Next Article in Journal
Improved Photostability in Fluorinated 2D Perovskite Single Crystals
Next Article in Special Issue
Nanostructural Materials with Rare Earth Ions: Synthesis, Physicochemical Characterization, Modification and Applications
Previous Article in Journal
Excitation Wavelength and Intensity-Dependent Multiexciton Dynamics in CsPbBr3 Nanocrystals
Previous Article in Special Issue
Nanomaterials Application in Orthodontics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Quenching of the Eu3+ Luminescence by Cu2+ Ions in the Nanosized Hydroxyapatite Designed for Future Bio-Detection

1
Institute of Low Temperature and Structure Research, PAS, Okolna 2, 50-422 Wroclaw, Poland
2
Faculty of Chemistry, University of Wroclaw, Joliot-Curie 14, 50-383 Wroclaw, Poland
3
International Institute of Translational Medicine, Jesionowa 11 St., 55–124 Malin, Poland
*
Authors to whom correspondence should be addressed.
Nanomaterials 2021, 11(2), 464; https://doi.org/10.3390/nano11020464
Submission received: 31 December 2020 / Revised: 5 February 2021 / Accepted: 8 February 2021 / Published: 11 February 2021

Abstract

:
The hydroxyapatite nanopowders of the Eu3+-doped, Cu2+-doped, and Eu3+/Cu2+-co-doped Ca10(PO4)6(OH)2 were prepared by a microwave-assisted hydrothermal method. The structural and morphological properties of the products were investigated by X-ray powder diffraction (XRD), transmission electron microscopy techniques (TEM), and infrared spectroscopy (FT-IR). The average crystal size and the unit cell parameters were calculated by a Rietveld refinement tool. The absorption, emission excitation, emission, and luminescence decay time were recorded and studied in detail. The 5D07F2 transition is the most intense transition. The Eu3+ ions occupied two independent crystallographic sites in these materials exhibited in emission spectra: one Ca(1) site with C3 symmetry and one Ca(2) sites with Cs symmetry. The Eu3+ emission is strongly quenched by Cu2+ ions, and the luminescence decay time is much shorter in the case of Eu3+/Cu2+ co-doped materials than in Eu3+-doped materials. The luminescence quenching mechanism as well as the schematic energy level diagram showing the Eu3+ emission quenching mechanism using Cu2+ ions are proposed. The electron paramagnetic resonance (EPR) technique revealed the existence of at least two different coordination environments for copper(II) ion.

1. Introduction

Apatite-type materials can be applied in many industrial fields, e.g., as sorbents, biocompatible and biodegradable materials for bone and teeth reconstruction, catalysts, materials for the wastewater treatment, fertilizers, and luminescent materials [1,2]. Hydroxyapatite (Ca10(PO4)6(OH)2—abbr. HAp) is used in medicine as a bone implant material due to its biocompatibility, bioactivity, and similarity to bone mineral [3,4]. However, it is still widely investigated in order to improve its properties by obtaining appropriate grain size, morphology, mechanical strength, and solubility and by adding some dopants that are, e.g., naturally built into bone apatite, ions possessing antibacterial properties, or ions enabling bio-imaging [5,6]. Infections after grafting of bone implant material are a serious problem in surgery. The idea is that doping with antibacterial ions into the grafted biomaterial will prevent bacterial biofilm formation and infection development. Inorganic antibacterial agents possess advantages such as stability and safety. The antibacterial agents include ions such as copper, silver, and zinc [7,8,9], and several studies have shown that they can play important roles in the prevention or minimization of initial bacterial adhesion [10,11]. Metal ions can react with microbial membrane, causing structural changes and permeability. Then, Ag+ and Cu2+ ions have the ability to complex anions such as –NH2, –S–S–, and –CONH– of the proteins or enzymes in the bacterial cells. It provides bacterial DNA and RNA damage and inhibits proliferation [7,11,12]. Copper is an essential microelement that is involved in many metabolic processes that taken place in human bodies [7,11,13,14]. However, copper ions may have potentially toxic effects at higher amount in human beings due to their ability to generate ROSs (Reactive Oxygen Species). On the other hand, recent studies have shown that copper-doped apatite-type materials are very promising as a new kind of ow-toxic pigment that can be used in the paint and varnish industry [1,15,16,17,18].
Lanthanide(III)-doped nanomaterials are promising candidates for fluorescent bio-labels due to their stable luminescence over time, high photochemical stability, sharp emission peak, low levels of photobleaching, and toxicity compared with organic fluorophores. Eu3+ ions are structural and luminescence probe-sensitive to changes in the local environment around the ion. Furthermore, the luminescence of Eu3+ ions are identified by a narrow emission band and long lifetimes of the excited state [19,20].
Apatite is a big family of compounds, and it is widely investigated due to its outstanding properties such as good biocompatibility or possibility to be doped with different ions in a broad concentration range for applications in the industry, in medicine, etc. There are a lot of papers focusing on apatite synthesis [21,22,23,24]; its doping with antibacterial ions [9,10,25]; as well as its doping with luminescence ions such as Eu3+ [5,19,26,27], Tb3+ [28,29], Eu3+/Tb3+ [30], Er3+/Yb3+, or Eu3+/Cu2+ [17]. Moreover, there is a lot of research focusing on anion-substituted apatite such as silicate [31,32], vanadate [33], borate [34], or carbonate [35].
In the presented work, the synthesis, structural, morphological, and luminescence properties of Eu3+/Cu2+ co-doped HAp were investigated attentively. To the best of our knowledge, this is the first time that the quenching mechanism in this system has been elucidated.

2. Materials and Methods

2.1. Synthesis

The Ca10(PO4)6(OH)2 nanopowders doped with Eu3+ and Cu2+ ions were synthesized by a microwave-assisted hydrothermal method. The starting materials used were CaCO3 (99.0%, Alfa Aesar, Karlsruhe, Germany), NH4H2PO4 (99.0%, Fluka, Bucharest, Romania), Eu2O3 (99.99%, Alfa Aesar, Karlsruhe, Germany), Cu(NO3)2∙2.5H2O (98.0–102.0%, Alfa Aesar, Karlsruhe, Germany), and NH3∙H2O (99%, Avantor, Gliwice, Poland) as a pH regulation reagent. The concentration of dopants was calculated based on inductively coupled plasma-optical emission spectrometer (ICP-OES) results. The concentrations of europium ions were 0.5 mol%, 1 mol%, 2 mol%, and 5 mol%, and the concentrations of the cooper ions were 2 mol% and 5 mol% to the overall molar content of calcium cations. First, the stoichiometric amounts of CaCO3 as well as Eu2O3 were separately digested in excess of HNO3 (suprapur Merck, Darmstadt, Germany) to obtain water-soluble nitrates. The obtained europium nitrate hydrate was recrystallized three times to remove excess HNO3. Then, the stoichiometric amount of Eu(NO3)3 was dissolved in deionized water, and then, the Cu(NO3)2 was added to the stoichiometric amount of calcium nitrate. After this, NH4H2PO4 was added to the abovementioned mixture and the pH value was adjusted to 9 by ammonia. The suspension was transferred to a Teflon vessel and was placed into the microwave reactor (ERTEC MV 02-02, Wrocław, Poland). The reaction system was heat-treated at 280 °C for 90 min under autogenous pressure of 60 atm. The obtained product was washed several times with deionized water and dried at 70 °C for 24 h.

2.2. Powder Characterization

The crystal structure and phase purity were studied using a PANalytical X’Pert Pro diffractometer (Malvern Panalytical Ltd., Malvern, UK) equipped with Ni-filtered Cu Kα radiation (V = 40 kV and I = 30 mA). The recorded X-ray powder diffraction patterns (XRD) were compared with the reference standard of hexagonal calcium hydroxyapatite (P63/m) from the Inorganic Crystal Structure Database (ICSD-2866) and analyzed. Rietveld structural refinement was performed with the aid of a Maud program (version 2.93) (University of Trento-Italy, Department of Industrial Engineering, Trento, Italy) [36,37] based on the apatite hexagonal crystal structure with better approximation and indexing of the Crystallographic Information File (CIF). The quality of structural refinement was checked by R-values (Rw, Rwnb, Rall, Rnb, and σ), which were followed to get a structural refinement with better quality and reliability.
The morphology was investigated by high-resolution transmission electron microscopy (HRTEM) using a Philips CM-20 SuperTwin microscope (Eindhoven, The Netherlands), operating at 200 kV. The specimen for the HRTEM measurement was obtained by dispersing a small amount of powder in methanol and by putting a droplet of the suspension onto a copper microscope grid covered with carbon.
Fourier transform infrared spectra were measured using a Thermo Scientific Nicolet iS50 FT-IR spectrometer (Waltham, MA, USA) in the range of 4000–400 cm−1 at 295 K. Absorption spectra were recorded with an Agilent Cary 5000 spectrophotometer, employing a spectral bandwidth (SBW) of 0.1 nm in the visible and ultraviolet range and of 0.7 nm in the infrared. The spectra were recorded at room temperature.
The excitation spectra were recorded with the aid of an FLS980 Fluorescence Spectrometer (Edinburgh Instruments, Kirkton Campus, UK) equipped with 450 W Xenon lamp. The excitation of 300 mm focal length monochromator was in Czerny–Turner configuration and the excitation arm was supplied with holographic grating of 1800 lines/mm grating blazed at 250 nm. The excitation spectra were corrected to the excitation source intensity. The emission spectra were measured by using a Hamamatsu PMA-12 photonic multichannel analyzer (Hamamatsu, Hamamatsu City, Japan) equipped with BT-CCD line (Hamamatsu, Hamamatsu City, Japan). As an excitation source, a pulsed 266 nm line of Nd:YAG laser (3rd harmonic; LOTIS TII, Minsk, Belarus) was chosen (ƒ = 10 Hz, t < 10 ns). The detection setup was calibrated and had a flat response for the whole working range (350–1100 nm). The measurements were carried out at 300 K.
The time-resolved luminescence spectrum was obtained by recording decay curves during changing observed wavelength and by creating a two-dimensional map (i.e., intensity vs. time and wavelength). It was recorded by an in-house developed software that controlled the equipment. A Dongwoo Optron DM711 monochromator (Hoean-Daero, Opo-Eup, Gyeonggi-Do, Korea) with a focal length of 750 mm was used to select the observed wavelength, while luminescence decay curves were acquired with a Hamamatsu R3896 photomultiplier (Hamamatsu, Hamamatsu City, Japan) connected to a digital Tektronix MDO 4054B oscilloscope (Bracknell, UK). An optical parametric oscillator Opotek Opolette 355 LD (Carlsbad, CA, USA) emitting 5 ns pulses was used as an excitation source.
The luminescence kinetics were measured by using a Jobin-Yvon THR1000 monochromator (HORIBA Jobin-Yvon, Palaiseu, France) equipped with a Hamamatsu R928 photomultiplier (Hamamatsu, Hamamatsu City, Japan) as a detector and a LeCroy WaveSurfer as a digital oscilloscope (Teledyne LeCroy, Chestnut Ridge, NY, USA). As an excitation source, a pulsed 266 nm line from an Nd:YAG laser was used. The luminescence kinetics were monitored at 618 nm according to the most intense electric dipole transition (5D07F2), and the effective emission lifetimes were calculated using the following equation:
τ m = 0 t I ( t ) d t 0 I ( t ) d t 0 t max t I ( t ) d t 0 t max I ( t ) d t
where I(t) is the luminescence intensity at time t corrected for the background and the integrals are calculated over the range of 0 < t < tmax, where tmax >> τm.
The effective content of elements was determined by using an Agilent 720 bench-top optical emission spectrometer with inductively coupled Ar plasma (Ar-ICP-OES) and was corrected to an effective value. The ICP standard solutions were used to record the calibration curves to determine the Ca2+, P5+, Cu2+, and Eu3+ ion content. The samples for elemental analysis were prepared by digesting in the pure HNO3 acid (65% suprapur Merck).
The electron paramagnetic resonance (EPR) spectra were measured at 295 K and 77 K using a Bruker Elexsys 500 CW-EPR (Bruker GmbH, Rheinstetten, Germany) spectrometer operating at the X-band frequency (≈9.7 GHz), equipped with frequency counter (E 41 FC) and NMR teslameter (ER 036TM). The spectra were measured with a modulation frequency of 100 kHz, microwave power of 10 mW, modulation amplitude of 10 G, time constant of 40 ms, and a conversion time of 160 ms. The first derivative of the absorption power was recorded as a function of the magnetic field value. An analysis of the EPR spectra was carried out using the WinEPR software package, version 1.26b (Bruker WinEPR GmbH, Rheinstetten, Germany).

3. Results and Discussion

3.1. Structural Analysis

The structural characterization of the HAp nanocrystals doped with xEu3+ (where x = 0.5, 1, and 3 mol%) and co-doped with xEu3+ and yCu2+ (where x = 0.5, 1, and 4 mol% and y = 0.5 and 1 mol%) was carried out by powder X-Ray diffraction measurements as a function of doping ion concentration (see Figure 1). Detectable crystallinity and pure hexagonal phase corresponding to the reference standard (ICSD—180315 [38]) were observed. Only in the case of the 4 mol% Eu3+/0.5 mol% Cu2+:HAp, an extra peak at 29.5° of 2θ was observed (assigned as asterisk in Figure 1).
Structural refinement was performed to obtain the unit cell parameters and the average grain sizes of synthesized materials. Hexagonal phase formation and the successful incorporation of Eu3+ and Cu2+ ions were verified. The theoretical fit with the observed XRD pattern was found to be in good agreement, which indicated the success of the Rietveld refinement method (see Figure 2). More details are displayed in Table 1. As can be seen, it was possible to observe an increase in the cell volume and a parameters with the increase in Eu3+ ion concentration in single-doped materials, which was caused by a smaller ionic radii of the dopant (Ca2+ (coordination number—CN9), 1.18 Å; Eu3+ (CN9), 1.12 Å; Ca2+ (CN7), 1.06 Å; and Eu3+ (CN7), 1.01 Å) [39]. Moreover, shrinkage of the average grain size with an increase in the Eu3+ ion concentration in the host lattice was observed. In the case of co-doped materials, no straightforward dopant concentration dependence on cell parameters (a, c, and V) or average grain size was observed.
The morphology of the calcium hydroxyapatite was investigated by HRTEM. Nanoparticles are crystalline in nature and elongated, as can be seen in Figure 3. The particle size distribution is relatively wide, and the mean grain sizes of particle is in the range between 60 and 120 nm in length and about 40 nm in width.
The infrared spectra of the copper-doped, europium-doped, and co-doped hydroxyapatite materials are presented in Figure 4. The most intense peaks are the triply degenerated antisymmetric stretching bands of phosphate groups ν3(PO43−) located at 1044.5 cm−1 and 1097.8 cm−1. The peaks observed at 566.0 cm−1 and 603.1 cm−1 correspond to the triply degenerated ν4(PO43−) vibrations. The peaks at 963.0 cm−1 are assigned to the non-degenerated symmetric stretching ν1(PO43−) band. Two peaks corresponding to OH group at 3571.5 cm−1 and 633.5 cm−1 are observed on the infrared spectra. The existence of these peaks clearly confirms the hydroxyapatite structure with a hydroxyl group in the host lattice. The broad bands between 3690 and 3290 cm−1 were connected with H2O vibration.

3.2. Absorption, Excitation, and Emission Spectra

The absorption spectra of the pure, copper-doped, europium-doped, and co-doped hydroxyapatite nanopowders were recorded in the visible range from 350 nm to 800 nm at room temperature (see Figure 5). The pure hydroxyapatite matrix is transparent for these wavelengths. The copper-doped materials absorbed the blue radiation in the range from 350 nm to 420 nm. All materials are relatively transparent for the radiation from 450 nm to 550 nm of the wavelength. The copper-doped materials absorbed the radiation from 550 nm to 800 nm, and the absorption coefficient increases with the increase in wavelength. The broad absorption band is attributed to the 2E → 2T2 intra-configurational (d-d) transition of the Cu2+ ions [40,41]. In the absorption spectra, the peaks related to the 4f-4f transitions of Eu3+ ions are observed. These peaks are attributed to the following transitions: the 7F05D4, 5L8 at 362 nm and 7F05G6, 5L7, 5G3 at 376 nm. The 7F05L6 transition with a maximum at 394 nm was observed in the case of europium and the copper co-doped materials. This transition is the most intense f-f transition of Eu3+ ions.
The excitation emission spectra, which were recorded at room temperature by monitoring the intense red emission at 618 nm (5D07F2), of investigated materials are presented in Figure 6. The representative excitation spectra of the 1 mol% Eu3+:HAp and 1 mol% Eu3+/1 mol% Cu2+:HAp are presented. As demonstrated, the excitation spectra consisted of visible intra-configurational 4f-4f transitions with sharp lines characteristic of Eu3+ ions. Particularly, these narrow bands located at around 320, 363, 383, 395, 416, and 466 nm originated from the 7F0 5HJ; 7F0 5D4, 5L8; 7F0 G2, 5L7, 5G3; 7F0 5L6; 7F0 5D3 transitions of Eu3+ ions, respectively [31,35]. The absorption peak of the 7F0 5HJ transition at 320 nm indicates that the energy band-gap of HAp is considerably larger than that in, e.g., Eu2Ti2O7 oxide [42], in which the 5H3,6-related transition peak is completely masked by the charge transfer band due to its lower energy band-gap nature. The f-f electron transitions are weakly affected by the crystal field; thus, their positions remain almost steady due to good isolations of lanthanide’s f orbitals by an external shell [19,31,43]. As can be seen, the intensity of the emission excitation spectra is much lower in the case of the co-doped material than in that single doped with Eu3+ in the HAp host.
The spectroscopic properties of Eu3+ ions allow us to receive vital information about the symmetry of the Eu3+ ions surrounding the crystal lattice; the amount of crystallographic positions; and therefore, potential sites of substitution, structural changes occurring in the matrix caused by external factors, etc. The emission spectra of the Eu3+ ions consist of characteristic bands present in the red region of the electromagnetic radiation assigned to the electron transitions developing in the 4f-4f shell of Eu3+ ions. The 5D07F0,1,2 transitions are the most important in analysis, particularly in correlation with the structural properties. The 5D07F0 transition can provide direct information about the number of crystallographic sites occupied by Eu3+ ions in the host lattice [19].
The emission spectra of y mol% Eu3+:HAp (where y—0.5, 1, and 3 mol%) and x mol% Eu3+/y mol% Cu2+:HAp (where x—0.5, 1, and 4 mol% and y—0.5 and 1 mol%) were measured by excitation wavelength at 266 nm at room temperature and are shown in Figure 7. The spectra were normalized to the 5D0 7F1 magnetic dipole transition. The emission spectra were dominated by an intense red emission band situated at about 618 nm corresponding to the 5D07F2 transition of Eu3+ ions. Meanwhile, four weaker emission bands peaking at around 578, 589, 652, and 698 nm were also detected and ascribed to the 5D07F0, 5D07F1, 5D07F3, and 5D07F4 transitions of Eu3+ ions, respectively [31,35]. The presence of a 5D07F0 transition confirms that europium ions are located in a low-symmetry environment. Furthermore, the number of lines directly indicate the number of occupied crystallographic positions in the investigated lattice by Eu3+ ions. In the apatite molecule, ten calcium atoms are found in two non-equal crystallographic positions, in agreement with the results showed in Figure 7.
In Figure 8, the time-resolved emission spectrum of the 1 mol% Eu3+/1 mol% Cu2+:HAp is presented.

3.3. Decay Profiles

The luminescence decay curves were registered and analyzed for the synthesized materials to determine the comprehensive characteristics of the luminescence properties. The decay curves presented in Figure 9 are not single-exponential, which is compatible with the presence of nonequivalent crystallographic sites of Eu3+ ions accordingly. The lifetimes values were calculated as the effective emission decay time by using Equation (1). The average lifetimes obtained for the sample single-doped by Eu3+ are equal to 0.93; 0.82, and 0.91 for concentrations of optically active ions at 0.5, 1.0, and 3.0 mol%, respectively.
The emission kinetic of Eu3+ ions strongly depends on the presence of Cu2+ ions, which effectively quenched the 5D0 level. The average lifetime obtained for co-doped materials is much shorter than that for Eu3+-doped materials, and the decay time values are estimated at about 0.33, 0.22, and 0.23 ms for the 0.5 mol% Eu3+/1 mol% Cu2+, 1 mol% Eu3+/1 mol% Cu2+, and 4 mol% Eu3+/0.5 mol% Cu2+ co-doped materials, respectively. The emission quenching of the Eu3+ ions may be interpreted as nonradiative energy transfer between the Eu3+ and Cu2+ ions. The efficiency of energy transfer was estimated by Equation (2) for pairs of materials: single-doped with Eu3+ ions and co-doped with Eu3+ and Cu2+ ions with the same concentration of Eu3+ ions. The calculated lifetimes of Eu3+ ions (donor) in the absence and presence of Cu2+ ions (acceptor) are used [40,44]. The results of energy transfer efficiency are presented in Table 2.
η Eu 3 + Cu 2 + = 1 ( τ Eu 3 + Cu 2 + τ Eu 3 + )
The obtained efficiency is equal to 65 and 73% for 0.5 mol% Eu2+/1 mol% Cu2+ and 1 mol% Eu2+/1 mol% Cu3+ co-doped HAp, respectively. The observed luminescence properties of europium ions in apatite lattice in the presence of copper ions are dominated by emission quenching of Eu3+ by Cu2+ ions. With an increase in Eu3+ concentration, the efficiency of quenching grew. This would suggest that the relatively huge probability of Eu3+ Cu2+ nonradiative energy transfer probably behaves by electric dipole interaction [40,44].
The simplified energy level diagram of Eu3+ and Cu2+ ions was proposed and is shown in Figure 10 in order to explain the quenching mechanism occurring in hydroxyapatite co-doped with Eu3+ and Cu2+ ions. When the materials were excited by 266 nm wavelength, a charge transfer transition O2− → Eu3+ occurred. Then, nonradiative relaxation to the 5D0 first excited state was performed. From this state, the energy can be relaxed in two ways: by radiative transition to the ground state of Eu3+ ions (7F0–6) or by energy transfer to the 2T2g energy level of Cu2+ and then nonradiative relaxation to the ground state of Cu2+ ions. These two manners of energy relaxation compete among themselves, and doping with Cu2+ ions causes Eu3+ emission quenching.

3.4. The EPR Spectra Analysis

The EPR spectroscopy is especially predisposed to identifying the structural properties of paramagnetic compounds. The unpaired electron interacts (couples) with the nuclear spin (I) to form a 2I + 1 line hyperfine structure centered on g and spaced with the distance quantified by the hyperfine coupling parameter A. The coupling between the nuclear and electron spins becomes stronger as the A parameter becomes larger. The combination of g and A parameters can be utilized to differentiate between electron environments of ion.
There are two distinct Ca coordination sites in the HAp unit cell, that is the Ca(1) site with the Ca2+ ion surrounded by 9 oxygen atoms from 6 PO43− groups and the Ca(2) site with the Ca2+ ion surrounded by 7 oxygen atoms from the 5 PO43− and 1 OH anions. The Ca2+ ions in both coordination sites can be replaced by Eu3+ and Cu2+ ions [45]. The EPR properties of trivalent europium (Eu3+) is relatively little because it is a non-Kramer ion, and its EPR spectrum should be silent because of the short spin-lattice relaxation time [46]. Therefore, in the EPR spectra recorded for the samples, only signals due to Cu2+ ions are observed.
The spectra recorded at room temperature and at 77 K (Figure 11) are anisotropic as a consequence of the Jahn–Teller effect operating for the d9 electron configuration of Cu2+ ions that leads to considerable departure from a regular symmetry of the coordination sphere. The spectra reveal a weakly resolved hyperfine interaction between the spins of unpaired electrons and copper nuclei (I = 3/2), which for powder spectrum suggest a large distance between paramagnetic centers (Cu2+ ions).
Spectral analysis revealed the existence of at least two different coordination environments for copper(II) ions. The EPR spectrum of the 1 mol% Eu3+/1 mol% Cu2+ co-doped HAp can be decomposed into two superimposed resonance signals due to two different Cu(II) coordination sites, hereinafter referred as “a” and “b”. This stays in line with the fact that there are two distinct Ca coordination sites in HAp in which Cu(II) can be doped. From the simulation of the spectrum recorded at 77 K, the estimated parameters are gz(a) = 2.41, gz(b) = 2.37, gy = 2.11, and gx = 2.08 with Az(a) = Az(b) = 110 G. However, the accuracy of these parameters is inevitably limited due to the fact that the Cu(II) signals are superimposed.
The observed EPR parameters contrast their counterparts determined for synthetic hydroxyapatite doped with Cu2+, in which also two different Cu(II) coordination sites were identified: gz = 2.485, gy = 2.17, gx = 2.08, and Az = 52 G for one coordination site and gz(a) = 2.420, gy = 2.17, gx = 2.08, and Az = 92 G for the second [47]. This difference in gz and Az parameters clearly stems from the structural divergence between the 1 mol% Eu3+/1 mol% Cu2+:HAp and SHA. At the same time, the gz and Az parameters for 1 mol% Eu3+/1 mol% Cu2+:HAp are similar to the ones reported for copper(II) ions bonded to lattice oxygens in montmorillonite((Cu(AlO)n(H2O)4-n)x): gz and Az in the ranges 2.37–2.41 and 100–140 G, respectively [48]. Therefore, the geometry of Cu(II) coordination sites in 1 mol% Eu3+/1 mol% Cu2+:HAp are expected to structurally resemble ((Cu(AlO)n(H2O)4-n)x).
Trends were found that enabled the Cu(II) EPR parameters to be correlated to the copper(II) ligands and the overall charge of the complexes [49,50,51,52]. According to these general trends, the gz and Az parameters for 1 mol% Eu3+/1 mol% Cu2+:HAp are characteristic of positively charged Cu–O complexes [49,50]. This fact indicates that, in the crystal lattice of the 1 mol% Eu3+/1 mol% Cu2+:HAp, the negative charge of the PO43− and OH anions are primarily neutralized by remained Ca2+ cations. Moreover, the observed difference between gz values found for two Cu coordination sites can be used to determine their possible assignment to Ca(1) and Ca(2). The increase in gz is associated with the rise in positive charges for the Cu–O complex. Hence, the higher value of gz(a) indicates that this Cu(II) ion is surrounded by a lower number of oxygen atoms, which are the primary carriers of a negative charge, and therefore should be labeled as Cu(II) ion doped into the Ca(2) site.

4. Conclusions

The pure crystal hydroxyapatite powder doped co-doped with Eu3+ and Cu2+ ions was successfully synthesized by a microwave-assisted hydrothermal method that was confirmed by the X-ray powder diffraction method. The nanometric size of the obtained materials was confirmed by Rietveld refinement and TEM techniques. In the absorption spectra, the transitions occurring in Eu3+ as well as Cu2+ ions were observed. In the emission spectra, the typical transition of Eu3+ ions (5D07FJ) were observed and the 5D07F2 transition is the most intense. The 5D07F0 transition consists of two lines, which means that the Eu3+ ions are localized in two independent crystallographic sites: in Ca(1) with C3 point symmetry and in Ca(2) with Cs symmetry. The emission decay times of Eu3+/Cu2+:HAp are much shorter than the decay times of Eu3+:HAp, which indicates that the Eu3+ emission is quenched by the Cu2+ ions. The simplified energy level diagram was proposed, and the quenching mechanism was explained. Based on the EPR measurement, the existence of at least two different coordinations surrounding copper(II) ions was detected.

Author Contributions

Conceptualization, K.S. and R.J.W.; methodology, K.S.; investigation, K.S., S.T., and A.L.; data curation, K.S. and S.T.; writing—original draft preparation, K.S.; writing—review and editing, K.S., S.T., A.L., A.W., and R.J.W.; visualization, K.S.; supervision, K.S. and R.J.W.; project administration, R.J.W.; funding acquisition, R.J.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Science Centre (NCN) within the projects “Preparation and characterisation of biocomposites based on nanoapatites for theranostics” (No. UMO-2015/19/B/ST5/01330) and “Elaboration and characteristics of biocomposites with anti-virulent and anti-bacterial properties against Pseudomonas aeruginosa” (No. UMO-2016/21/B/NZ6/01157).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We are grateful to E. Bukowska for the XRD measurements and to J. Komar for help with the emission spectra map measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pogosova, M.A.; González, L.V. Influence of anion substitution on the crystal structure and color properties of copper-doped strontium hydroxyapatite. Ceram. Int. 2018, 44, 20140–20147. [Google Scholar] [CrossRef]
  2. Zawisza, K.; Sobierajska, P.; Renaudin, G.; Nedelec, J.-M.; Wiglusz, R.J. Effects of crystalline growth on structural and luminescence properties of Ca(10−3x)Eu2x(PO4)6F2 nanoparticles fabricated by using a microwave driven hydrothermal process. CrystEngComm 2017, 19, 6936–6949. [Google Scholar] [CrossRef]
  3. Bal, Z.; Kaito, T.; Korkusuz, F.; Yoshikawa, H. Bone regeneration with hydroxyapatite-based biomaterials. Emergent Mater. 2019. [Google Scholar] [CrossRef]
  4. Pajor, K.; Pajchel, L.; Kolmas, J. Hydroxyapatite and fluorapatite in conservative dentistry and oral implantology-a review. Materials (Basel) 2019, 12, 2683. [Google Scholar] [CrossRef] [Green Version]
  5. 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]
  6. Szyszka, K.; Rewak-Soroczynska, J.; Dorotkiewicz-Jach, A.; Ledwa, K.A.; Piecuch, A.; Giersig, M.; Drulis-Kawa, Z.; Wiglusz, R.J. Structural modification of nanohydroxyapatite Ca10(PO4)6(OH)2 related to Eu3+ and Sr2+ ions doping and its spectroscopic and antimicrobial properties. J. Inorg. Biochem. 2020, 203, 110884. [Google Scholar] [CrossRef]
  7. Shanmugam, S.; Gopal, B. Copper substituted hydroxyapatite and fluorapatite: Synthesis, characterization and antimicrobial properties. Ceram. Int. 2014, 40, 15655–15662. [Google Scholar] [CrossRef]
  8. Wiglusz, R.J.; Drulis-Kawa, Z.; Pazik, R.; Zawisza, K.; Dorotkiewicz-Jach, A.; Roszkowiak, J.; Nedelec, J.M. Multifunctional lanthanide and silver ion co-doped nano-chlorapatites with combined spectroscopic and antimicrobial properties. Dalt. Trans. 2015, 44, 6918–6925. [Google Scholar] [CrossRef] [Green Version]
  9. Zawisza, K.; Sobierajska, P.; Nowak, N.; Kedziora, A.; Korzekwa, K.; Pozniak, B.; Tikhomirov, M.; Miller, J.; Mrowczynska, L.; Wiglusz, R.J. Preparation and preliminary evaluation of bio-nanocomposites based on hydroxyapatites with antibacterial properties against anaerobic bacteria. Mater. Sci. Eng. C 2020, 106, 110295. [Google Scholar] [CrossRef]
  10. 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 2018, 90, 308–313. [Google Scholar] [CrossRef] [PubMed]
  11. Stanić, V.; Dimitrijević, S.; Antić-Stanković, J.; Mitrić, M.; Jokić, B.; Plećaš, I.B.; Raičević, S. Synthesis, characterization and antimicrobial activity of copper and zinc-doped hydroxyapatite nanopowders. Appl. Surf. Sci. 2010, 256, 6083–6089. [Google Scholar] [CrossRef]
  12. Kim, T.N.; Feng, Q.L.; Kim, J.O.; Wu, J.; Wang, H.; Chen, G.C.; Cui, F.Z. Antimicrobial effects of metal ions (Ag+, Cu2+, Zn2+) in hydroxyapatite. J. Mater. Sci. Mater. Med. 1998, 9, 129–134. [Google Scholar] [CrossRef]
  13. Kolmas, J.; Groszyk, E.; Kwiatkowska-Rózycka, D. Substituted hydroxyapatites with antibacterial properties. Biomed Res. Int. 2014, ID 178123, 1–15. [Google Scholar] [CrossRef]
  14. Huang, Y.; Hao, M.; Nian, X.; Qiao, H.; Zhang, X.; Zhang, X.; Song, G.; Guo, J.; Pang, X.; Zhang, H. Strontium and copper co-substituted hydroxyapatite-based coatings with improved antibacterial activity and cytocompatibility fabricated by electrodeposition. Ceram. Int. 2016, 42, 11876–11888. [Google Scholar] [CrossRef]
  15. Pogosova, M.A.; Kasin, P.E.; Tretyakov, Y.D.; Jansen, M. Synthesis, structural features, and color of calcium-yttrium hydroxyapatite with copper ions in hexagonal channels. Russ. J. Inorg. Chem. 2013, 58, 381–386. [Google Scholar] [CrossRef]
  16. Kazin, P.E.; Zykin, M.A.; Zubavichus, Y.V.; Magdysyuk, O.V.; Dinnebier, R.E.; Jansen, M. Identification of the chromophore in the apatite pigment [Sr10(PO4)6(CuxOH1-x-y)2]: Linear OCuO- featuring a resonance raman effect, an extreme magnetic anisotropy, and slow spin relaxation. Chem. A Eur. J. 2014, 20, 165–178. [Google Scholar] [CrossRef] [PubMed]
  17. Pogosova, M.A.; Eliseev, A.A.; Kazin, P.E.; Azarmi, F. Synthesis, structure, luminescence, and color features of the Eu- and Cu-doped calcium apatite. Dye. Pigment. 2017, 141, 209–216. [Google Scholar] [CrossRef]
  18. Kazin, P.E.; Zykin, M.A.; Schnelle, W.; Felser, C.; Jansen, M. Rich diversity of single-ion magnet features in the linearOCuIIIO- ion confined in the hexagonal channels of alkaline-earth phosphate apatites. Chem. Commun. 2014, 50, 9325–9328. [Google Scholar] [CrossRef]
  19. Zawisza, K.; Strzep, A.; Wiglusz, R.J. Influence of annealing temperature on the spectroscopic properties of hydroxyapatite analogues doped with Eu3+. New J. Chem. 2017, 41, 9990–9999. [Google Scholar] [CrossRef]
  20. Syamchand, S.S.; Sony, G. Europium enabled luminescent nanoparticles for biomedical applications. J. Lumin. 2015, 165, 190–215. [Google Scholar] [CrossRef]
  21. Hassan, M.N.; Mahmoud, M.M.; El-Fattah, A.A.; Kandil, S. Microwave-assisted preparation of Nano-hydroxyapatite for bone substitutes. Ceram. Int. 2016, 42, 3725–3744. [Google Scholar] [CrossRef]
  22. Chen, J.; Liu, J.; Deng, H.; Yao, S.; Wang, Y. Regulatory synthesis and characterization of hydroxyapatite nanocrystals by a microwave-assisted hydrothermal method. Ceram. Int. 2020, 46, 2185–2193. [Google Scholar] [CrossRef]
  23. 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] [PubMed] [Green Version]
  24. Mondal, S.; Hoang, G.; Manivasagan, P.; Kim, H.; Oh, J. Nanostructured hollow hydroxyapatite fabrication by carbon templating for enhanced drug delivery and biomedical applications. Ceram. Int. 2019, 45, 17081–17093. [Google Scholar] [CrossRef]
  25. Targonska, S.; Rewak-Soroczynska, J.; Piecuch, A.; Paluch, E.; Szymanski, D.; Wiglusz, R.J. Preparation of a New Biocomposite Designed for Cartilage Grafting with Antibiofilm Activity. ACS Omega 2020, 5, 24546–24557. [Google Scholar] [CrossRef]
  26. Targonska, S.; Wiglusz, R.J. Investigation of Physicochemical Properties of the Structurally Modified Nanosized Silicate-Substituted Hydroxyapatite Co-Doped with Eu3+ and Sr2+ Ions. Nanomaterials 2021, 11. [Google Scholar]
  27. Escudero, A.; Calvo, M.E.; Rivera-Fernández, S.; de la Fuente, J.M.; Ocaña, M. Microwave-Assisted Synthesis of Biocompatible Europium-Doped Calcium Hydroxyapatite and Fluoroapatite Luminescent Nanospindles Functionalized with Poly(acrylic acid). Langmuir 2013, 29, 1985–1994. [Google Scholar] [CrossRef] [Green Version]
  28. Wang, C.; Jeong, K.-J.; Kim, J.; Kang, S.W.; Kang, J.; Han, I.H.; Lee, I.-W.; Oh, S.-J.; Lee, J. Emission-tunable probes using terbium(III)-doped self-activated luminescent hydroxyapatite for in vitro bioimaging. J. Colloid Interface Sci. 2021, 581, 21–30. [Google Scholar] [CrossRef]
  29. Muresan, L.E.; Perhaita, I.; Prodan, D.; Borodi, G. Studies on terbium doped apatite phosphors prepared by precipitation under microwave conditions. J. Alloys Compd. 2018, 755, 135–146. [Google Scholar] [CrossRef]
  30. Ma, B.; Zhang, S.; Qiu, J.; Li, J.; Sang, Y.; Xia, H.; Jiang, H.; Claverie, J.; Liu, H. Eu/Tb codoped spindle-shaped fluorinated hydroxyapatite nanoparticles for dual-color cell imaging. Nanoscale 2016, 8, 11580–11587. [Google Scholar] [CrossRef]
  31. Targonska, S.; Szyszka, K.; Rewak-Soroczynska, J.; Wiglusz, R.J. A new approach to spectroscopic and structural studies of the nano-sized silicate-substituted hydroxyapatite doped with Eu3+ ions. Dalt. Trans. 2019. [Google Scholar] [CrossRef] [PubMed]
  32. Priyadarshini, B.; Vijayalakshmi, U. Development of cerium and silicon co-doped hydroxyapatite nanopowder and its in vitro biological studies for bone regeneration applications. Adv. Powder Technol. 2018, 29, 2792–2803. [Google Scholar] [CrossRef]
  33. Yoder, C.H.; Havlusch, M.D.; Dudrick, R.N.; Schermerhorn, J.T.; Tran, L.K.; Deymier, A.C. The synthesis of phosphate and vanadate apatites using an aqueous one-step method. Polyhedron 2017, 127, 403–409. [Google Scholar] [CrossRef] [Green Version]
  34. Nakagawa, D.; Nakamura, M.; Nagai, S.; Aizawa, M. Fabrications of boron-containing apatite ceramics via ultrasonic spray-pyrolysis route and their responses to immunocytes. J. Mater. Sci. Mater. Med. 2020, 31. [Google Scholar] [CrossRef]
  35. 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]
  36. Rietveld, H.M. A profile refinement method for nuclear and magnetic structures. J. Appl. Crystallogr. 1969, 2, 65–71. [Google Scholar] [CrossRef]
  37. Lutterotti, L.; Matthies, S.; Wenk, H.-R. MAUD: A friendly Java program for Material Analysis Using Diffraction. IUCr Newsl. CPD 1999, 21, 14–15. [Google Scholar] [CrossRef] [Green Version]
  38. Balan, E.; Delattre, S.; Roche, D.; Segalen, L.; Morin, G.; Guillaumet, M.; Blanchard, M.; Lazzeri, M.; Brouder, C.; Salje, E.K.H. Line-broadening effects in the powder infrared spectrum of apatite. Phys Chem Miner. 2011, 38, 111–122. [Google Scholar] [CrossRef]
  39. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. 1976, A32, 751–767. [Google Scholar] [CrossRef]
  40. Jiménez, J.A. Photoluminescence of Eu3+-doped glasses with Cu2+ impurities. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 145, 482–486. [Google Scholar] [CrossRef]
  41. Lakshmi Reddy, S.; Endo, T.; Siva Reddy, G. Electronic (Absorption) Spectra of 3d Transition Metal Complexes. In Advanced Aspects of Spectroscopy; Farrukh, M.A., Ed.; IntechOpen: London, UK, 2012; pp. 3–48. [Google Scholar]
  42. Orihashi, T.; Nakamura, T.; Adachi, S. Synthesis and Unique Photoluminescence Properties of Eu2Ti2O7 and Eu2TiO5. J. Am. Ceram. Soc. 2016, 99, 3039–3046. [Google Scholar] [CrossRef]
  43. Wang, S.; Wei, W.; Zhou, X.; Li, Y.; Hua, G.; Li, Z.; Wang, D.; Mao, Z.; Zhang, Z.; Ying, Z. Comparative study of the luminescence properties of Ca2+xLa8-x(SiO4)6-x(PO4)xO2:Eu3+ (x = 0, 2) red phosphors. J. Lumin. 2020, 221, 1170743. [Google Scholar] [CrossRef]
  44. Jiménez, J.A. Eu3+ amidst ionic copper in glass: Enhancement through energy transfer from Cu+, or quenching by Cu2+? Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2017, 173, 979–985. [Google Scholar] [CrossRef]
  45. Cawthray, J.F.; Creagh, A.L.; Haynes, C.A.; Orvig, C. Ion exchange in hydroxyapatite with lanthanides. Inorg. Chem. 2015, 54, 1440–1445. [Google Scholar] [CrossRef]
  46. Yong, G.; Chunshan, S. The spectral properties of BaB8O13:Eu, Tb phosphors. J. Phys. Chem. Solids 1996, 57, 1303–1306. [Google Scholar] [CrossRef]
  47. Sutter, B.; Wasowicz, T.; Howard, T.; Hossner, L.R.; Ming, D.W. Characterization of Iron, Manganese, and Copper Synthetic Hydroxyapatites by Electron Paramagnetic Resonance Spectroscopy. Soil Sci. Soc. Am. J. 2002, 66, 1359–1366. [Google Scholar] [CrossRef]
  48. Bahranowski, K.; Dula, R.; Labanowska, M.; Serwicka, E.M. ESR study of Cu centers supported on Al-, Ti-, and Zr-pillared montmorillonite clays. Appl. Spectrosc. 1996, 50, 1439–1445. [Google Scholar] [CrossRef]
  49. Peisach, J.; Blumberg, W. Analysis of EPR Copper structural implications derived from the analysis of EPR spectra of natural and artificial Cu-proteins. Arch Biochem Biophys 1974, 165, 691–708. [Google Scholar] [CrossRef]
  50. Kumar, S.; Sharma, R.P.; Venugopalan, P.; Witwicki, M.; Ferretti, V. Synthesis, characterization, single crystal X-ray structure, EPR and theoretical studies of a new hybrid inorganic-organic compound [Cu(Hdien)2(H2O)2](pnb)4·4H2O and its structural comparison with related [Cu(en)2(H2O)2](pnb)2. J. Mol. Struct. 2016, 1123, 124–132. [Google Scholar] [CrossRef]
  51. Maślewski, P.; Wyrzykowski, D.; Witwicki, M.; Dołęga, A. Histaminol and Its Complexes with Copper(II) – Studies in Solid State and Solution. Eur. J. Inorg. Chem. 2018, 2018, 1399–1408. [Google Scholar] [CrossRef]
  52. Kumar, S.; Pal Sharma, R.; Venugopalan, P.; Gondil, V.S.; Chhibber, S.; Aree, T.; Witwicki, M.; Ferretti, V. Hybrid inorganic-organic complexes: Synthesis, spectroscopic characterization, single crystal X-ray structure determination and antimicrobial activities of three copper(II)-diethylenetriamine-p-nitrobenzoate complexes. Inorganica Chim. Acta 2018, 469, 288–297. [Google Scholar] [CrossRef]
Figure 1. X-ray powder diffraction patterns of the Eu3+-doped and Eu3+/Cu2+ co-doped Ca10(PO4)6(OH)2.
Figure 1. X-ray powder diffraction patterns of the Eu3+-doped and Eu3+/Cu2+ co-doped Ca10(PO4)6(OH)2.
Nanomaterials 11 00464 g001
Figure 2. Representative results for the 0.5 mol% Eu3+/1 mol% Cu2+:Ca10(PO4)6(OH)2:, Rietveld analysis (red—fitted diffraction, green—differential pattern, and blue column—reference phase peak position).
Figure 2. Representative results for the 0.5 mol% Eu3+/1 mol% Cu2+:Ca10(PO4)6(OH)2:, Rietveld analysis (red—fitted diffraction, green—differential pattern, and blue column—reference phase peak position).
Nanomaterials 11 00464 g002
Figure 3. Representative TEM images (ac) and particle size distribution (d) of the 1 mol% Eu3+:Ca10(PO4)6(OH)2.
Figure 3. Representative TEM images (ac) and particle size distribution (d) of the 1 mol% Eu3+:Ca10(PO4)6(OH)2.
Nanomaterials 11 00464 g003
Figure 4. Infrared spectra of the Eu3+-doped and Eu3+/Cu2+ co-doped Ca10(PO4)6(OH)2.
Figure 4. Infrared spectra of the Eu3+-doped and Eu3+/Cu2+ co-doped Ca10(PO4)6(OH)2.
Nanomaterials 11 00464 g004
Figure 5. The absorption spectra of the Eu3+-doped and Eu3+/Cu2+-co-doped Ca10(PO4)6(OH)2.
Figure 5. The absorption spectra of the Eu3+-doped and Eu3+/Cu2+-co-doped Ca10(PO4)6(OH)2.
Nanomaterials 11 00464 g005
Figure 6. The excitation spectra of the 1 mol% Eu3+ and 1 mol% Eu3+/1 mol% Cu2+-co-doped Ca10(PO4)6(OH)2.
Figure 6. The excitation spectra of the 1 mol% Eu3+ and 1 mol% Eu3+/1 mol% Cu2+-co-doped Ca10(PO4)6(OH)2.
Nanomaterials 11 00464 g006
Figure 7. The emission spectra of the Eu3+-doped and Eu3+/Cu2+ co-doped Ca10(PO4)6(OH)2.
Figure 7. The emission spectra of the Eu3+-doped and Eu3+/Cu2+ co-doped Ca10(PO4)6(OH)2.
Nanomaterials 11 00464 g007
Figure 8. Representative emission spectra map of the 1 mol% Eu3+/1 mol% Cu2+:Ca10(PO4)6(OH)2.
Figure 8. Representative emission spectra map of the 1 mol% Eu3+/1 mol% Cu2+:Ca10(PO4)6(OH)2.
Nanomaterials 11 00464 g008
Figure 9. Decay times of the Eu3+-doped and Eu3+/Cu2+ co-doped Ca10(PO4)6(OH)2.
Figure 9. Decay times of the Eu3+-doped and Eu3+/Cu2+ co-doped Ca10(PO4)6(OH)2.
Nanomaterials 11 00464 g009
Figure 10. Simplified energy level scheme of Eu3+ and Cu2+ explaining quenching of Eu3+ ion emission.
Figure 10. Simplified energy level scheme of Eu3+ and Cu2+ explaining quenching of Eu3+ ion emission.
Nanomaterials 11 00464 g010
Figure 11. Experimental and simulated electron paramagnetic resonance (EPR) spectra of the 1 mol% Eu3+/1 mol% Cu2+:Ca10(PO4)6(OH)2.
Figure 11. Experimental and simulated electron paramagnetic resonance (EPR) spectra of the 1 mol% Eu3+/1 mol% Cu2+:Ca10(PO4)6(OH)2.
Nanomaterials 11 00464 g011
Table 1. Unit cell parameters (a and c), cell volume (V), grain size, as well as refine factor (RW) for the Eu3+-doped and Eu3+/Cu2+ co-doped Ca10(PO4)6(OH)2.
Table 1. Unit cell parameters (a and c), cell volume (V), grain size, as well as refine factor (RW) for the Eu3+-doped and Eu3+/Cu2+ co-doped Ca10(PO4)6(OH)2.
Samplea (Å)c (Å)V3)Size (nm)Rw (%)
single crystal9.424(4)6.879(4)529.09(44)
doped with x mol% Eu3+
0.5 mol% Eu3+9.4139(6)6.8905(0)528.83(54)55.6(2)3.0
1 mol% Eu3+9.4259(6)6.8881(3)530.19(10)45.5(9)2.8
3 mol% Eu3+9.4276(8)6.8891(5)530.26(80)38.0(1)3.3
co-doped with x mol% Eu3+ and y mol% Cu2+
0.5 mol% Eu3+/1 mol% Cu2+9.4300(2)6.8875(1)530.41(48)42.0(2)3.1
1 mol% Eu3+/1 mol% Cu2+9.4264(5)6.8858(9)529.87(91)41.2(4)2.7
4 mol% Eu3+/0.5 mol% Cu2+9.4288(2)6.8893(5)530.41(84)58.2(0)3.2
Table 2. The average lifetime of Eu3+-doped (τEu), Eu3+/1 mol% Cu2+EuCu) co-doped Ca10(PO4)6(OH)2 and energy transfer efficiency (ηEuCu).
Table 2. The average lifetime of Eu3+-doped (τEu), Eu3+/1 mol% Cu2+EuCu) co-doped Ca10(PO4)6(OH)2 and energy transfer efficiency (ηEuCu).
τEu (ms)τEuCu (ms)ηEuCu (%)
0.5 mol% Eu3+0.930.3365
1 mol% Eu3+0.820.2273
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Szyszka, K.; Targońska, S.; Lewińska, A.; Watras, A.; Wiglusz, R.J. Quenching of the Eu3+ Luminescence by Cu2+ Ions in the Nanosized Hydroxyapatite Designed for Future Bio-Detection. Nanomaterials 2021, 11, 464. https://doi.org/10.3390/nano11020464

AMA Style

Szyszka K, Targońska S, Lewińska A, Watras A, Wiglusz RJ. Quenching of the Eu3+ Luminescence by Cu2+ Ions in the Nanosized Hydroxyapatite Designed for Future Bio-Detection. Nanomaterials. 2021; 11(2):464. https://doi.org/10.3390/nano11020464

Chicago/Turabian Style

Szyszka, Katarzyna, Sara Targońska, Agnieszka Lewińska, Adam Watras, and Rafal J. Wiglusz. 2021. "Quenching of the Eu3+ Luminescence by Cu2+ Ions in the Nanosized Hydroxyapatite Designed for Future Bio-Detection" Nanomaterials 11, no. 2: 464. https://doi.org/10.3390/nano11020464

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