Next Article in Journal
Metal-Free Aerobic C–N Bond Formation of Styrene and Arylamines via Photoactivated Electron Donor–Acceptor Complexation
Next Article in Special Issue
Excitation-Controlled Host–Guest Multicolor Luminescence in Lanthanide-Doped Calcium Zirconate for Information Encryption
Previous Article in Journal
Rhenium(I) Block Copolymers Based on Polyvinylpyrrolidone: A Successful Strategy to Water-Solubility and Biocompatibility
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Series of Red-Light Phosphor Ca9−xZnxGd0.9(PO4)7:0.1Eu3+ (x = 0–1)

by
Yury Yu. Dikhtyar
1,
Dmitry A. Spassky
2,3,*,
Vladimir A. Morozov
1,
Sergey N. Polyakov
4,
Valerya D. Romanova
1,
Sergey Yu. Stefanovich
1,
Dina V. Deyneko
1,5,
Oksana V. Baryshnikova
1,
Ivan V. Nikiforov
1 and
Bogan I. Lazoryak
1
1
Chemistry Department, Lomonosov Moscow State University, 119991 Moscow, Russia
2
Skobeltsyn Institute of Nuclear Physics, Lomonosov Moscow State University, 119991 Moscow, Russia
3
Institute of Physics, University of Tartu, W. Ostwald str. 1, 50411 Tartu, Estonia
4
Technological Institute for Superhard and Novel Carbon Materials, Troitsk, 108840 Moscow, Russia
5
Laboratory of Arctic Mineralogy and Material Sciences, Kola Science Centre, Russian Academy of Sciences, 14 Fersman str.,184209 Apatity, Russia
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(1), 352; https://doi.org/10.3390/molecules28010352
Submission received: 1 December 2022 / Revised: 26 December 2022 / Accepted: 28 December 2022 / Published: 1 January 2023

Abstract

:
In this study, a new series of phosphors, Ca9−xZnxGd0.9(PO4)7:0.1Eu3+ (x = 0.00–1.00, step dx 0.05), was synthesized, consisting of centro- and non-centrosymmetric phases with β-Ca3(PO4)2-type structure. Crystal structures with space groups R3c (0.00 ≤ x < 0.35) and R 3 ¯ c (x > 0.8) were determined using X-ray powder diffraction and the method of optical second harmonic generation. In the region 0.35 ≤ x ≤ 0.75, phases R3c and R 3 ¯ c were present simultaneously. Refinement of the Ca8ZnGd(PO4)7 crystal structure with the Rietveld method showed that 71% of Gd3+ ions are in M3 sites and 29% are in M1 sites. A luminescent spectroscopy study of Ca9−xZnxGd0.9(PO4)7:0.1Eu3+ indicated the energy transfer from the crystalline host to the Gd3+ and Eu3+ luminescent centers. The maximum Eu3+ luminescence intensity corresponds to the composition with x = 1.

1. Introduction

Structural type β-Ca3(PO4)2 (space group (SG): R3c, Z = 21) [1] is a promising base for the development of materials with various interesting properties: luminescent [2], laser [3], non-linear optical [4,5,6], ferroelectric [4,5,6], antiferroelectric [7,8], catalytic [9], bioregenerative [10] and antibacterial [11]. In particular, these compounds can be expected to be of practical interest as narrow-band red light phosphors for the creation of white-light emission diodes (WLEDs), or as bone substitute material with ability to use imaging techniques.
The creation of new synthetic bone substitutes is an important task due to the rapid growth of the aging population and the increasing number of people with diseases of the musculoskeletal system [12]. Ceramics, based on calcium phosphates, are the most common material for filling bone defects [13]. Of all possible calcium orthophosphates (hydroxyapatite Ca10(PO4)6(OH)2 [14], β-tricalcium phosphate β-Ca3(PO4)2 [15], a-tricalcium phosphate a-Ca3(PO4)2 [16], brushite CaHPO4·2H2O [17], octacalcium phosphate Ca8(HPO4)2(PO4)4·5H2O [18]), β-Ca3(PO4)2 is one of the most applicable for bone tissue restoration [19].
The development of a highly efficient red phosphor is an urgent task because commercially available WLED based on (Ga,In)N (465 nm) blue-chip [20] and bright-yellow phosphor Y3Al2(AlO4)3:Ce3+ (YAG:Ce3+) [21,22] suffers from several disadvantages: low color rendering index (CRI) [23,24,25,26], high correlated color temperature [27], bad influence on a person’s psychological state and health of the eyes [28,29]. Using the UV-chip [30] and varying blue, red [31,32,33] and green phosphors, a WLED with the necessary characteristics can be obtained [34,35].
The unique structure of β-Ca3(PO4)2 has wide possibilities for cationic substitution. For example, incorporation of Eu3+ in β-Ca3(PO4)2 makes it possible to obtain red-light phosphors, while co-doping of Eu3+ and Gd3+ allows for the acquirement of even more superior luminescent properties. In [36] it was shown that the Ca9Gd0.1Eu0.9(PO4)7 phosphor emitted red light by 4.13 times brighter than the commercially available Y2O3:Eu3+. In [37,38], the authors found that sample Ca8MgGd(PO4)7:Eu3+ had the highest luminescent intensity in comparison with Ca8MgY(PO4)7:Eu3+ and Ca8MgLa(PO4)7:Eu3+. This may be related to the sensibilization effect of Gd3+ in the energy transfer processes to Eu3+. Thus, the simultaneous incorporation of Eu3+ and Gd3+ can improve the luminescent characteristics of material (due to energy transfer processes from the matrix to Gd3+ and further to Eu3+-centers), and also can act as a contrast agent for MRI (magnetic resonance imaging) and X-ray dual imaging, which consists of combining two radiographs acquired at two different lanthanides [39].
Further improvement of luminescent properties is possible due to isovalent cation substitutions within the β-Ca3(PO4)2 host. This can also lead to a change in SG (supporting information in [40]), which should be taken into account. Sometimes, the authors did not define the crystal structure of Ca8MGd(PO4)7:Eu3+ (M2+ = Zn, Mg, Cd) with SG R3c correctly. Substitution Ca2+ → Zn2+ in the M5 site of β-Ca3(PO4)2 leads to the change in SG R3cR 3 ¯ c, the improvement of materials’ luminescent properties (the explanation of this phenomena is provided in detail in [2,41]) and antibacterial characteristics [42] (Figure 1). Given the highest luminescent characteristics of phosphor with M = Zn2+ in Ca8MEu(PO4)7 [41] (M = Ca2+, Mg2+, Zn2+, Cd2+), it can be assumed that Ca8ZnGd(PO4)7:Eu3+ will show improved luminescent properties.
In the present study, a number of solid solutions with the general formula Ca9−xZnxGd0.9Eu0.1(PO4)7 (x = 0.00–1.00, dx = 0.05) are synthesized and examined for the first time. The following questions were in the focus of the study: (1) the boundary of the two-phase region with the change of SG R3cR 3 ¯ c at gradual substitution Ca2+ → Zn2+; (2) the distribution of cations in the Ca8ZnGd(PO4)7 structure; (3) the modification of luminescence properties, in particular the energy transfer to Eu3+ with crystal compositions.

2. Results and Discussions

2.1. Elemental Composition and Preliminary Characterization

The quantitative ratio of elements was determined by EDX analysis. The results for samples with x = 0.35, 0.50, 0.75 and 1.00 in Ca9−xZnxGd0.9Eu0.1(PO4)7 series show the insignificant deviation from the theoretical composition. Table 1 summarizes the results of the EDX analysis.

2.2. Scanning Electron Microscopy

Figure 2 shows the images obtained with the SEM method for x = 0.00, 0.35, 0.75 and 1.00 in the Ca9−xZnxGd0.9Eu0.1(PO4)7 series. The shape of the particles becomes sharper, and the particles form larger agglomerates under Ca2+→ Zn2+ substitution in Ca9xZnxGd0.9Eu0.1(PO4)7. This is correlated with the transition from the non-centrosymmetric to centrosymmetric (R3c → R 3 ¯ c) state.

2.3. PXRD Analysis

A number and peaks position of synthesized solid solutions in Ca9−xZnxGd(PO4)7:0.1Eu3+ PXRD patterns related to similar compounds with β-Ca3(PO4)2-type structure [2]. The absence of impurity phases shows that Gd3+ (rVI = 0.94 Å), Eu3+ (rVI = 0.95 Å) and Zn2+ (rVI = 0.74 Å [43]) ions are completely incorporated into the β-Ca3(PO4)2-type structure.
The unit cells parameters decrease with increasing of Zn2+ concentration because the ionic radius of Zn2+ (rVI = 0.74 Å) is smaller than that of Ca2+ (rVI = 1.00 Å) (Table S1). However, this decrease in parameters is nonlinear. Figure 3a shows that the slope of the curve in the region of 0.00 ≤ x < 0.30 is less than in the region of 0.30 ≤ x ≤ 0.80, and there is a sharp jump in the change of parameters for the composition x = 0.50. Such comprehensive behavior of unit cell parameters on Zn2+ concentration can be explained by a change in the SG R3cR 3 ¯ c. In a routine laboratory experiment, PXRD patterns of compounds with these SGs are indistinguishable [44,45].
In the region x = 0.35–0.75, a decrease in the crystalline size and an increase in the FWHM in comparison with x = 1 is observed. This circumstance may also indicate the coexistence of two phases—R3c and R 3 ¯ c—further confirmed by the SHG method.

2.4. SHG Study

The above-defined limitations on the existence of regions with different symmetries are consistent with the SHG data (Table S1). As Zn2+ ions concentration rises from 0 to 0.35, there is a slight decrease in SHG signal (Figure 3b). More rapid and nonmonotonic decrease of SHG is observed in an interphase region with 0.35 < x < 0.75, where centro- and non-centrosymmetry fragments of whitlockite-like structures are mixed. Very small SHG activity at 0.75 < x ≤ 1.00 corresponds to centrosymmetric phase (SG R 3 ¯ c) distorted with defects near the smaller boundary of x. For x > 0.75, the SHG signal is null or at the background level in accordance with macroscopic center of symmetry in this phase.

2.5. Crystal Structure Refinement of Ca8ZnGd(PO4)7

Compounds with a β-Ca3(PO4)2-type structure have extended possibilities for cationic substitution. There are six crystallographic positions with different sizes M1–M5 and M6 for Ca2+ in β-Ca3(PO4)2. The occupancy of M4 site can vary from 0 to 1, while M6 is always vacant. Large numbers of positions and vacancies suggests a wide opportunity for iso- and heterovalent cationic substitution, including the lanthanoids Ln3+. These substitutions may lead to changing of SG from polar to non-polar (Supplementary information of [40]).
Atomic coordinates of phosphate Ca8ZnLa(PO4)7 (SG R 3 ¯ c) [46] were used as an initial model for the structural refinement of Ca8ZnGd(PO4)7. The M1 and M3 sites (36f) are jointly occupied by La3+ and Ca2+; M5 (6b) is occupied by Zn2+ in Ca8ZnLa(PO4)7. There is no M2 site in this structure with R 3 ¯ c SG, since M2 is equivalent to M1 (M1 = M2). In centrosymmetric model with R 3 ¯ c SG, the name of the M3 site was left the same, as for the polar model with R3c SG of β-TCP-type structure. There are two phosphorus atoms in 12c and 36f Wyckoff positions. There is one oxygen atom in 12c Wyckoff position, and the other five are in 36f positions.
At the first step of the refinement, the f-curves for Ca2+ (in M1 and M3 sites) and Zn2+ (in M5) were used to form the determination of the atoms’ positions. This analysis (Table 2) shows that the Gd3+ ions are distributed between the positions M1, M3 (exceeding the maximum occupancy—1 for M1 and 0.5 for M3), while the M5 position is completely occupied by Zn2+ ions.
The M3 and P1 sites must be in special Wyckoff positions (18d) (0.5, 0, 0) and (6a) (0, 0, 0.25), respectively, in Ca8ZnGd(PO4)7 (SG R 3 ¯ c). The structure refinement of this model led to large parameters of atomic displacement, Uiso. = 0.162(2) Å2 for Ca2+/Gd3+ in the M3 site, and Uiso. = 0.173(4) Å2 for P1. For this reason, the refinement of the Ca8ZnGd(PO4)7 structure was performed with a shift of the phosphorus atom P1 from a special (6a) position to a half-occupied (12c) position. The Ca2+ in M3 was shifted from a special (18d) position to a half-occupied (36f) site. Moreover, the positions of Ca2+ and Gd3+ in M1 and M3 sites were additionally split. This led to a significant decrease in M1 and M3 Uiso.
After refinement, there is a good agreement between the calculated and experimental X-ray diffraction patterns (Figure 4) with acceptable R-factors (Table 2). Fractional atomic coordinates, site symmetry, isotropic displacement of atomic parameters and site occupation for Ca8ZnGd(PO4)7 are shown in Table S2. The main interatomic distances are shown in Table S3. The distribution of Gd3+ ions in Ca8ZnGd(PO4)7 over crystal sites was found to be 71% in M3, and 29% in M1. In Ca8MgGd(PO4)7 [47], the distribution of Gd3+ ions was 77% in M3, and 23% in M1 sites. Thus, the M3 site is a preferable location for relatively big Gd3+ (rVI = 0.94 Å) ions.
The average length of the M-O bonds in the M1O9 and M3O8 polyhedra are d<M1-O> = 2.347 Å and d<M3-O> = 2.526 Å, respectively. At the same time, the deviations of M-O distances from the average values of d<M1-O> and d<M3-O> are significant and indicate a large distortion of the polyhedra. Calculated distortion indexes (DI [48]) for M1O9 and M3O8 polyhedra are DIM1O = 0.0362 and DIM3O = 0.0563. These values were calculated according to the Equation (1)
D I = 1 n i = 1 n d i d
where n—number of bonds, di—length of a bond and <d>—average bond length for polyhedra. In Ca9Gd(PO4)7, DI’s for M1O8, M2O8 and M3O8 are DIM1O = 0.0486, DIM2O = 0.0234 and DIM3O = 0.0500 (M1 and M2 sites in SG R 3 ¯ c are equivalent) [49]. It indicates that M3 site (77% Gd3+) in Ca8ZnGd(PO4)7 (SG R 3 ¯ c) is more distorted than M3 site in Ca9Gd(PO4)7 (SG R3c).

2.6. Luminescent Properties of Ca9−xZnxGd(PO4)7:0.1Eu3+

The luminescence spectra of Ca9−xZnxGd0.9Eu0.1(PO4)7 (x = 0, 1) compounds are presented in Figure 5a,b at 6 and 300 K. The spectra consist of a set of narrow emission lines in the region of 305–317 nm and 575–720 nm, related to 4f-4f transitions in Gd3+ and Eu3+, respectively. The features of the structure of Eu3+ emission change with the incorporation of Zn2+. It can be shown for the 5D0–7F0 transition (Figure 6), which is not split by the crystal field and sensitive to the number of nonequivalent Eu3+ positions. Two peaks can be observed for the sample with x = 0, while only one for x = 1. The number of peaks correlate with the decrease of the number of cationic positions. However, presence of slightly different crystallographic sites with similar crystallographic properties may also result in a single peak consisting of several superimposed peaks. For x = 0 (SG R3c), there are three cationic positions—M1, M2 and M3, while for x = 1 (SG R 3 ¯ c) there are only two cationic positions—M1 and M3. Analysis of the emission spectra in the 305–317 nm region demonstrates that the number of peaks related to Gd3+ also depends on the crystal composition. Presence of additional features such as the shoulder at 310.6 nm and peak at 314 nm depends on the Zn2+ content in the sample. This is especially demonstrative when the temperature drops to 6 K (Figure 5b)—the peak at 314 nm becomes more prominent for the compound with x = 0.
The relative intensities of Gd3+ and Eu3+ emissions depend on the Zn2+ content in the sample under VUV excitation (163 nm). The given wavelength is tentatively related to the fundamental absorption region [2], and observed modifications are connected with the features of energy transfer to competitive emission centers. It was found that in the sample with x = 0, the Gd3+ luminescence band dominated in the spectrum, while for x = 1, the Eu3+ luminescence band intensity increased (Figure 5). Therefore, the energy transfer from the host to Eu3+ is more efficient in the sample with Zn2+.
PLE spectra of Eu3+ and Gd3+ emissions are presented in Figure 7. In the excitation spectra of Eu3+ a set of narrow lines in the region of 320–500 nm is connected with 4f-4f Eu3+ transitions, while the broad band peaking at ~245 nm to charge transfer transitions from the valence band (VB) to 4f Eu states. A narrow excitation band at 273 nm is related to 8S7/26IJ transitions in Gd3+, thus indicating the energy transfer from Gd3+ to Eu3+. The scheme of Eu3+ and Gd3+ energy levels position relative to the top of the valence and bottom of the conduction bands is presented in Figure 8. The position of Eu3+ and Gd3+ 4f ground states position were taken from [2], while 4f excited states were taken from a Dieke diagram [50]. In the excitation spectra of Gd3+ emissions, a set of narrow lines is observed at 246, 253 and 273 nm, which are connected with 8S7/26DJ and 8S7/26IJ Gd3+ transitions. A very weak broad band can be found in the region of 210–250 nm. The position of this band coincides with the position of the charge-transfer band (CTB) in the excitation spectra of Eu3+ and shows that the energy transfer from Eu3+ to Gd3+ is possible as well; however, its efficiency is low. Similarity of the PLE of Ca9Gd0.9Eu0.1(PO4)7 and Ca8ZnGd0.9Eu0.1(PO4)7 in the energy range up to the fundamental absorption region indicates that electronic states of Zn2+ do not form additional channels of energy transfer to the emission centers. In the region of the fundamental absorption edge, the excitation spectra of the studied samples differ considerably. In the sample with Zn2+, a broad band peaking at 157 nm (7.89 eV) is observed. The obtained value corresponds to the energy of the direct creation of excitons in Ca8ZnLn(PO4)7 compounds—7.94 eV [2]. Therefore, we also attribute the peak at 7.89 eV to the direct exciton creation in Ca9−xZnxGd0.9Eu0.1(PO4)7 (x = 0, 1). A small hump can also be found at 172 nm (7.21 eV) in the excitation spectrum of Eu3+ for the sample with x = 1, while this peak dominates in the excitation spectra of both Eu3+ and Gd3+ emissions in the sample with x = 0. Previously, a set of sharp peaks related to 4f-5d transitions in Eu3+ were observed in the VUV spectral region in some wide bandgap compounds [51]. A similar origin could be supposed for the detected peak at 172 nm. However, this sharp peak is intensive in the excitation spectra of Gd3+ emission, thus indicating efficient energy transfer from Eu3+ to Gd3+. However, according to the analysis of the excitation spectra in UV spectral region such kind of energy transfer is inefficient (CTB is barely observed in the excitation spectra of Gd3+ emission) and therefore the attribution of the peak at 172 nm to 4f-5d transitions in Eu3+ is low probable. This peak can be tentatively ascribed to excitons localized near Gd3+. This may be the reason for the enhanced energy transfer from the host to Gd3+ in this compound. In the sample with Zn2+, localization of excitons near Gd3+ is less effective, which results in the increase in intensity of Eu3+ luminescence increases.

3. Experimental Section

3.1. Sample Preparation

Ca9−xZnxGd0.9Eu0.1(PO4)7 (x = 0.00–1.00, dx = 0.05) compounds were synthesized by solid-state method from stoichiometric mixtures of CaHPO4·2H2O (99.9%), CaCO3 (99.9%), ZnO (99.9%) and Ln2O3 (Ln3+ = Gd, Eu) (99.9%), purchased from Sigma-Aldrich, according to the reaction:
7 C a H P O 4 · 2 H 2 O + 2 x C a C O 3 + x Z n O + 0.45 G d 2 O 3 + 0.05 E u 2 O 3 = C a 9 x Z n x G d 0.9 E u 0.1 P O 4 7 + 17.5 H 2 O + 2 x C O 2
All reagents were controlled by PXRD for the absence of impurity phases. The stoichiometric mixtures were carefully grounded and very slowly heated up to 600 K for 9 h and then annealed at 1273 K for 10 h with several intermediate grindings followed by slow cooling (10 h) to room temperature (TR).

3.2. Experimental Description

The powder X-ray diffraction (PXRD) patterns were collected on a Thermo ARL X’TRA powder diffractometer (CuKα radiation, λ = 1.5418 Å, Bragg–Brentano geometry, scintillator detector) at TR in 2θ range of 5–65° with steps of 0.02°. The phase analysis of the obtained samples was carried out using the Crystallographica Search-March program (Version 2.0.3.1) and JCPDS PDF-2 database. Le Bail analysis was performed using JANA2006 program package [52]. The Debye–Scherrer equation [53] was implemented to count coherent scattering regions (crystalline sizes). LaB6 (SRM 660c) as a line shape standard was applied to determine instrumental broadening.
Scanning electron microscopy (SEM) study was performed using Tescan VEGA3 microscope equipped with LaB6 cathode. The SEM images were obtained using secondary electron detector. The analysis of the quantitative of elements concentration was determined by energy-dispersive X-ray (EDX) analysis. CaK, EuL, GdL and ZnL lines in the EDX spectra were used for the element content determination.
An Empyrean X-ray diffractometer (PANalytical, Almelo, Netherlands) equipped with a PIXcel3D 2D solid-state hybrid detector providing for counting photons with a high spatial resolution and a high dynamic range was used for registration a powder diffraction patterns. Each pixel is 55 μm × 55 μm and detector array is 256 × 256 pixels. Bragg–Brentano geometry was realized with a Bragg—BrentanoHD X-ray optical module (parabolic multilayer mirror), which monochromatized the primary X-ray beam and provided it with high intensities compare to the commonly used divergence slits and beta filters as well as increase the peak-to-background ratio and minimize excitation of fluorescent radiation from the sample. The X-ray generator (CuKα-radiation) was operated at 40 kV and 40 mA. Diffraction patterns were recorded in the range of 10÷1100 (2θ) with the step size of 0.01310 using continuous scan mode. PIXcel3D operated in a scanning line detector (1D) mode over its total active length (14 mm), which corresponds 3.30 (2θ) on the Empyrean goniometer with the radius of 240 mm. To avoid an influence of transparency effect of a material, we used a zero-background sample holder consisting of an obliquely cut silicon single crystal with a 32 mm diameter and 2 mm thickness. Rietveld analysis [54] was performed using Jana2006 program package. Illustrations were made using the VESTA program.
The second harmonic generation signal was measured with a Q-switched YAG:Nd laser at λω = 1064 nm in reflection mode for powder series with particle sizes of 40–60 μm. The laser operated with a repetition rate of 10 impulses/s and an impulse duration of about 5 ns. The laser beam was split into two beams to excite the radiation at the halved wavelength, λ = 532 nm, simultaneously in samples to be measured and in reference sample polycrystalline α-SiO2. The incident beam peak power was about 0.1 mW on a spot 3 mm in diameter on the surface of the sample.
Photoluminescence emission (PL) and excitation (PLE) spectra were measured using specialized setups for ultraviolet (UV) and vacuum ultraviolet (VUV) luminescence spectroscopy. A DDS-400 deuterium lamp was used as an excitation source for measurements in UV-visible spectral regions (200–500 nm). Excitation wavelengths were monochromatized using primary prism monochromator DMR-4. PLE spectra were measured with 5 nm spectral resolution. The sample was installed in a Janis VPF-800 nitrogen cryostat, which allows for changing the temperature in the range of 80–700 K. The luminescence signal was registered using an ARC SpectraPro-300i monochromator and a Hamamatsu H28259-01 photon counting head used for the measurements of luminescence.
The measurements in UV-VUV spectral region (130–400 nm) were performed using deuterium lamp Hamamatsu L11798 with MgF2 output windows as an excitation source. The lamp radiation was monochromatized using a McPherson 234/302 vacuum primary monochromator. The PLE spectra were measured with 14 nm resolution. The samples were placed in an ARS vacuum cryostat, which allows for measurements in the temperature range of 5–300 K. Luminescence was registered using Shamrock 303i monochromator equipped with ANDOR iDUS CCD detector with a spectral resolution of 0.3 nm.

4. Conclusions

The as-synthesized Ca9-xZnxGd0.9Eu0.1(PO4)7 solid solutions with β-Ca3(PO4)2-type structures do not contain impurity phases and all crystallize in R3c or R 3 ¯ c SGs depending on the content of Zn2+ ions. Centrosymmetry in Ca9−xZnxGd0.9Eu0.1(PO4)7 phosphors with x > 0.75 was reliably established based on the absence of the SHG effect. Together with data from X-ray analysis, this proves the existence of a pure centrosymmetric variant of the β-Ca3(PO4)2-type compounds with R 3 ¯ c SG and the interphase region between this phase and compounds with polar SG R3c at lower x. It is worth noting that SGs R3c and R 3 ¯ c are practically indistinguishable in routine laboratory X-ray diffraction experiments because the regions of centro- and non-centrosymmetric phases can only be definitely determined using additional methods such as SHG. Namely, in a Ca9−xZnxGd0.9Eu0.1(PO4)7 system, the change of symmetry was confirmed by this method. Concentration phase boundaries in this system were established according to zero (or very small) SHG signal for the R 3 ¯ c phase in interval 0.80 < x ≤ 1.00. Then, at 0.35 ≤ x ≤ 0.75 a transient two-phase state was observed, and at 0.00 ≤ x < 0.35 a non-centrosymmetric SG R3c was fully stabilized according to nearly constant non-zero SHG signal.
In the single-phase phosphate Ca8ZnGd(PO4)7, it was shown using the Rietveld method that Zn2+ ions occupy the M5 position, thus relieving a geometric stress in the structure. The M3 position is shifted from the special position (18d) to the semi-occupied position (36f), and P1 position is shifted from the special position (6a) to the semi-occupied general position (12c). The rationale for these shifts was the large values of the Uiso parameters (atomic displacement) in refining the structure under the assumption of locating M3 and P1 in special positions. Thus, the M3 position is shifted from the third-order axis, and its occupancy is equal to 0.5. The distribution of Gd3+ ions in Ca8ZnGd(PO4)7 by position in the structure turned out to be 71% in M3 and 29% in M1. This is characteristic of large cations such as Gd3+. Thus, we can distinguish the following factors, which probably positively affect the intensity of Eu3+ luminescence:
(1)
Shifting of the M3 position from the third-order axis;
(2)
Distortion of M3O8 polyhedra (local decrease of the symmetry);
(3)
General increase in symmetry of the structure (R3cR 3 ¯ c).
The relative intensity of Eu3+/Gd3+ emissions depend on their composition. It is shown that in compounds with x = 1 the energy transfer from the host to Eu3+ is improved, which results in the increase in Eu3+ luminescent intensity.
Two nonequivalent positions of Ln3+ ions were deduced from the emission spectra of Eu3+ as well as Gd3+ in the sample with x = 0. Therefore, Gd3+ ions could also be used as a luminescent marker to study the crystal structure of the compound.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/molecules28010352/s1, Table S1: The unit cells parameters, SHG signals (I/I(SiO2)) and supposed SG for Ca9-xZnxGd(PO4)7:0.1Eu3+ compounds; Tablse S2: Fractional atomic coordinates, site symmetry, isotropic displacement atomic parameters (Uiso) and site occupation for Ca8ZnGd(PO4)7 from PXRD data; Table S3: Selected distances (Å) in Ca8ZnGd(PO4)7 from PXRD data.

Author Contributions

Conceptualization, S.Y.S., I.V.N. and B.I.L.; Methodology, D.A.S. and V.A.M.; Formal analysis, Y.Y.D., D.A.S., V.A.M., S.N.P., V.D.R. and S.Y.S.; Investigation, Y.Y.D., D.A.S. and V.A.M.; Resources, D.V.D.; Data curation, O.V.B.; Writing—original draft, Y.Y.D. and V.D.R.; Writing—review & editing, D.A.S., S.N.P., S.Y.S., D.V.D., O.V.B., I.V.N. and B.I.L.; Visualization, Y.Y.D.; Supervision, B.I.L.; Funding acquisition, B.I.L. All authors have read and agreed to the published version of the manuscript.

Funding

The study was supported by RFBR (Project No. 20-03-00929). The research was also supported by the Development Program of the Interdisciplinary Scientific and Educational School of Lomonosov Moscow State University’s “The future of the planet and global environmental change” and the state assignment of the Chemistry Department of Moscow State University (Agreement No. AAAA-A21-121011590086-0). D.V.D. is grateful to the Scholarship of the President of the Russian Federation (CП-859.2021.1) and the state of the Russian Federation, state registration number 122011300125-2. D.A.S. is grateful to the support from the Estonian Research Council (Project: PUT PRG111).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dickens, B.; Schroeder, L.W.; Brown, W.E. Crystallographic studies of the role of Mg as a stabilizing impurity in β-Ca3(PO4)2. The crystal structure of pure β-Ca3(PO4)2. J. Solid State Chem. 1974, 10, 232–248. [Google Scholar] [CrossRef]
  2. Dikhtyar, Y.Y.; Spassky, D.A.; Morozov, V.A.; Deyneko, D.V.; Belik, A.A.; Baryshnikova, O.V.; Nikiforov, I.V.; Lazoryak, B.I. Site occupancy, luminescence and dielectric properties of β-Ca3(PO4)2-type Ca8ZnLn(PO4)7 host materials. J. Alloys Compd. 2022, 908, 164521. [Google Scholar] [CrossRef]
  3. Sun, S.; Lin, Z.; Zhang, L.; Huang, Y.; Wang, G. Growth and spectral properties of a new nonlinear laser crystal of Nd3+:Ca9Y0.5La0.5(VO4)7. J. Alloys Compd. 2013, 551, 229–232. [Google Scholar] [CrossRef]
  4. Lazoryak, B.I.; Baryshnikova, O.V.; Stefanovich, S.Y.; Malakho, A.P.; Morozov, V.A.; Belik, A.A.; Leonidov, I.A.; Leonidova, O.N.; Tendeloo, G. Van Ferroelectric and Ionic-Conductive Properties of Nonlinear-Optical Vanadate, Ca9Bi(VO4)7. Chem. Mater. 2003, 15, 3003–3010. [Google Scholar] [CrossRef]
  5. Lazoryak, B.I.; Teterskii, A.V.; Morozov, V.A.; Stefanovich, S.Y. Dielectric and nonlinear optical properties of the Ca9R(PO4)7 (R = Ln) phosphates. Russ. J. Inorg. Chem. 2005, 50, 986–989. [Google Scholar]
  6. Lazoryak, B.I.; Baryshnikova, O.V.; Malakho, A.P.; Kobyletskii, K.K.; Fursina, A.A.; Leonidova, O.N.; Morozov, V.A.; Leonidov, I.A.; Stefanovich, S.Y. Ferroelectric solid solutions in the Ca-3(VO4)(2)-BiVO4 system. Russ. J. Inorg. Chem. 2005, 50, 823–832. [Google Scholar]
  7. Stefanovich, S.Y.; Belik, A.A.; Azuma, M.; Takano, M.; Baryshnikova, O.V.; Morozov, V.A.; Lazoryak, B.I.; Lebedev, O.I.; Van Tendeloo, G. Antiferroelectric phase transition in Sr9In(PO4)7. Phys. Rev. B 2004, 70, 172103. [Google Scholar] [CrossRef]
  8. Teterskii, A.V.; Stefanovich, S.Y.; Lazoryak, B.I.; Rusakov, D.A. Whitlockite solid solutions Ca9−xMxR(PO4)7 (x = 1, 1.5; M = Mg, Zn, Cd; R = Ln, Y) with antiferroelectric properties. Russ. J. Inorg. Chem. 2007, 52, 308–314. [Google Scholar] [CrossRef]
  9. Benarafa, A.; Kacimi, M.; Coudurier, G.; Ziyad, M. Characterisation of the active sites in butan-2-ol dehydrogenation over calcium–copper and calcium–sodium–copper phosphates. Appl. Catal. A Gen. 2000, 196, 25–35. [Google Scholar] [CrossRef]
  10. Wagner, D.E.; Eisenmann, K.M.; Nestor-Kalinoski, A.L.; Bhaduri, S.B. A microwave-assisted solution combustion synthesis to produce europium-doped calcium phosphate nanowhiskers for bioimaging applications. Acta Biomater. 2013, 9, 8422–8432. [Google Scholar] [CrossRef]
  11. Deyneko, D.V.; Fadeeva, I.V.; Borovikova, E.Y.; Dzhevakov, P.B.; Slukin, P.V.; Zheng, Y.; Xia, D.; Lazoryak, B.I.; Rau, J.V. Antimicrobial properties of co-doped tricalcium phosphates Ca3-2(M′M″) (PO4)2 (M = Zn2+, Cu2+, Mn2+ and Sr2+). Ceram. Int. 2022, 48, 29770–29781. [Google Scholar] [CrossRef]
  12. Safiri, S.; Kolahi, A.; Cross, M.; Hill, C.; Smith, E.; Carson-Chahhoud, K.; Mansournia, M.A.; Almasi-Hashiani, A.; Ashrafi-Asgarabad, A.; Kaufman, J.; et al. Prevalence, Deaths, and Disability-Adjusted Life Years Due to Musculoskeletal Disorders for 195 Countries and Territories 1990–2017. Arthritis Rheumatol. 2021, 73, 702–714. [Google Scholar] [CrossRef] [PubMed]
  13. Ginebra, M.-P.; Montufar, E.B. Cements as bone repair materials. In Bone Repair Biomaterials; Elsevier: Amsterdam, The Netherlands, 2019; pp. 233–271. [Google Scholar]
  14. Heise, U.; Osborn, J.F.; Duwe, F. Hydroxyapatite ceramic as a bone substitute. Int. Orthop. 1990, 14, 329–338. [Google Scholar] [CrossRef] [PubMed]
  15. Rau, J.V.; Fadeeva, I.V.; Forysenkova, A.A.; Davydova, G.A.; Fosca, M.; Filippov, Y.Y.; Antoniac, I.V.; Antoniac, A.; D’Arco, A.; Di Fabrizio, M.; et al. Strontium Substituted Tricalcium Phosphate Bone Cement: Short and Long-Term Time-Resolved Studies and In Vitro Properties. Adv. Mater. Interfaces 2022, 9, 2200803. [Google Scholar] [CrossRef]
  16. Yuan, Z.; Bi, J.; Wang, W.; Sun, X.; Wang, L.; Mao, J.; Yang, F. Synthesis and properties of Sr2+ doping α-tricalcium phosphate at low temperature. J. Appl. Biomater. Funct. Mater. 2021, 19, 228080002199699. [Google Scholar] [CrossRef]
  17. Moussa, H.; Jiang, W.; Alsheghri, A.; Mansour, A.; El Hadad, A.; Pan, H.; Tang, R.; Song, J.; Vargas, J.; McKee, M.D.; et al. High strength brushite bioceramics obtained by selective regulation of crystal growth with chiral biomolecules. Acta Biomater. 2020, 106, 351–359. [Google Scholar] [CrossRef]
  18. Komlev, V.S.; Barinov, S.M.; Bozo, I.I.; Deev, R.V.; Eremin, I.I.; Fedotov, A.Y.; Gurin, A.N.; Khromova, N.V.; Kopnin, P.B.; Kuvshinova, E.A.; et al. Bioceramics Composed of Octacalcium Phosphate Demonstrate Enhanced Biological Behavior. ACS Appl. Mater. Interfaces 2014, 6, 16610–16620. [Google Scholar] [CrossRef]
  19. Tavoni, M.; Dapporto, M.; Tampieri, A.; Sprio, S. Bioactive Calcium Phosphate-Based Composites for Bone Regeneration. J. Compos. Sci. 2021, 5, 227. [Google Scholar] [CrossRef]
  20. Cho, J.; Park, J.H.; Kim, J.K.; Schubert, E.F. White light-emitting diodes: History, progress, and future. Laser Photon. Rev. 2017, 11, 1600147. [Google Scholar] [CrossRef]
  21. Atuchin, V.V.; Beisel, N.F.; Galashov, E.N.; Mandrik, E.M.; Molokeev, M.S.; Yelisseyev, A.P.; Yusuf, A.A.; Xia, Z. Pressure-Stimulated Synthesis and Luminescence Properties of Microcrystalline (Lu,Y)3Al5O12:Ce3+Garnet Phosphors. ACS Appl. Mater. Interfaces 2015, 7, 26235–26243. [Google Scholar] [CrossRef]
  22. Ji, H.; Wang, L.; Molokeev, M.S.; Hirosaki, N.; Xie, R.; Huang, Z.; Xia, Z.; ten Kate, O.M.; Liu, L.; Atuchin, V.V. Structure evolution and photoluminescence of Lu3(Al,Mg)2(Al,Si)3O12:Ce3+ phosphors: New yellow-color converters for blue LED-driven solid state lighting. J. Mater. Chem. C 2016, 4, 6855–6863. [Google Scholar] [CrossRef] [Green Version]
  23. George, N.C.; Denault, K.A.; Seshadri, R. Phosphors for Solid-State White Lighting. Annu. Rev. Mater. Res. 2013, 43, 481–501. [Google Scholar] [CrossRef]
  24. Bois, C.; Bodrogi, P.; Khanh, T.; Winkler, H. Measuring, simulating and optimizing current LED phosphor systems to enhance the visual quality of lighting. J. Solid State Light. 2014, 1, 5. [Google Scholar] [CrossRef] [Green Version]
  25. Cao, R.; Xue, H.; Yu, X.; Xiao, F.; Wu, D.; Zhang, F. Luminescence properties and synthesis of SrMgAl10O17:Mn4+ red phosphor for white light-emitting diodes. J. Nanosci. Nanotechnol. 2016, 16, 3489–3493. [Google Scholar] [CrossRef]
  26. Morassuti, C.Y.; Andrade, L.H.C.; Silva, J.R.; Bento, A.C.; Baesso, M.L.; Guimarães, F.B.; Rohling, J.H.; Nunes, L.A.O.; Boulon, G.; Guyot, Y.; et al. Eu2+,3+/Pr3+ co-doped calcium aluminosilicate glass for tunable white lighting devices. J. Alloys Compd. 2020, 817, 153319. [Google Scholar] [CrossRef]
  27. Li, G.; Lin, J. Recent progress in low-voltage cathodoluminescent materials: Synthesis, improvement and emission properties. Chem. Soc. Rev. 2014, 43, 7099–7131. [Google Scholar] [CrossRef] [PubMed]
  28. Bauer, M.; Glenn, T.; Monteith, S.; Gottlieb, J.F.; Ritter, P.S.; Geddes, J.; Whybrow, P.C. The potential influence of LED lighting on mental illness. World J. Biol. Psychiatry 2018, 19, 59–73. [Google Scholar] [CrossRef]
  29. Behar-Cohen, F.; Martinsons, C.; Viénot, F.; Zissis, G.; Barlier-Salsi, A.; Cesarini, J.P.; Enouf, O.; Garcia, M.; Picaud, S.; Attia, D. Light-emitting diodes (LED) for domestic lighting: Any risks for the eye? Prog. Retin. Eye Res. 2011, 30, 239–257. [Google Scholar] [CrossRef] [PubMed]
  30. Zhang, X.; Xu, J.; Guo, Z.; Gong, M. Luminescence and Energy Transfer of Dual-Emitting Solid Solution Phosphors (Ca,Sr)10Li(PO4)7:Ce3+,Mn2+ for Ratiometric Temperature Sensing. Ind. Eng. Chem. Res. 2017, 56, 890–898. [Google Scholar] [CrossRef]
  31. Qian, B.; Wang, Z.; Wang, Y.; Zhao, Q.; Zhou, X.; Zou, H.; Song, Y.; Sheng, Y. Comparative study on the morphology, growth mechanism and luminescence property of RE2O2S:Eu3+ (RE = Lu, Gd, Y) phosphors. J. Alloys Compd. 2021, 870, 159273. [Google Scholar] [CrossRef]
  32. Wang, H.; Li, Y.; Ning, Z.; Huang, L.; Zhong, C.; Wang, C.; Liu, M.; Lai, X.; Gao, D.; Bi, J. A novel red phosphor LixNa1-xEu(WO4)2 solid solution: Influences of Li/Na ratio on the microstructures and luminescence properties. J. Lumin. 2018, 201, 364–371. [Google Scholar] [CrossRef]
  33. Wang, H.; Yang, T.; Feng, L.; Ning, Z.; Liu, M.; Lai, X.; Gao, D.; Bi, J. Energy Transfer and Multicolor Tunable Luminescence Properties of NaGd0.5Tb0.5−xEux(MoO4)2 Phosphors for UV-LED. J. Electron. Mater. 2018, 47, 6494–6506. [Google Scholar] [CrossRef]
  34. Kumar, K.N.; Vijayalakshmi, L.; Choi, J.; Kim, J.S. Efficient red-luminescence of CaLa2ZnO5 phosphors co-doped by Ce3+ and Eu3+ ions. J. Alloys Compd. 2019, 787, 711–719. [Google Scholar] [CrossRef]
  35. Zhao, M.; Liao, H.; Ning, L.; Zhang, Q.; Liu, Q.; Xia, Z. Next-Generation Narrow-Band Green-Emitting RbLi(Li3SiO4)2:Eu2+: Phosphor for Backlight Display Application. Adv. Mater. 2018, 30, 1802489. [Google Scholar] [CrossRef] [PubMed]
  36. Zhang, Z.; Liu, L.; Liu, R.; Zhang, X.; Peng, X.; Wang, C.; Wang, D. High-brightness Eu3+-doped Ca9Gd(PO4)7 red phosphor for NUV light-emitting diodes application. Mater. Lett. 2016, 167, 250–253. [Google Scholar] [CrossRef]
  37. Huang, Y.; Zhao, W.; Cao, Y.; Jang, K.; Lee, H.S.; Cho, E.; Yi, S.-S. Photoluminescence of Eu3+-doped triple phosphate Ca8MgR(PO4)7 (R=La, Gd, Y). J. Solid State Chem. 2008, 181, 2161–2164. [Google Scholar] [CrossRef]
  38. Huang, Y.; Jiang, C.; Cao, Y.; Shi, L.; Seo, H.J. Luminescence and microstructures of Eu3+-doped in triple phosphate Ca8MgR(PO4)7 (R=La, Gd, Y) with whitlockite structure. Mater. Res. Bull. 2009, 44, 793–798. [Google Scholar] [CrossRef]
  39. Wang, P.; Hao, L.; Wang, Z.; Wang, Y.; Guo, M.; Zhang, P. Gadolinium-Doped BTO-Functionalized Nanocomposites with Enhanced MRI and X-ray Dual Imaging to Simulate the Electrical Properties of Bone. ACS Appl. Mater. Interfaces 2020, 12, 49464–49479. [Google Scholar] [CrossRef]
  40. Deyneko, D.V.; Morozov, V.A.; Hadermann, J.; Savon, A.E.; Spassky, D.A.; Stefanovich, S.Y.; Belik, A.A.; Lazoryak, B.I. A novel red Ca8.5Pb0.5Eu(PO4)7 phosphor for light emitting diodes application. J. Alloys Compd. 2015, 647, 965–972. [Google Scholar] [CrossRef]
  41. Deyneko, D.V.; Morozov, V.A.; Zhukovskaya, E.S.; Nikiforov, I.V.; Spassky, D.A.; Belik, A.A.; Lazoryak, B.I. The influence of second coordination-sphere interactions on the luminescent properties of β-Ca3(PO4)2-related compounds. J. Alloys Compd. 2020, 815, 152352. [Google Scholar] [CrossRef]
  42. Fadeeva, I.V.; Goldberg, M.A.; Preobrazhensky, I.I.; Mamin, G.V.; Davidova, G.A.; Agafonova, N.V.; Fosca, M.; Russo, F.; Barinov, S.M.; Cavalu, S.; et al. Improved cytocompatibility and antibacterial properties of zinc-substituted brushite bone cement based on β-tricalcium phosphate. J. Mater. Sci. Mater. Med. 2021, 32, 99. [Google Scholar] [CrossRef] [PubMed]
  43. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  44. Deyneko, D.V.; Nikiforov, I.V.; Spassky, D.A.; Dikhtyar, Y.Y.; Aksenov, S.M.; Stefanovich, S.Y.; Lazoryak, B.I. Luminescence of Eu3+ as a probe for the determination of the local site symmetry in β-Ca3(PO4)2 -related structures. CrystEngComm 2019, 21, 5235–5242. [Google Scholar] [CrossRef]
  45. Lazoryak, B.I. Comment on “Tuning of Photoluminescence and Local Structures of Substituted Cations in xSr2Ca(PO4)2−(1−x)Ca10Li(PO4)7:Eu2+ phosphors”. Chem. Mater. 2017, 29, 3800–3802. [Google Scholar] [CrossRef] [Green Version]
  46. Dikhtyar, Y.Y.; Deyneko, D.V.; Boldyrev, K.N.; Baryshnikova, O.V.; Belik, A.A.; Morozov, V.A.; Lazoryak, B.I. Crystal structure, dielectric and optical properties of β-Ca3(PO4)2-type phosphates Ca9-xZnxLa(PO4)7:Ho3+. J. Lumin. 2021, 236, 118083. [Google Scholar] [CrossRef]
  47. Belik, A.A.; Morozov, V.A.; Deyneko, D.V.; Savon, A.E.; Baryshnikova, O.V.; Zhukovskaya, E.S.; Dorbakov, N.G.; Katsuya, Y.; Tanaka, M.; Stefanovich, S.Y.; et al. Antiferroelectric properties and site occupations of R3+ cations in Ca8MgR(PO4)7 luminescent host materials. J. Alloys Compd. 2017, 699, 928–937. [Google Scholar] [CrossRef]
  48. Baur, W.H. The geometry of polyhedral distortions. Predictive relationships for the phosphate group. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1974, 30, 1195–1215. [Google Scholar] [CrossRef]
  49. El Khouri, A.; Elaatmani, M.; Della Ventura, G.; Sodo, A.; Rizzi, R.; Rossi, M.; Capitelli, F. Synthesis, structure refinement and vibrational spectroscopy of new rare-earth tricalcium phosphates Ca9RE(PO4)7 (RE = La, Pr, Nd, Eu, Gd, Dy, Tm, Yb). Ceram. Int. 2017, 43, 15645–15653. [Google Scholar] [CrossRef]
  50. Goodwin, D.W. Spectra and Energy Levels of Rare Earth Ions in Crystals. Phys. Bull. 1969, 20, 525. [Google Scholar] [CrossRef]
  51. van Pieterson, L.; Reid, M.F.; Wegh, R.T.; Soverna, S.; Meijerink, A. Meijerink, 4fn →4fn-15d transitions of the light lanthanides: Experiment and theory. Phys. Rev. B 2002, 65, 045113. [Google Scholar] [CrossRef]
  52. Petříček, V.; Dušek, M.; Palatinus, L. Crystallographic Computing System JANA2006: General features. Zeitschrift für Krist. Cryst. Mater. 2014, 229, 345–352. [Google Scholar] [CrossRef]
  53. Scherrer, P. Bestimmung dergrosse und der inneren struktur yon kolloiteilchen mittels. Gott. Nachr Math. Phys 1918, 2, 98–100. [Google Scholar]
  54. Rietveld, H.M. A profile refinement method for nuclear and magnetic structures. J. Appl. Crystallogr. 1969, 2, 65–71. [Google Scholar] [CrossRef]
Figure 1. β-Ca3(PO4)2-type structure with R3c (Ca9Gd(PO4)7) and R 3 ¯ c (Ca8ZnGd(PO4)7) space groups. Half of the spheres show occupancy of 0.5 (for example, in M3 site).
Figure 1. β-Ca3(PO4)2-type structure with R3c (Ca9Gd(PO4)7) and R 3 ¯ c (Ca8ZnGd(PO4)7) space groups. Half of the spheres show occupancy of 0.5 (for example, in M3 site).
Molecules 28 00352 g001
Figure 2. The SEM images for x = 0.00 (a), 0.35 (b), 0.50 (c), 0.75 (d) and 1.00 (e) in Ca9−xZnxGd0.9Eu0.1(PO4)7 series.
Figure 2. The SEM images for x = 0.00 (a), 0.35 (b), 0.50 (c), 0.75 (d) and 1.00 (e) in Ca9−xZnxGd0.9Eu0.1(PO4)7 series.
Molecules 28 00352 g002
Figure 3. Dependence of unit cell parameters (a) and the second harmonic generation signal (I/I(SiO2)) (b) on the Zn2+ concentration for Ca9−xZnxGd(PO4)7:0.1Eu3+ solid solutions.
Figure 3. Dependence of unit cell parameters (a) and the second harmonic generation signal (I/I(SiO2)) (b) on the Zn2+ concentration for Ca9−xZnxGd(PO4)7:0.1Eu3+ solid solutions.
Molecules 28 00352 g003
Figure 4. Fragment of observed, calculated and difference PXRD data for Ca8ZnGd(PO4)7. Tick marks denote the peak positions of possible Bragg reflections. The inset shows detailed fragment of PXRD data for 2θ = 24–40°.
Figure 4. Fragment of observed, calculated and difference PXRD data for Ca8ZnGd(PO4)7. Tick marks denote the peak positions of possible Bragg reflections. The inset shows detailed fragment of PXRD data for 2θ = 24–40°.
Molecules 28 00352 g004
Figure 5. PL spectra of Ca9-xZnxGd0.9Eu0.1(PO4)7 x = 0 (red), 1 (black) (λex = 163 nm) at 300 K (a) and 6K (b). The inset shows the detailed PL spectra in the 300–320 nm region for both temperatures.
Figure 5. PL spectra of Ca9-xZnxGd0.9Eu0.1(PO4)7 x = 0 (red), 1 (black) (λex = 163 nm) at 300 K (a) and 6K (b). The inset shows the detailed PL spectra in the 300–320 nm region for both temperatures.
Molecules 28 00352 g005
Figure 6. Detailed PL spectra in the 576–582 nm region (5D0–7F0 transition in Eu3+) measured at 6 K for Ca9−xZnxGd0.9Eu0.1(PO4)7 x = 0 (red), 1 (black).
Figure 6. Detailed PL spectra in the 576–582 nm region (5D0–7F0 transition in Eu3+) measured at 6 K for Ca9−xZnxGd0.9Eu0.1(PO4)7 x = 0 (red), 1 (black).
Molecules 28 00352 g006
Figure 7. Normalized PLE spectra of Eu3+ (λem = 620 nm, black), Gd3+ (λem = 313 nm, red) in Ca9−xZnxGd0.9Eu0.1(PO4)7 x = 0 (a), 1 (b).
Figure 7. Normalized PLE spectra of Eu3+ (λem = 620 nm, black), Gd3+ (λem = 313 nm, red) in Ca9−xZnxGd0.9Eu0.1(PO4)7 x = 0 (a), 1 (b).
Molecules 28 00352 g007
Figure 8. Energy position of the ground and excited levels of 4f Ln3+ (orange squares), ground levels 5d Ln3+ (cyan squares), 4f Ln2+ (pink dots) for Ca8ZnLn(PO4)7 (Ln3+ = Gd, Eu). Lower black bold line corresponds to the top of the valence band (VB), upper green bold line—bottom of the conduction band (CB). Dashed line near CB—the exciton formation energy (Eex).
Figure 8. Energy position of the ground and excited levels of 4f Ln3+ (orange squares), ground levels 5d Ln3+ (cyan squares), 4f Ln2+ (pink dots) for Ca8ZnLn(PO4)7 (Ln3+ = Gd, Eu). Lower black bold line corresponds to the top of the valence band (VB), upper green bold line—bottom of the conduction band (CB). Dashed line near CB—the exciton formation energy (Eex).
Molecules 28 00352 g008
Table 1. EDX analysis results, estimated crystalline size and average full width at half maximum (FWHM) for 0 2 10 (hkl) reflection at 2θ~31° (CuKa1Ka2 radiation) of the diffraction peaks for Ca9−xZnxGd0.9Eu0.1(PO4)7.
Table 1. EDX analysis results, estimated crystalline size and average full width at half maximum (FWHM) for 0 2 10 (hkl) reflection at 2θ~31° (CuKa1Ka2 radiation) of the diffraction peaks for Ca9−xZnxGd0.9Eu0.1(PO4)7.
xCa, at.%Zn, at.%Gd, at.%Eu, at.%Zn:Ca ratioCrystalline sizes, nmFWHM, °
0.3587.29 ± 0.293.19 ± 0.168.80 ± 0.190.72 ± 0.110.32 ± 0.03:8.73 ± 0.1175 ± 100.130 ± 0.013
0.5083.56 ± 0.505.74 ± 0.239.90 ± 0.180.80 ± 0.100.57 ± 0.02:8.36 ± 0.05103 ± 80.104 ± 0.006
0.7581.39 ± 0.468.34 ± 0.379.44 ± 0.200.83 ± 0.130.83 ± 0.04:8.14 ± 0.05108 ± 100.101 ± 0.006
1.0078.25 ± 0.1411.12 ± 0.229.65 ± 0.180.98 ± 0.081.11 ± 0.02:7.83 ± 0.01270 ± 200.072 ± 0.002
Table 2. Crystallographic data for Ca8ZnGd(PO4)7 (SG R 3 ¯  c, Z = 6, T = 293 K).
Table 2. Crystallographic data for Ca8ZnGd(PO4)7 (SG R 3 ¯  c, Z = 6, T = 293 K).
Sample CompositionCa8ZnGd(PO4)7
Lattice parameters: a, Å10.3796(6)
c, Å37.1316(3)
Unit cell volume, Å33460.69(4)
Calculated density, g/cm33.452(2)
Data Collection
DiffractometerEmpyrean X-ray
Radiation/wavelength (λ, Å)CuKα/1.540593, 1.544427
2θ range (o)10.013–109.983
Step scan (2θ)0.013
Imax51281
Number of points7691
Refinement
RefinementRietveld
Background functionPseudo-Voigt, 16 terms
No. of reflections (all/observed)488/473
No. of refined parameters/refined atomic parameters64/54
R and Rw (%) for Bragg reflection (Rall/Robs)8.81/9.07, 9.10/9.08
RP, RwP, Rexp (%)4.73, 6.64, 1.96
Max/min residual density (e) (Å3)1.58/−1.78
nf-CaM11.263(8)
nf-CaM30.605(5)
nf-ZnM50.864(8)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Dikhtyar, Y.Y.; Spassky, D.A.; Morozov, V.A.; Polyakov, S.N.; Romanova, V.D.; Stefanovich, S.Y.; Deyneko, D.V.; Baryshnikova, O.V.; Nikiforov, I.V.; Lazoryak, B.I. New Series of Red-Light Phosphor Ca9−xZnxGd0.9(PO4)7:0.1Eu3+ (x = 0–1). Molecules 2023, 28, 352. https://doi.org/10.3390/molecules28010352

AMA Style

Dikhtyar YY, Spassky DA, Morozov VA, Polyakov SN, Romanova VD, Stefanovich SY, Deyneko DV, Baryshnikova OV, Nikiforov IV, Lazoryak BI. New Series of Red-Light Phosphor Ca9−xZnxGd0.9(PO4)7:0.1Eu3+ (x = 0–1). Molecules. 2023; 28(1):352. https://doi.org/10.3390/molecules28010352

Chicago/Turabian Style

Dikhtyar, Yury Yu., Dmitry A. Spassky, Vladimir A. Morozov, Sergey N. Polyakov, Valerya D. Romanova, Sergey Yu. Stefanovich, Dina V. Deyneko, Oksana V. Baryshnikova, Ivan V. Nikiforov, and Bogan I. Lazoryak. 2023. "New Series of Red-Light Phosphor Ca9−xZnxGd0.9(PO4)7:0.1Eu3+ (x = 0–1)" Molecules 28, no. 1: 352. https://doi.org/10.3390/molecules28010352

Article Metrics

Back to TopTop