Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

A crystal defect led zero thermal quenching β-NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]Eu3+,Dy3+ red emitting phosphor

Shaokun Lingb, Chang Chenb, Kai Mengb, Yifeng Yanb, Junyun Mingb, Sen Liao*ab and Yingheng Huang*b
aSchool of Chemistry and Chemical Engineering, Guangxi University, Nanning, Guangxi 530004, China. E-mail: liaosen@gxu.edu.cn; liaosen380@sina.com; Fax: +86 771 3233718; Tel: +86 771 3233718
bSchool of Resources, Environment and Materials, Guangxi University, Nanning, Guangxi 530004, China. E-mail: huangyingheng@163.com; lsk0708@163.com; 249443176@qq.com; 1006779688@qq.com; 1024487385@qq.com; 2945539376@qq.com

Received 8th September 2022 , Accepted 28th September 2022

First published on 27th October 2022


Abstract

Red-light phosphors with extraordinary and stable thermal luminous properties must urgently be explored under the circumstances that commercial phosphors are suffering from serious thermal quenching effects and a lack of red-light components. Synthesized by a one-step hydrothermal method, a new type of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+ phosphor with notable thermal luminous stability is reported in this study. As well as energy transfer between Dy3+ and Eu3+, this novel red-light phosphor manifests zero thermal quenching (ZTQ) performance under an increasing temperature of measurement. The ZTQ property stems from the interior defects of the crystal produced by the non-equivalence replacement between distinct ions. Density Functional Theory (DFT) calculations were utilized to verify the formation energy of two kinds of defects that make a vital contribution to the ZTQ performance of the NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+ phosphor. This finding could make some contributions towards research into improving thermal luminous properties and stability.


1. Introduction

Since the birth of the commercial light emitting diode (LED) in 1968, LEDs have been applied in various fields due to its merits of being energy saving and environmentally friendly1–3 such as basic illumination,4 cultivation of plants,5,6 indoor illumination in hospitals for physical health care,7 and as a backlight source for large scale LCDs.8,9 The WLED is the most important and famous application of LEDs and is used in various application fields. Up to now, it is the most successful method that combines a YAG[thin space (1/6-em)]:[thin space (1/6-em)]Ce3+ yellow emitting phosphor with a blue-light emitting semiconductor InGaN chip to obtain white light.10,11 However, current commercial WLEDs illumination systems suffer from a shortage of high correlated color temperatures (CCTs) and low color rendering indexes (CRIs) due to the deficiency of red-light components.12,13 YAG[thin space (1/6-em)]:[thin space (1/6-em)]Ce3+ phosphors present serious quenching of luminous intensity at high operating temperatures,14–16 which diminishes the luminous properties of WLEDs. To overcome these shortcomings, research into stable and reliable red-light performance is prompted by the shortage of current commercial WLEDs.

Eu2+ doped nitride-based red phosphors provide a new approach to resolve the problem of a lack of a red-light source with high chemical and luminous stability for WLEDs. Sr[LiAl3N4][thin space (1/6-em)]:[thin space (1/6-em)]Eu2+ was reported as a novel type of red phosphor with zero thermal quenching red light emitting property in a temperature range from 300 K to 600 K.17 Other researchers have reported that Li2CaSi2N4 based phosphors present zero thermal quenching and color tunable properties.18 These reports indicate that nitride-based phosphors have an enormous potential to be an ideal matrix for red-light emission phosphors. Unfortunately, the major obstacle limiting further application of nitride-based red-light phosphors is high synthesis costs induced by the rigorous synthesis condition that require high synthetic temperature and purity of atmosphere. To successfully synthesize nitride-based phosphors, it is necessary to keep the raw materials in a highly pure N2 atmosphere. For example, Sr[LiAl3N4][thin space (1/6-em)]:[thin space (1/6-em)]Eu2+ was synthesized in a mixed atmosphere (N2[thin space (1/6-em)]:[thin space (1/6-em)]H2 = 95[thin space (1/6-em)]:[thin space (1/6-em)]5) at 1000 °C and lasted for 2 hours.17 Li2CaSi2N4 was prepared in a 0.4 MPa N2 atmosphere heated to 1000 °C and lasted for 4 hours.18

The trivalent europium ion, Eu3+, is considered as an ideal red-light source, due to its extraordinary red photoluminescence performance in inorganic substrates.19 Furthermore, Eu3+ red-light luminous materials, such as Y2O2S[thin space (1/6-em)]:[thin space (1/6-em)]Eu3+ and Y2O3[thin space (1/6-em)]:[thin space (1/6-em)]Eu3+, have been commercialized in LEDs and are the most well-known and successful red-light sources. Unfortunately, Y2O3[thin space (1/6-em)]:[thin space (1/6-em)]Eu3+ phosphors suffer from serious thermal quenching that diminishes the luminous performance with increasing temperature.20 Furthermore, Y2O2S[thin space (1/6-em)]:[thin space (1/6-em)]Eu3+ phosphors may decompose and release sulfides.21 This sulfide will not only lead to degeneration of luminous properties, but may also have a harmful effect on human health. These fatal defects constrain the utilization of Y2O2S[thin space (1/6-em)]:[thin space (1/6-em)]Eu3+ and Y2O3[thin space (1/6-em)]:[thin space (1/6-em)]Eu3+ red-light phosphors. It is urgent to synthesize a novel type of host material for red-light Eu3+ phosphors which possesses outstanding thermal luminous and chemical stability.

Fluorides, due to their superior chemical stability, low possibility of non-radiative transitions, and low phono energy, have a high potential as a matrix for luminescence ions. Mn4+-doped fluoride phosphors, such as K2TiF6[thin space (1/6-em)]:[thin space (1/6-em)]Mn4+ and K2SiF6[thin space (1/6-em)]:[thin space (1/6-em)]Mn4+, were reported.22,23 These phosphors also exhibit outstanding red-light properties. However, unfortunately, the synthesis of Mn4+-doped fluoride phosphors utilize HF frequently, which inevitably could cause physical harm to the human body and is not suitable for commercial production. Most importantly, fluorides have superb compatibility for luminous lanthanide ions.24–26 Possessing the various merits of fluorides, NaYF4 is considered as one of the most suitable host materials for lanthanide luminous ions. As reported previously, Yb3+ and Er3+ co-doped NaYF4 nanoparticles were employed to improve the accuracy of miRNA detection,27 indicating selected chemical stability and safety of NaYF4 for the human body. Unlike nitride-based phosphors, there are distinct synthesis strategies to obtain NaYF4 phosphors with high chemical stability and luminous properties, such as electrospinning,28 thermal decomposition,29 co-precipitation,30 and the hydrothermal method.31,32 In these mentioned synthesis methods, the hydrothermal method is the most cost-effective synthesis compared with thermal decomposition and electrospinning. Moreover, the hydrothermal method has better safety and operational processes.33 Compared with the co-precipitation method that synthesizes cubic NaYF4 (α-NaYF4), the hydrothermal method can obtain hexagonal NaYF4 (β-NaYF4) which possesses outstanding optical property compared to α-NaYF4.

In this study, Eu3+ and Dy3+ were chosen as the activator and sensitizer, respectively. The energy transfer between Dy3+ and Eu3+ could occur due to the higher energy level of the 4F9/2 of the Dy3+ ion compared to the 5D0 and 5D1 of the Eu3+ ion.34–36 The existence of Dy3+ leads to an improvement of the red-light emitting properties for Eu3+. A one-step hydrothermal method was utilized to synthesize a Eu3+ and Dy3+ co-doped hexagonal NaYF4 phosphor with outstanding chemical stability and thermal luminous performance. DFT calculations were used to determine the factor responsible for the extraordinary thermal luminous properties, which is distinct from other Eu3+-doped NaYF4 phosphors37,38 without intriguing thermal luminous properties. This study could make a contribution to tackling the problem of the lack of stable red-light components in high temperature working conditions.

2. Experimental

2.1 Experiment process

A series of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,xDy3+ (x = 0, 0.001, 0.002, 0.003, 0.004, and 0.005) were synthesized by a one-step hydrothermal method. The details of the synthesis are as follows: (i) NaNO3 (2 mol L−1), Y(NO3)3 (1 mol L−1), Eu(NO3)3 (1 mol L−1), Dy(NO3)3 (1 mol L−1), and NH4F (6 mol L−1) solutions were prepared. (ii) NaNO3, Y(NO3)3, Eu(NO3)3, and Dy(NO3)3 solutions were mixed with a stoichiometric ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)](1–0.065−x)[thin space (1/6-em)]:[thin space (1/6-em)]0.065[thin space (1/6-em)]:[thin space (1/6-em)]x, corresponding to the ion ratio of Na+[thin space (1/6-em)]:[thin space (1/6-em)]Y3+: Eu3+[thin space (1/6-em)]:[thin space (1/6-em)]Dy3+ and named solution A. Solution A was stirred for 20 minutes inside Teflon containers. (iii) 10 ml NH4F solution was added into solution A, which produced a white colloidal precursor. The mixture solution containing the precursor was continually stirred for 30 minutes. (iv) After even stirring, the Teflon containers with the mixed solution and precursors produced in step (iii) were inserted into the hydrothermal autoclave and heated to 230 °C. They lasted at this temperature for 72 hours. (v) The final products were washed several times and dried at 120 °C for 2 hours.

2.2 Characterization

The crystal structure of all samples was characterized by X-ray powder diffraction (Bruker D8). Refinement of the X-ray data was processed by the GSAS program. The morphology and energy disperse spectrometer of the sample were observed and detected by a scanning electron microscope from Hitachi (SU4800). High resolution transmission electron microscopy (HRTEM) and selected area electron diffraction were implemented by JEM-2100. To characterize the fluorescent performance, measurement of PL and PLE spectra, as well as the decay of the sample were implemented by a Horiba FluoroMax4 spectrophotometer. High temperature PL spectra and decay were measured on the same optical platform. ESCALAB-250+ X-ray photoelectron spectra were employed to analyze the elements and chemical bonds qualitatively. Crystal defects were characterized by electron paramagnetic resonance (EPR, Bruker EMXplus-6/1).

2.3 Calculations

The band structure, density of states, and formation energy of the crystal defects were calculated by the VASP program. The PAW-PBE and DFT+U methods were utilized for calculation of the band structure and density of states (DOS). The files VASP needed were prepared by VASPKIT.39 The process for determining the formation energy of defects can be described as the following: (i) geometry optimization of a single cell was implemented. (ii) On the basis of the optimization of a single cell, a supercell with 2 × 2 × 2 was built. (iii) Na and Y ions in a supercell were replaced by Y and Na ions, respectively. (iv) Geometry optimization for the whole system was implemented to find out the minimum energy.

3. Results and discussion

3.1 X-ray powder diffraction results

As illustrated in Fig. 1a, compared with the standard PDF card, the X-ray powder diffraction patterns of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,xDy3+ (x = 0, 0.001, 0.002, 0.003, 0.004, and 0.005) are in good agreement with the diffraction peaks of PDF#28-1192 which represent the phase of Na0.57Y0.39Yb0.06ErF4. From the XRD patterns, impurity phases were not detected, which proved the successful synthesis of a high purity single phase of a series of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,xDy3+ samples. As shown in Fig. 1b, compared with the standard PDF card, between a 2 theta value from 92° to 120°, diffraction peaks of all samples present a slight skewing to lower angles due to the larger ion radius of Eu3+ and Dy3+ compared to Yb3+ and Er3+.
image file: d2ra05674g-f1.tif
Fig. 1 X-ray patterns of the samples (i–vi) NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,xDy3+((i–vi), x = 0.00, 0.001, 0.002, 0.003, 0.004, and 0.005). (a) Full patterns. (b) Local expanded patterns.

3.2 Analysis of crystal morphology and elements

As illustrated in Fig. 2a–d, the morphology of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+ was observed by SEM at magnifications of 2.0 K, 3.0 K, 4.0 K, and 5.0 K, respectively. The sample was ground up with an agate mortar before observation. Crystal grains with an even prism shape and a smooth crystal grain surface were obtained. The size of the grains was detected to be 4 × 10 μm. From Fig. 2e–i, EDS mapping graphs of the sample indicate that all the elements Na, Y, F, Eu, Dy were uniformly dispersed in the crystal grains. Elements Eu and Dy were successfully incorporated into the β-NaYF4 substrate.
image file: d2ra05674g-f2.tif
Fig. 2 SEM images observed at magnifications of (a) 2.0 K, (b) 3.0 K, (c) 4.0 K, and (d) 5.0 K. EDS mapping of (e) Na, (f) Y, (g) F, (h) Eu, and (i) Dy.

The X-ray photoelectron spectrum of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+ is shown in Fig. 3a. Fig. 3b and c illustrate the high-resolution XPS spectra of Na+ and Y3+. Fig. 3e, the high-resolution XPS spectra of Eu3+, presents two different peaks owing to the 5d5/2 energy level. The peak located at a binding energy of 1134.83 eV and 1138.03 eV demonstrated by the fitted result may be ascribed to the splitting of the 5d5/2 energy level. Satellites peaks were present at 1152.56 eV to 1159.60 eV. Finally, the binding energy located at 1164.30 eV and 1164.83 eV can be ascribed to the 3d3/2 energy level. Fig. 3f shows the high-resolution XPS spectra of Dy3+ and its fitted result, the fitted peaks at a binding energy of 1296.62 eV could be attributed to the 3d5/2 energy level, respectively. The XPS results further indicate that the Dy3+ ion was successfully incorporated into the NaYF4 matrix.


image file: d2ra05674g-f3.tif
Fig. 3 (a) XPS survey spectrum of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+. High resolution spectra of (b) Na, (c) Y, (d) F, (e) Eu, and (f) Dy.

3.3 TEM, X-ray refinement, and crystal structure

As demonstrated in Fig. 4a, the TEM graph of the crystal grain is shown. Graphs of HRTEM and SAED were obtained based on Fig. 4a. As illustrated in Fig. 4b, the graph of the HRTEM shows the distance of the crystal face of (100) and (110), corresponding to the crystal face distance of 0.506 Å and 0.279 Å. As illustrated in Fig. 4c, redundant light spots are not observed, indicating the pure phase of β-NaYF4. Results of HRTEM and SAED were in good agreement with the XRD patterns which indicate the purity of phase. The diffraction spots of (100) and (110) were detected and shown in the SAED graph. The results of the HRTEM and SAED also provide reliable information on the X-ray refinement. The χ2, Rwp, Rp, and Rbragg values of the X-ray diffraction refinement are equal to 1.892, 8.50%, 6.27%, and 6.29, respectively, which indicates an extraordinary fitness result and reliable information of the crystal structure. The refinement pattern of XRD is shown in Fig. 4d, which is consistent with the experimental data.
image file: d2ra05674g-f4.tif
Fig. 4 (a) The TEM graph of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+. (b) The HRTEM graph of the sample. (c) The SAED graph of the sample. (d) The X-ray refinement pattern of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+.

Combining the X-ray diffraction data and GSAS program, the crystal information of the NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+ sample was obtained and is listed in Table 1. The result of the refinement indicates that the crystal system and space group of the sample are hexagonal and P63/m (176), respectively. Due to the relatively low Dy3+ concentration, the Dy3+ ion was not taken into consideration due to the few contributions to the XRD pattern. Na and Y atoms exist at two distinct positions in the β-NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+ structure. In the coordinate value (0.0, 0.0, 0.11413), Na and Y occupy an equivalent position, named Na1 and Y2, respectively. Similarly, at a position for which the coordinate value is (0.6667, 0.3333, 0.25), the Y and Na atoms occupy an equivalent atom position, labelled Y1 and Na2, respectively.

Table 1 The results of the Rietveld refinement of β-NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+
Formula sum β-Na(0.935Y)F4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+    
Formula weight 191.9889    
Crystal system Hexagonal    
Space group P63/m (176)    
Cell parameters a = 5.985309(27) Å b = 5.984911 Å c = 3.52072(5) Å
Cell volume 109.212(20)    
Z 6    
Atom x/a y/b z/c
Na1 0.0 0.0 0.11413(10)
Na2 0.6667 0.3333 0.25
Y1 0.6667 0.3333 0.25
Y2 0.0 0.0 0.11413(10)
Eu1 0.6667 0.3333 0.25
F1 0.3090(4) 0.2795(4) 0.25


As shown in Fig. 5a–e, the crystal structure of β-NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+ was portrayed by the Diamon-3 program. Watching the structure from the perspective parallel to the connection line of the vertex of the a and c axis, the integrated crystal structure is constituted of Na–F octahedron and Y–F decahedron. Named [YF8], the Y–F decahedron is composed of a Y atom and nine F atoms, as depicted in Fig. 5b. Consisting of two central Na atoms and six F atoms, the Na–F octahedron is named as [Na2F6], as illustrated in Fig. 5c. Noticeably, the [YF8] decahedrons and [Na2F6] octahedron occupy the centre of the crystal cell and the edge line parallel to the direction of the connected line between a and c axis. Fig. 5d illustrates the process by which Eu3+ and Dy3+ ions replace the Y3+ inside the [YF8] decahedron.


image file: d2ra05674g-f5.tif
Fig. 5 (a) The crystal structure graph of β-NaYF4. (b) Y–F decahedron. (c) Na–F octahedron. (d) The mechanism of incorporation of the Eu3+ and Dy3+ ions. (e) The ultraviolet visible absorption spectra of β-NaYF4. (f) The crystal structure graph of β-NaYF40.065Eu3+ (2 × 2 × 2 supercell).

To obtain the energy gap of β-NaYF4, ultraviolet absorption spectroscopy was employed. The ultraviolet absorption spectra and photo energy spectra of β-NaYF4 are shown in Fig. 5d. The energy gap is calculated by eqn (1):

 
(αhν)n = A(hν − Eg) (1)
In this equation, α, h, and ν represent the absorption coefficient, Planck's constant, and frequency, respectively. A is a constant. Eg refers to the energy gap. n can be equal to 1/2 or 2, which represents the indirect or direct energy gap, respectively. In this study, n is chosen to be 1/2 (ref. 40). The energy gap of β-NaYF4 is calculated to be 5.321 eV. A 2 × 2 × 2 supercell was built to illustrate the crystal structure model of β-NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+ clearly. The crystal structure of β-NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+ is shown in Fig. 5f, which demonstrates the crystal structure after incorporation of Eu3+.

The band structure and density of states are portrayed in Fig. 6a and b, respectively. The band structure shows that the energy gap calculated by the DFT + U algorithm is 5.246 eV, which is slightly lower than the experimental result. According to Fig. 6b, the p-orbital of the Y element makes an enormous contribution for the deep energy level. For the conduction band, the s-orbital of Na is the predominant contributor. Otherwise, the p-orbital of the element Na forms a continued DOS in the range of −15 eV to −37 eV. Finally, the density of states located in the top of the valence band is supplied by the d-orbital of the element Y and the p-orbital of element F. Evidently, the large energy gap provides extraordinary circumstances for incorporation and transition of Eu3+ and Dy3+, according to the results of the band structure and density of states.


image file: d2ra05674g-f6.tif
Fig. 6 (a) The band structure of β-NaYF4. (b) The total and projected density of states of β-NaYF4.

3.4 Luminescence properties of β-NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,xDy3+

As exhibited in Fig. 7a, the emission spectra of β-NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,xDy3+ (x = 0.001, 0.002, 0.003, 0.004, and 0.005), all samples present strong red-light emission. Otherwise, a main emission peak emerges at 615 nm, which could be ascribed to the 5D07F2 transition. At a wavelength of 592 nm, orange-red light emission was detected, which contributed to the 5D07F2 transition. According to the Judd–Ofelt theory,41,42 the luminous intensity ratio of 5D07F2 and 5D07F1 (R = I2/I1) is a vital criterion to confirm the emission mode determined by the magnetic dipole (MD) transition (5D07F2) or electric dipole (ED) transition (5D07F1). When the R value is larger than 1, the emission mode would be determined by the ED transition. On the contrary, if the R value is smaller than 1, the emission mode would be determined by the MD transition. As the Dy3+concentration increases, the R values of a series of β-NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,xDy3+ are 1.997, 1.983, 2.050, 2.049, 1.904, and 2.032, which suggests that the emission mode of all the samples are decided by the electric dipole (ED) transition.
image file: d2ra05674g-f7.tif
Fig. 7 (a) Emission spectra of samples incorporated with different concentrations of Dy3+. (b) Emission spectra of samples incorporated with different concentrations of Dy3+ shown by a 2-dimensional graph. (c) Emission intensity of the major emission peak. (d) CIE chromaticity diagram and luminous photo of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+.

As illustrated in Fig. 7a–c, the luminous intensities of a series of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,xDy3+ increase in a range of concentration of Dy3+ from 0 to 0.03, a declining trend of emission intensity emerges when the concentration of Dy3+ exceeds 0.003. The luminous intensities of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0,065Eu3+,xDy3+ (x = 0.001, 0.002, 0.003, 0.004, and 0.05) are 399.11%, 522.65%, 561.67%, 545.60%, and 492.45% of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0,065Eu3+, respectively. The results of the emission spectra indicate that concentration quenching occurs when the concentration of Dy3+ reaches 0.003. The optimal incorporating concentration of Dy3+ is 0.003. To elucidate the mechanism of quenching, eqn (2)43 is utilized:

 
image file: d2ra05674g-t1.tif(2)
where V, xc, and N are equal to the crystal cell volume, concentration of sample corresponding to supreme luminous intensity, and the number of possible replacement sites of Eu3+ and Dy3+, respectively. In this work, according to Table 1, V and N are equal to 109.228 and 6, while xc is 0.068. The Rc value is calculated to be 8.035, which is far away from 5 Å. The calculation results of the Rc value suggest that the electric multipolar interaction is the chief factor for concentration quenching. As illustrated in Fig. 7d, the sample NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+ shows a brilliant orange-red emitting color under the excitation of a UV-lamp.

The quantum yield (QY) and CIE coordinates of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,xDy3+ are listed in Table 2. The QY increases from 64.65% to 77.49% as the concentration of Dy3+ increases from 0.001 to 0.003. When the concentration of Dy3+ is 0.004 and 0.005, the QYs of those two samples are 73.02% and 71.19%, respectively. As the concentration of the Dy3+ ion increases, the CIE coordinates of x slightly decrease from 0.6272 to 0.6251, while the y value of the CIE coordinates increases from 0.3724 to 0.3744.

Table 2 Quantum yield and CIE coordinates of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,xDy3+
Sample Quantum yield (%) CIE coordinate x CIE coordinate y
NaY4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.001Dy3+ 64.65 0.6272 0.3724
NaY4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.002Dy3+ 69.53 0.6263 0.3733
NaY4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+ 77.49 0.626 0.3736
NaY4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.004Dy3+ 73.02 0.6259 0.3737
NaY4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.005Dy3+ 71.19 0.6251 0.3744


As demonstrated in Fig. 8a, the decay of all samples was measured at room temperature and fitted by the following equation:44

 
image file: d2ra05674g-t2.tif(3)
where It and I0 are the luminescence intensity at a selected time t and 0. A1 and A2 are constants and τ1 and τ2 are the brisk and unhurried time for decay constants, respectively. According to eqn (3), the average decay time is expressed as the following expression:
 
τ = (A1τ12 + A2τ22) / (A1τ1 + A2τ2) (4)
as the concentration of Dy3+ increases from 0 to 0.005, the lifetimes of the samples are fitted to be 6.63 ms, 6.48 ms, 6.48 ms, 6.08 ms, 5.99 ms, and 5.97 ms, respectively, indicating a decreasing trend with increasing concentration of Dy3+, which can be ascribed to a poor impact stemming from concentration quenching. As shown in Fig. 8b, the lifetimes of Dy3+ in NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+ were detected and are depicted. The lifetimes of Dy3+incorporated into NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+ are fitted to be 0.89 ms, 0.87 ms, 0.78 ms, 0.77 ms, and 0.72 ms, corresponding to concentrations of Dy3+ of 0.01, 0.02, 0.03, 0.04, and 0.05, respectively. The fitted results of decay demonstrate a linear decrease with the increase in the concentration of Dy3+, suggesting the existence of energy transfer from Dy3+ to the Eu3+ ion. To calculate the efficiency of energy transfer, eqn (5)45 is employed:
 
image file: d2ra05674g-t3.tif(5)
where τ and τ0 are the decay of Dy3+ in the matrix with and without Eu3+, respectively. As illustrated in Fig. 8c, when the Dy3+ was incorporated solely into the NaYF4 matrix with a concentration range from 0.001 to 0.005, the lifetimes of Dy3+ were calculated to be 1.10 ms, 1.14 ms, 1.07 ms, 1.01 ms, and 1.00 ms, respectively. Hence, energy transfer efficiency can be calculated to be 19.09%, 23.68%, 27.10%, 23.76%, and 28.00%. An increasing trend is shown with the increasing concentration of Dy3+. Despite the higher ηT of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.005Dy3+ than NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+, the luminous intensity of the former is still lower than the latter due to the stronger concentration quenching effect.


image file: d2ra05674g-f8.tif
Fig. 8 (a) Decay of Eu3+ in NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,xDy3+. (b) Decay of Dy3+ in NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+. (c) Decay of Dy3+ in NaYF4.

To elucidate the energy transfer mechanism, it is necessary to express the relationship between the luminous intensity and concentration of ions by a suitable method. The following equation45 is utilized:

 
I0/ICn/3 (6)
where I0, I, and C represent the luminous intensity of the Dy3+ion in the NaYF4 matrix, the luminous intensity of Dy3+ in the NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+ samples, and the total concentration of Dy3+ and Eu3+, respectively. Otherwise, n can be equal to 6, 8, and 10 which represent the dipole–dipole, the dipole–quadrupole, and the quadrupole–quadrupole interactions.

The luminous intensity of Dy3+ in the absence and presence of Eu3+ was measured by emission spectra presented in Fig. 9a and b to acquire I0 and I. The normalized intensities of I0 and I are compared in Fig. 9c. It is remarkable that the changing trend for luminous intensity of Dy3+ is identical to the intensity of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,xDy3+ excited by 393 nm. According to Fig. 9d that shows the fitted results of I0/I and Cn/3, when the n value is equal to 8, the R2 is the maximum value that equals 0.40286. The fitted result indicates that the energy transfer between Dy3+ and Eu3+ is ascribed to dipole–dipole interactions. Based on Fig. 9e–h, which show the excitation and emission spectra of Eu3+ and Dy3+, the schematic diagram of energy transfer is portrayed in Fig. 9i. Under the excitation of 393 nm, electrons in the ground state of Eu3+ are excited to the excited state (5L6) and transition to 4F1 and 4F2 from 5D0. Otherwise, the electrons in the ground state of Dy3+ are excited to the 4I13/2 and 4I15/2 energy level and transition to 6H13/2 and 6H15/2 from 5F9/2. Due to the higher energy of 5F9/2 compared to 5D0, the electrons will transition from 5F9/2 to the 5D0 energy level, which will lead to energy transfer of Dy3+ to Eu3+.


image file: d2ra05674g-f9.tif
Fig. 9 (a) Emission spectra of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]xDy3+. (b) Emission spectra of Dy3+ in NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,xDy3+. (c) Values of I and I0. (d) The fitted results of n = 6, n = 8, and n = 10. (e) Excitation spectra and (f) emission spectra of Eu3+ in NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+. (g) Excitation spectra and (h) Emission spectra of Dy3+ in NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+. (i) A schematic diagram of energy transfer from Dy3+ to Eu3+.

3.5 Thermal luminous properties of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+

As depicted in Fig. 10a and b, under the excitement wavelength of 393 nm, the emission intensity of the sample demonstrates a linear increase in the temperature range from 303 K to 363 K. The luminous intensity of the sample was determined at 363 K and the optimal temperature for luminous intensity is 121.07% of intensity at 303 K. When the measurement temperature was higher than 363 K, the luminous intensity shows a linear decreasing trend. The luminous intensity measured at 443 K and 483 K remains 104.99% and 83.03% of luminous intensity at 303 K, respectively. There are not any additional emission peaks as measurement temperature increases. The shape of the emission spectra measured at high temperatures is identical to the spectra at room temperature. This result suggests that the sample of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.03Dy3+ presents outstanding thermal luminous performance. Based on the emission spectra measured at distinct temperatures, as illustrated in Fig. 10c, the change of CIE coordinate at different measurement temperatures shows a negligible change after 443 K. Before being heated to 443 K, the CIE coordinate in the temperature range from 323 K to 423 K can be considered as invariable, indicating the extraordinary red-light stability presented by the sample.
image file: d2ra05674g-f10.tif
Fig. 10 (a) Emission spectrum of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+ at a distinct temperature. (b) Luminous intensity of the primary emission peak at temperatures 303 K to 543 K. (c) Change of CIE color coordinates from 303 K to 543 K. (d) Decay measured at 303 K, 363 K, 423 K, 483 K, and 543 K.

As portrayed in Fig. 10d, the decay measured at 303 K and 363 K indicates a slight decline from 5.97 ms to 5.81 ms which corresponds to increasing luminous intensity from 303 K to 363 K. A sharp decrease of decay from 5.52 ms to 4.84 ms represents the dwindling process of thermal luminous intensity led by the thermal quenching effect at a temperature range from 423 K to 543 K.

3.6 The zero thermal quenching mechanism of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+

The ZTQ performance presented by the sample can be ascribed to the crystal defect, as reported previously.46,47 The source of the crystal defect should be taken into consideration. As illustrated in Fig. 3b and c, the high resolution XPS of Na and Y present two distinct chemical bonds of Na–F and Y–F, respectively. As Fig. 3b presents, two peaks are portrayed at 1070.95 eV and 1071.84 eV, which are close to 1071.4 eV[thin space (1/6-em)]48 and 1072.5 eV,49 corresponding to the standard binding energy of Na–F, respectively. As portrayed in Fig. 3c, the Y–F bond presents two kinds of binding energy, corresponding to 158.82 eV and 160.88 eV. Both those values are close to 159.10 eV[thin space (1/6-em)]50 and 161.20 eV,51 representing the standard binding energy of YF3. The existence of the distinct chemical bonds of Na–F and Y–F is also confirmed by the crystal information listed in Table 1. The table presents two different occupancies for Na+ and Y3+ ions. Combined with the result of X-ray refinement, nonequivalent replacement between Na+ and Y3+ ions can be confirmed.

As shown in Fig. 11a, the Y3+ ion replaces the Na+ ion inside the [Na2F6] octahedron, which will produce a negative vacancy named image file: d2ra05674g-t4.tif and a cation vacancy image file: d2ra05674g-t5.tif. Similarly, the Na+ ion will replace the interior Y3+ ion of the [YF8] decahedrons, as illustrated in Fig. 11a, which would introduce the negative vacancy image file: d2ra05674g-t6.tif and the cation vacancy image file: d2ra05674g-t7.tif. Two pairs of vacancies would be produced due to the requirement of electric neutrality for a whole crystal. A nonequivalent replacement process can be inferred and expressed by the following expression:

 
image file: d2ra05674g-t8.tif(7)
 
image file: d2ra05674g-t9.tif(8)


image file: d2ra05674g-f11.tif
Fig. 11 (a) Replacement between Na+ and Y3+. (b) Calculation model for the VY vacancy. (c) Calculation model for the VNa vacancy. (d) EPR spectra at 100 K, 200 K, and 298 K. (e) Calculated g values at 100 K, 200 K, and 298 K based on EPR spectra. (f) A zero thermal quenching mechanism schematic diagram of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+.

EPR, a well acknowledged method to characterize defects, was utilized for the purpose of characterizing crystal defects. The EPR signal was measured at 100 K, 200 K, and 298 K, as shown in Fig. 11b, manifesting a single kind of signal. The intensity of the signal decreases from 100 K to 298 K. Comparing the signal at 298 K with 200 K, the intensity of the signal shows a slight decline. The EPR signal was converted to a g value, which is a critical parameter for crystal defects, according to the following equation:52

 
image file: d2ra05674g-t10.tif(9)

In this equation, the ν refers to the measurement frequency. B0 represents the intensity of the magnetic field. As illustrated in Fig. 11c, the g value was calculated to be 1.9882, 1.9883, and 1.9878 at 100 K, 200 K, and 298 K, respectively. The calculated results clearly show the stable existence of crystal defects at room temperature. Furthermore, the sample keeps a stable g value at a distinct measurement temperature. Typical g values for the element Eu are 0.6134 and 1.3887, corresponding to 153Eu and 151Eu.53 Similarly, representative g values for element Dy are 0.5122 and 0.7177, corresponding to 161Dy and 163Dy,53 respectively. Those four values are far away from the calculated g values according to EPR, which suggests that the EPR signal does not stem from the unpaired electrons of elements Eu and Dy. It is confirmed that the crystal defect leads to the signal detected by EPR.

To further confirm the type of crystal defect, VASP was employed to calculated the formation energy of a defect. Eqn (10)54 is necessary for the calculation:

 
image file: d2ra05674g-t11.tif(10)
where Ef (defect) refers to the formation energy of the defect. Et (defect) and Et (perfect) represent the total energy of a crystal with a present and absent defect, respectively. The n and μ correspond to the number of atoms and chemical potential of the element. The q is the charge state of the crystal defect. Finally, εVBM and εF are the maximum values of the valence band and the energy of Fermi level, respectively. The chemical potential of Na originates from the minimized total energy of Na (P63/mmc) divided by two Na atoms. Similarly, the chemical potential of Y and F originate from the minimum total energy of Y (P63/mmc) and F2 (P3/mn) divided by two Y atoms and eight F atoms, respectively. The calculation model is presented in Fig. 11b and c.

Table 3, listing all the values of formation energy, suggests that the formation energy of image file: d2ra05674g-t12.tif (−1310.7589 eV) and image file: d2ra05674g-t13.tif (−1267.5217 eV) is much higher than image file: d2ra05674g-t14.tif (−43.2101 eV) and image file: d2ra05674g-t15.tif (18.2062 eV). The DFT calculation results demonstrate that the formation of image file: d2ra05674g-t16.tif and image file: d2ra05674g-t17.tif vacancies will occur due to the requirement for electrons to balance, rather and image file: d2ra05674g-t18.tif and image file: d2ra05674g-t19.tif. Furthermore, image file: d2ra05674g-t20.tif and image file: d2ra05674g-t21.tif will be hard to produce due to the high formation energy, despite the nonequivalent replacement between Na+ and Y3+ located in the [YF8] decahedron.

Table 3 Formation energy of the distinct defect
 

image file: d2ra05674g-t22.tif

image file: d2ra05674g-t23.tif

image file: d2ra05674g-t24.tif

image file: d2ra05674g-t25.tif

Et (perfect) −426.2295 −426.2295 −426.2295 −426.2295
Et (defect) −1724.1516 −1732.7422 −449.95187 −440.36326
εVBM 7.6929 7.6929 7.6929 7.6929
εF 5.246 5.246 5.246 5.246
niμi −13.0771 −13.0771 −6.4261 −6.4261
Ef (defect) −1310.7589 −1267.5217 −43.2101 18.2062


On the other hand, the crystal defect plays the vital role of the electron trap.55,56 Based on this property, unpaired electrons induced by nonequivalent replacement are captured by the electron trap to maintain the electric neutrality of the whole system. Combining the EPR spectra, image file: d2ra05674g-t26.tif and image file: d2ra05674g-t27.tif vacancies show a single signal in Fig. 11a and b, due to the identical chemistry environment of the two types of vacancies. The expression expressing the formation of crystal defect should be modified as the following:

 
image file: d2ra05674g-t28.tif(11)

Based on the analysis of EPR and DFT calculations, image file: d2ra05674g-t29.tif and image file: d2ra05674g-t30.tif vacancies, treated as electronic traps, make a predominant contribution to the zero thermal quenching performance of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+. As illustrated in Fig. 11f, with an increase of temperature, under thermal stimulation, electrons caught in electronic traps will be excited to the 5L6 energy level and transition to 4FJ (J = 0, 1, 2, 3, 4) level, to compensate for the loss of luminous intensity induced by thermal quenching. In the temperature range of 303 K to 363 K, the luminous intensity increases linearly due to the increasing number of excited electrons stored in electronic traps, corresponding to the stage of increasing luminous intensity. This process is presented by the lifetimes measured at 303 K and 363 K. Comparing the decay times of 363 K with 303 K, the lifetimes decrease by 2.68%, while the decay at 423 K, 483 K, and 543 K declines by 7.82%, 14.572%, and 18.972%, respectively. The change of decay time shows the evident compensation effect induced by electrons in defects at 303 K to 363 K. After heating to 386 K, the luminous intensity begins to decline due to the decreasing number of excited electrons in electronic traps. Noticeably, the thermal luminous intensities remain 104.99% and 83.03% at 443 K and 483 K, respectively. The NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+ red-light phosphor demonstrates extraordinary thermal luminous properties.

3.7 Fabrication of the prototype red light LED

As shown in Fig. 12a, the prototype red LED was fabricated by coating the NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+ phosphor onto the surface of the 395 nm UV-chip. The red light ratio reached 35.4% under the 9 mA driving current, indicating the extraordinary red-light performance of the NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+ red-light phosphor. A white light emitting diode was fabricated by combining the YAG[thin space (1/6-em)]:[thin space (1/6-em)]Ce3+ yellow phosphor with the NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+ phosphor and 395 nm UV chip. Under a current of 9 mA, the correlated color temperature and color-rendering index of this fabricated WLED are 5698 K and 82.2, respectively. The CIE coordinate of the fabricated WLED is (x, y) = (0.3284, 0.3243), which is the white light area. The result shows that NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+ has outstanding thermal stability and the red-light source is an extraordinary candidate for WLEDs. Fig. 12b shows the color region of the prototype LED in the CIE chromaticity diagram, unveiling a close color area to white light. This result provides powerful proof that the NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+ phosphor is an ideal red-light component for WLEDs.
image file: d2ra05674g-f12.tif
Fig. 12 (a) Luminescent performance of prototype LEDs (NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+ + 395 nm UV chip) and luminescent graph. (b) The CIE chromaticity diagram of fabricated LEDs. (c) The CIE coordinate and fabricated WLEDs combining the YAG[thin space (1/6-em)]:[thin space (1/6-em)]Ce3+ yellow phosphor with NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+ phosphor and 395 nm UV chip.

4. Conclusions

A series of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,xDy3+ (x = 0.001, 0.002, 0.003, 0.004, 0.005) red-light phosphors were synthesized by a one-step hydrothermal method. The energy transfer effect was determined to be due to dipole–quadrupole interactions. With the increasing concentration of Dy3+, the luminous intensity of β-NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,xDy3+ is 399.11%, 522.65%, 561.67%, 545.60%, and 492.45% of NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0,065Eu3+, respectively. The zero thermal quenching property of the NaYF4[thin space (1/6-em)]:[thin space (1/6-em)]0.065Eu3+,0.003Dy3+ phosphor can be ascribed to the thermal stimulation of electrons stored in the crystal defect which is treated as an electronic trap. DFT calculations were employed to elucidate that image file: d2ra05674g-t31.tif and image file: d2ra05674g-t32.tif vacancies are predominant in the crystal and lead to the zero thermal quenching property. The formation energy of the image file: d2ra05674g-t33.tif and image file: d2ra05674g-t34.tif vacancies was calculated to be −43.2101 eV and 18.2062 eV, respectively.

Conflicts of interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgements

This research is supported by the National Natural Science Foundation of China (Grant No. 21661006 and No. 21965004), the Natural Science Foundation of Guangxi Zhuang Autonomous Region, China (Grant No. 2019GXNSFDA245022, No. 2020GXNSFAA159036), the Scientific Research Foundation of Guangxi University (Grant No. XDZ140116), the innovation Project of Guangxi Graduate Education (Grant No. YCSW2020015), and the Students Experimental Skills and Innovation Ability Training Fund Project of Guangxi University (No. S202210593143). The calculation part was supported by the high performance calculation center of Guangxi University.

References

  1. S. Liang, P. Dang, G. Li, M. S. Molokeev, Y. Wei, Y. Wei, H. Lian, M. Shang, A. A. Al Kheraif and J. Lin, J. Mater. Chem. C, 2018, 6, 6714–6725 RSC.
  2. S. Pimputkar, J. S. Speck, S. P. DenBaars and S. Nakamura, Nat. Photonics, 2009, 3, 180–182 CrossRef CAS.
  3. E. F. Schubert and J. K. Kim, Science, 2005, 308, 1274–1278 CrossRef CAS PubMed.
  4. P. M. Pattison, J. Y. Tsao, G. C. Brainard and B. Bugbee, Nature, 2018, 563, 493–500 CrossRef CAS PubMed.
  5. R. J. Bula, R. C. Morrow, T. W. Tibbitts, D. J. Barta, R. W. Ignatius and T. S. Martin, HortScience, 1991, 26, 203–205 CAS.
  6. C. Kubota, T. Eguchi and M. Kroggel, Sci. Hortic., 2017, 226, 366–371 CrossRef CAS.
  7. S. A. Rahman, M. A. St Hilaire and S. W. Lockley, Physiol. Behav., 2017, 177, 221–229 CrossRef CAS PubMed.
  8. J. Chang, D.-s. Yang, Z.-h. Sun, Z.-c. Weng, H.-l. Jiang and T. Ren, Chin. J. Lumin., 2008, 29, 195–199 Search PubMed.
  9. J. W. Liu, G. Wang, L. Ma, K. L. Zhang, L. L. Zhang, Q. J. Wang and L. F. Wan, East China Normal University, Shanghai, Peoples R China, 2007.
  10. Z. Xia and Q. Liu, Prog. Mater. Sci., 2016, 84, 59–117 CrossRef CAS.
  11. Z. Zhang, C. Ma, R. Gautier, M. S. Molokeev, Q. Liu and Z. Xia, Adv. Funct. Mater., 2018, 28, 1804150 CrossRef.
  12. H. Liu, Y. Luo, Z. Mao, L. Liao and Z. Xia, J. Mater. Chem. C, 2014, 2, 1619–1627 RSC.
  13. W. Lu, M. Jiao, B. Shao, L. Zhao and H. You, Inorg. Chem., 2015, 54, 9060–9065 CrossRef CAS PubMed.
  14. L. Chen, C.-C. Lin, C.-W. Yeh and R.-S. Liu, Materials, 2010, 3, 2172–2195 CrossRef CAS.
  15. C. C. Lin and R. S. Liu, J. Phys. Chem. Lett., 2011, 2, 1268–1277 CrossRef CAS PubMed.
  16. J. Ueda, P. Dorenbos, A. J. J. Bos, A. Meijerink and S. Tanabe, J. Phys. Chem. C, 2015, 119, 25003–25008 CrossRef CAS.
  17. P. Pust, V. Weiler, C. Hecht, A. Tucks, A. S. Wochnik, A. K. Henss, D. Wiechert, C. Scheu, P. J. Schmidt and W. Schnick, Nat. Mater., 2014, 13, 891–896 CrossRef CAS PubMed.
  18. Q. Wei, J. Ding and Y. Wang, Chem. Eng. J., 2020, 386, 124004 CrossRef CAS.
  19. K. Binnemans, Coord. Chem. Rev., 2015, 295, 1–45 CrossRef CAS.
  20. H. Peng, H. Song, B. Chen, J. Wang, S. Lu, X. Kong and J. Zhang, J. Chem. Phys., 2003, 118, 3277–3282 CrossRef CAS.
  21. Y. Zhou, Z. Song, Y. Li, Q. Liu, Z. Xu, R. Hu, X. Zhang, J. Qiu and Z. Yang, J. Rare Earths, 2016, 34, 1188–1192 CrossRef CAS.
  22. H. Zhu, C. C. Lin, W. Luo, S. Shu, Z. Liu, Y. Liu, J. Kong, E. Ma, Y. Cao, R. S. Liu and X. Chen, Nat. Commun., 2014, 5, 4312 CrossRef CAS PubMed.
  23. F. Tang, Z. Su, H. Ye, W. Gao, X. Pan and S. Xu, ACS Omega, 2018, 3, 13704–13710 CrossRef CAS PubMed.
  24. H.-X. Mai, Y.-W. Zhang, R. Si, Z.-G. Yan, L.-D. Sun, L.-P. You and C.-H. Yan, J. Am. Chem. Soc., 2006, 128, 6426–6436 CrossRef CAS PubMed.
  25. J. W. Stouwdam and F. van Veggel, Nano Lett., 2002, 2, 733–737 CrossRef CAS.
  26. F. Vetrone, V. Mahalingam and J. A. Capobianco, Chem. Mater., 2009, 21, 1847–1851 CrossRef CAS.
  27. L. Mao, Z. Lu, N. He, L. Zhang, Y. Deng and D. Duan, Sci. China: Chem., 2016, 60, 157–162 CrossRef.
  28. B. Dong, H. W. Song, H. Q. Yu, H. Zhang, R. F. Qin, X. Bai, G. H. Pan, S. Z. Lu, F. Wang, L. B. Fan and Q. L. Dai, J. Phys. Chem. C, 2008, 112, 1435–1440 CrossRef CAS.
  29. H. X. Mai, Y. W. Zhang, R. Si, Z. G. Yan, L. D. Sun, L. P. You and C. H. Yan, J. Am. Chem. Soc., 2006, 128, 6426–6436 CrossRef CAS PubMed.
  30. G. S. Yi, H. C. Lu, S. Y. Zhao, G. Yue, W. J. Yang, D. P. Chen and L. H. Guo, Nano Lett., 2004, 4, 2191–2196 CrossRef CAS.
  31. M. Calil Júnior, Á. Melo, E. Rodrigues, F. Sigoli and M. Rodrigues, J. Braz. Chem. Soc., 2017, 28, 1816–1821 Search PubMed.
  32. X. P. Chen, Q. Y. Zhang, C. H. Yang, D. D. Chen and C. Zhao, Spectrochim. Acta, Part A, 2009, 74, 441–445 CrossRef CAS PubMed.
  33. Y. Luo, R. Yang, X. Zhang, B. Hu, S. Hu, L. Zhou and J. Yang, CrystEngComm, 2015, 17, 7762–7771 RSC.
  34. X. Zhang, H. Zou, C. Xu, Z. An, R. Dong, K. Zheng, Y. Sheng and Y. Song, Opt. Mater., 2019, 89, 512–520 CrossRef CAS.
  35. A. R. Kadam and S. J. Dhoble, J. Alloys Compd., 2021, 884, 161138 CrossRef CAS.
  36. Z. Wang, L. Cheng, H. Tang, X. Yu, J. Xie, X. Mi, Q. Liu and X. Zhang, J. Solid State Chem., 2021, 301, 122295 CrossRef CAS.
  37. Q. Han, A. Hao, W. Gao, J. Qi, Y. Wang and J. Dong, Ceram. Int., 2020, 46, 11132–11136 CrossRef CAS.
  38. X. Wang, S. Zhao, Y. Zhang and G. Sheng, J. Rare Earths, 2010, 28, 222–224 CrossRef.
  39. V. Wang, N. Xu, J.-C. Liu, G. Tang and W.-T. Geng, Comput. Phys. Commun., 2021, 267, 108033 CrossRef CAS.
  40. M. G. Brik, A. M. Srivastava and A. I. Popov, Opt. Mater., 2022, 127, 112276 CrossRef CAS.
  41. B. R. Judd, Phys. Rev., 1962, 127, 750–761 CrossRef CAS.
  42. G. S. Ofelt, J. Chem. Phys., 1962, 37, 511–520 CrossRef CAS.
  43. Y. Sato, H. Kato, M. Kobayashi, T. Masaki, D. H. Yoon and M. Kakihana, Angew. Chem., Int. Ed. Engl., 2014, 53, 7756–7759 CrossRef CAS PubMed.
  44. K. Li, J. Fan, M. Shang, H. Lian and J. Lin, J. Mater. Chem. C, 2015, 3, 9989–9998 RSC.
  45. C. Cao, H. K. Yang, J. W. Chung, B. K. Moon, B. C. Choi, J. H. Jeong and K. H. Kim, J. Mater. Chem., 2011, 21, 10342–10347 RSC.
  46. J. Tian and W. Zhuang, Inorg. Chem. Front., 2021, 8, 4933–4954 RSC.
  47. X. Wang, Z. Zhao, Q. Wu, C. Wang, Q. Wang, L. Yanyan and Y. Wang, J. Mater. Chem. C, 2016, 4, 8795–8801 RSC.
  48. V. I. Nefedov, Y. V. Salyn, G. Leonhardt and R. Scheibe, J. Electron Spectrosc. Relat. Phenom., 1977, 10, 121–124 CrossRef CAS.
  49. H. Seyama, J. Chem. Soc., Faraday Trans. 1, 1985, 81, 485–495 RSC.
  50. R. P. Vasquez, M. C. Foote and B. D. Hunt, J. Appl. Phys., 1989, 66, 4866–4877 CrossRef CAS.
  51. U. Yoshinori, T. Akira, I. Toshio and Y. Hideo, Bull. Chem. Soc. Jpn., 1986, 59, 2263–2267 CrossRef.
  52. K. Ye, K. Li, Y. Lu, Z. Guo, N. Ni, H. Liu, Y. Huang, H. Ji and P. Wang, TrAC, Trends Anal. Chem., 2019, 116, 102–108 CrossRef CAS.
  53. J. R. B. J. A. Weil, Electron Paramagnetic Resonance: Elementary Theory Electron and Practical applications, John Wiley & Sons, Hoboken, 2nd edn, 2007 Search PubMed.
  54. L. Jia, M. Gu, G. Song and J. Zhu, Opt. Mater., 2019, 93, 15–18 CrossRef CAS.
  55. S. Fang, T. Lang, T. Han, J. Wang, J. Yang, S. Cao, L. Peng, B. Liu, A. N. Yakovlev and V. I. Korepanov, Chem. Eng. J., 2020, 389, 124297 CrossRef CAS.
  56. G. Zhu, Z. Li, C. Wang, X. Wang, F. Zhou, M. Gao, S. Xin and Y. Wang, Dalton Trans., 2019, 48, 1624–1632 RSC.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2ra05674g

This journal is © The Royal Society of Chemistry 2022