scholar] [CrossRef]
  40. Donnelly, F.C.; Purcell-Milton, F.; Framont, V.; Cleary, O.; Dunne, P.W.; Gun’k, Y.K. Synthesis of CaCO3 nano- and micro-particles by dry ice carbonation. Chem. Commun. 2017, 53, 6657–6660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Tada, H.; Bronkema, J.L.; Bell, A. Application of in situ surface-enhanced Raman spectroscopy (SERS) to the study of citrate oxidation on silica-supported silver nanoparticles. Catal. Lett. 2004, 92, 93–99. [Google Scholar] [CrossRef]
  42. Lucas, S.; Champion, E.; Bernache-Assollant, D.; Leroy, G. Rare earth phosphate powders RePO4·nH2O (Re = La, Ce or Y) II. Thermal behavior. J. Solid State Chem. 2004, 177, 312–1320. [Google Scholar] [CrossRef]
  43. Cho, J.; Kim, C.H. Solid-state phase transformation mechanism from hexagonal GdPO4:Eu3+ nanorods to monoclinic nanoparticles. RSC Adv. 2014, 4, 31385–31392. [Google Scholar] [CrossRef]
  44. Kijkowska, R. Thermal decomposition of lanthanide orthophosphates synthesized through crystallisation from phosphoric acid solution. Thermochim. Acta 2003, 404, 81–88. [Google Scholar] [CrossRef]
  45. Colomer, M.T.; Delgado, I.; Ortiz, A.L.; Fariñas, J.C. Microwave-assisted hydrothermal synthesis of single-crystal nanorods of rhabdophane-type Sr-doped LaPO4.nH2O. J. Am. Ceram. Soc. 2014, 97, 750–758. [Google Scholar] [CrossRef]
  46. Jabalera, Y.; Oltolina, F.; Prat, M.; Jimenez-Lopez, C.; Fernández-Sánchez, J.F.; Choquesillo-Lazarte, D.; Gómez-Morales, J. Eu-Doped Citrate-Coated Carbonated Apatite Luminescent Nanoprobes for Drug Delivery. Nanomaterials 2020, 10, 199. [Google Scholar] [CrossRef] [Green Version]
  47. Lundqvist, V.; Stigler, J.; Elia, G.; Lynch, I.; Cedervall, T.; Dawson, K.A. Nanoparticle size and surface properties determine the protein corona with possible implications for biological impacts. Proc. Natl. Acad. Sci. USA 2008, 105, 14265–14270. [Google Scholar] [CrossRef] [Green Version]
  48. Hemmilä, I.; Dakubu, S.; Mukkala, V.-M.; Siitari, H.; Lövgren, T. Europium as a label in time-resolved immunofluorometric assays. Anal. Biochem. 1984, 137, 335. [Google Scholar] [CrossRef]
  49. Siqueira, K.P.F.; Lima, P.P.; Ferreira, R.A.S.; Carlos, L.D.; Bittar, E.M.; Matinaga, F.M.; Paniago, R.; Krambrock, K.; Moreira, R.L.; Dias, A.J. Influence of the Matrix on the Red Emission in Europium Self-Activated Orthoceramics. Phys. Chem. C 2015, 119, 17825. [Google Scholar] [CrossRef]
  50. Siqueira, K.P.F.; Lima, P.P.; Ferreira, R.A.S.; Carlos, L.D.; Bittar, E.M.; Granado, E.; González, J.C.; Abelenda, A.; Moreira, R.L.; Dias, A. Lanthanide Orthoantimonate Light Emitters: Structural, Vibrational, and Optical Properties. Chem. Mater. 2014, 26, 6351. [Google Scholar] [CrossRef]
  51. Medina-Velazquez, D.Y.; Caldiño, U.; Morales-Ramirez, A.; Reyes-Miranda, J.; López, R.E.; Escudero, R.; Ruiz-Guerrero, R.; Morales Perez, M.F. Synthesis of luminescent terbium-thenoyltriflouroacetone MOF nanorods for green laser application. Opt. Mater. 2019, 87, 3–10. [Google Scholar] [CrossRef] [Green Version]
  52. Richardson, F.S. Terbium(III) and europium(III) ions as luminescent probes and stains for biomolecular systems. Chem. Rev. 1982, 82, 541. [Google Scholar] [CrossRef]
  53. Guilbault, G.G. General Aspects of Luminescence Spectroscopy, Practical Fluorescence; Guilbault, G.G., Ed.; Marcel Dekker: New York, NY, USA, 1990. [Google Scholar]
  54. ISO 10993-5 Biological Evaluation of Medical Devices Part 5: Tests for In Vitro Cytotoxicity; International Standard Organization: Geneva, Switzerland, 2009.
  55. López-Macipe, A.; Gómez-Morales, J.; Rodriguez-Clemente, R. The role of pH in the adsorption of citrate ions on hydroxyapatite. J. Colloid Interface Sci. 1998, 200, 114–120. [Google Scholar] [CrossRef]
  56. Borroni, E.; Miola, M.; Ferraris, S.; Ricci, G.; Žužek Rožman, K.; Kostevšek, N.; Catizone, A.; Rimondini, L.; Prat, M.; Verné, E.; et al. Tumor targeting by lentiviral vectors combined with magnetic nanoparticles in mice. Acta Biomater. 2017, 59, 303–316. [Google Scholar] [CrossRef] [PubMed]
  57. Mooney, C.L. X-ray diffraction study of cerium phosphate and related crystals. I. Hexagonal modification. Acta Cryst. 1950, 3, 337–340. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Evolution of XRD patterns with maturation time (4 h, green; 24 h, red; 96 h, blue) of samples precipitated from Ca2+/Tb3+/cit/phosphate/carbonate solutions in presence of increasing concentrations of Ca2+: (a) x = 0.01 M Ca2+; (b) x = 0.03 M Ca2+; (c) x = 0.05 M Ca2+, and (d) x = 0.07 M Ca2+. * hexagonal phase TbPO4·nH2O PDF 20-1244.
Figure 1. Evolution of XRD patterns with maturation time (4 h, green; 24 h, red; 96 h, blue) of samples precipitated from Ca2+/Tb3+/cit/phosphate/carbonate solutions in presence of increasing concentrations of Ca2+: (a) x = 0.01 M Ca2+; (b) x = 0.03 M Ca2+; (c) x = 0.05 M Ca2+, and (d) x = 0.07 M Ca2+. * hexagonal phase TbPO4·nH2O PDF 20-1244.
Nanomaterials 11 00322 g001
Figure 2. Crystalline domain sizes of Ca2+:TbPO4·nH2O nanoparticles prepared with, (a) x = 0.01 M; (b) x = 0.03 M; (c) x = 0.05 M and (d) x = 0.07 M Ca2+ for 96 h, calculated by analysis of the full XRD pattern using TOPAS 6.0. The two XRD vaterite reflections are excluded. Rwp, weighted profile R-factor as described in Young [37].
Figure 2. Crystalline domain sizes of Ca2+:TbPO4·nH2O nanoparticles prepared with, (a) x = 0.01 M; (b) x = 0.03 M; (c) x = 0.05 M and (d) x = 0.07 M Ca2+ for 96 h, calculated by analysis of the full XRD pattern using TOPAS 6.0. The two XRD vaterite reflections are excluded. Rwp, weighted profile R-factor as described in Young [37].
Nanomaterials 11 00322 g002
Figure 3. (ac) TEM micrographs of cit-Ca2+:TbPO4·H2O nanorods prepared with different Ca2+ doping concentrations at 96 h: (a) x = 0.01 M; (b) x = 0.03 M; (c) x = 0.05 M. Insets display SAED patterns of the nanoparticles. (d) HAADF-STEM micrograph of sample x = 0.01 M Ca2+. (e,f,h,i) EDX element mapping analysis of Tb, Ca, P, and O in nanoparticles of image (d). (g) HR-TEM image showing lattice fringes and d-spacing of 2.78 Å corresponding to plane (102) of nanocrystals of sample x = 0.01 M Ca2+. Inset in (g) shows the corresponding Fourier transform (FFT) image displaying the (102) plane.
Figure 3. (ac) TEM micrographs of cit-Ca2+:TbPO4·H2O nanorods prepared with different Ca2+ doping concentrations at 96 h: (a) x = 0.01 M; (b) x = 0.03 M; (c) x = 0.05 M. Insets display SAED patterns of the nanoparticles. (d) HAADF-STEM micrograph of sample x = 0.01 M Ca2+. (e,f,h,i) EDX element mapping analysis of Tb, Ca, P, and O in nanoparticles of image (d). (g) HR-TEM image showing lattice fringes and d-spacing of 2.78 Å corresponding to plane (102) of nanocrystals of sample x = 0.01 M Ca2+. Inset in (g) shows the corresponding Fourier transform (FFT) image displaying the (102) plane.
Nanomaterials 11 00322 g003
Figure 4. (a) FTIR and (b) Raman spectra of samples precipitated at 96 h using Ca2+-doping concentrations x = 0.01 M (purple), x = 0.03 M (green), x = 0.05 M (red), x = 0.07 M (light blue).
Figure 4. (a) FTIR and (b) Raman spectra of samples precipitated at 96 h using Ca2+-doping concentrations x = 0.01 M (purple), x = 0.03 M (green), x = 0.05 M (red), x = 0.07 M (light blue).
Nanomaterials 11 00322 g004
Figure 5. (a) Volume-based CSD and (b) cumulative volume-based CSD of cit-Ca2+:TbPO4·nH2O nanocrystals prepared with Ca2+ doping concentrations x = 0.01 (purple), 0.03 (green), 0.05 (red) and 0.07 M (light blue) for 96 h, measured in aqueous suspensions (pHs 8.4, 8.9, 8.3, and 7.80, respectively). (cf) ζ-potential versus pH of aqueous suspensions of cit-Ca2+:TbPO4·nH2O nanocrystals.
Figure 5. (a) Volume-based CSD and (b) cumulative volume-based CSD of cit-Ca2+:TbPO4·nH2O nanocrystals prepared with Ca2+ doping concentrations x = 0.01 (purple), 0.03 (green), 0.05 (red) and 0.07 M (light blue) for 96 h, measured in aqueous suspensions (pHs 8.4, 8.9, 8.3, and 7.80, respectively). (cf) ζ-potential versus pH of aqueous suspensions of cit-Ca2+:TbPO4·nH2O nanocrystals.
Nanomaterials 11 00322 g005
Figure 6. Excitation (dashed lines) and emission (solid lines) uncorrected spectra of solid cit-Ca2+:TbPO4·nH2O samples prepared with x = 0.01 M Ca2+ at maturation times of 4 h, 24 h, 96 h and 7 days.
Figure 6. Excitation (dashed lines) and emission (solid lines) uncorrected spectra of solid cit-Ca2+:TbPO4·nH2O samples prepared with x = 0.01 M Ca2+ at maturation times of 4 h, 24 h, 96 h and 7 days.
Nanomaterials 11 00322 g006
Figure 7. Luminescence properties of cit-Ca2+:TbPO4·nH2O dispersed in aqueous media with x = 0.01 M Ca2+, 96 h maturation time and 25 °C versus pH; (a) Excitation (dashed line) and emission (solid line) spectra; (b) Variation of R.L.I. versus pH; (c) Luminescence decay curve; and (d) Variation of luminescence lifetime with the pH.
Figure 7. Luminescence properties of cit-Ca2+:TbPO4·nH2O dispersed in aqueous media with x = 0.01 M Ca2+, 96 h maturation time and 25 °C versus pH; (a) Excitation (dashed line) and emission (solid line) spectra; (b) Variation of R.L.I. versus pH; (c) Luminescence decay curve; and (d) Variation of luminescence lifetime with the pH.
Nanomaterials 11 00322 g007
Figure 8. Viability of GTL-16 cells (a) and of A549 cells (b) incubated with cit-Ca2+:TbPO4·nH2O particles prepared with different Ca2+-doping concentration for three days. Viability was assessed in MTT assays. Data represent means ± SD of three independent experiments performed in triplicate and statistical analyses were carried on using One-way ANOVA, with Bonferroni comparison test. For statistical analysis all data were compared to untreated samples and only samples treated with H2O2 displayed statistically significant difference (**** p = 0.0005).
Figure 8. Viability of GTL-16 cells (a) and of A549 cells (b) incubated with cit-Ca2+:TbPO4·nH2O particles prepared with different Ca2+-doping concentration for three days. Viability was assessed in MTT assays. Data represent means ± SD of three independent experiments performed in triplicate and statistical analyses were carried on using One-way ANOVA, with Bonferroni comparison test. For statistical analysis all data were compared to untreated samples and only samples treated with H2O2 displayed statistically significant difference (**** p = 0.0005).
Nanomaterials 11 00322 g008
Table 1. Quantitative analyses of Ca, P, Tb, C, H, and estimated percent weights of structural citrate, H2O, CaCO3 and hydration number n of the samples prepared with Ca2+ doping concentration x ranging from 0.01 to 0.07 M, at 96 h.
Table 1. Quantitative analyses of Ca, P, Tb, C, H, and estimated percent weights of structural citrate, H2O, CaCO3 and hydration number n of the samples prepared with Ca2+ doping concentration x ranging from 0.01 to 0.07 M, at 96 h.
x [Ca2+] mol/LCa (wt%)P (wt%)Tb (wt%)C (wt%)H (wt%)Cit (wt%)H2O str (wt%)CaCO3 (wt%)n
0.010.4210.561.40.460.721.215.000.88
0.032.9110.939.80.420.741.105.440.20.86
0.056.76, 6.09 *12.027.50.520.751.126.680.760.93
0.0710.32, 8.11 *11.4516.91.721.043.966.331.750.95
* wt% of dopant Ca in the TbPO4·nH2O structure.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gómez-Morales, J.; Fernández-Penas, R.; Romero-Castillo, I.; Verdugo-Escamilla, C.; Choquesillo-Lazarte, D.; D’Urso, A.; Prat, M.; Fernández-Sánchez, J.F. Crystallization, Luminescence and Cytocompatibility of Hexagonal Calcium Doped Terbium Phosphate Hydrate Nanoparticles. Nanomaterials 2021, 11, 322. https://doi.org/10.3390/nano11020322

AMA Style

Gómez-Morales J, Fernández-Penas R, Romero-Castillo I, Verdugo-Escamilla C, Choquesillo-Lazarte D, D’Urso A, Prat M, Fernández-Sánchez JF. Crystallization, Luminescence and Cytocompatibility of Hexagonal Calcium Doped Terbium Phosphate Hydrate Nanoparticles. Nanomaterials. 2021; 11(2):322. https://doi.org/10.3390/nano11020322

Chicago/Turabian Style

Gómez-Morales, Jaime, Raquel Fernández-Penas, Ismael Romero-Castillo, Cristóbal Verdugo-Escamilla, Duane Choquesillo-Lazarte, Annarita D’Urso, Maria Prat, and Jorge Fernando Fernández-Sánchez. 2021. "Crystallization, Luminescence and Cytocompatibility of Hexagonal Calcium Doped Terbium Phosphate Hydrate Nanoparticles" Nanomaterials 11, no. 2: 322. https://doi.org/10.3390/nano11020322

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