Next Article in Journal
In Situ Test Study on Freezing Scheme of Freeze-Sealing Pipe Roof Applied to the Gongbei Tunnel in the Hong Kong-Zhuhai-Macau Bridge
Next Article in Special Issue
Implementation of a Motor Diagnosis System for Rotor Failure Using Genetic Algorithm and Fuzzy Classification
Previous Article in Journal
Reverse Circulation Drilling Method Based on a Supersonic Nozzle for Dust Control
Previous Article in Special Issue
Estimation of Noise Magnitude for Speech Denoising Using Minima-Controlled-Recursive-Averaging Algorithm Adapted by Harmonic Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of the Concentration of Eu3+ Ions and Synthesizing Temperature on the Luminescence Properties of Sr2−xEuxZnMoO6 Phosphors

1
Department of Aero-Electronic Engineering, Air Force Institute of Technology, Kaohsiung 820, Taiwan, R.O.C.
2
Department of Electronic Engineering, National Kaohsiung University of Applied Sciences, Kaohsiung 807, Taiwan, R.O.C.
3
Department of Aeronautics and Astronautics, Air Force Academy, Kaohsiung 820, Taiwan, R.O.C.
4
Department of Chemical and Material Engineering, National University of Kaohsiung, Kaohsiung 811, Taiwan, R.O.C.
*
Author to whom correspondence should be addressed.
Appl. Sci. 2017, 7(1), 30; https://doi.org/10.3390/app7010030
Submission received: 13 September 2016 / Revised: 5 December 2016 / Accepted: 22 December 2016 / Published: 27 December 2016

Abstract

:
The effect of Eu2O3 concentration on the luminescence properties of double perovskite (cubic) Sr2−xEuxZnMoO6 phosphors was thoroughly investigated using different synthesizing temperatures. Phosphors with the composition Sr2−xEuxZnMoO6, where Eu2O3 was substituted for SrO and x was changed from 0 to 0.12, were synthesized by the solid-state method at temperatures of 900–1200 °C, respectively. Analysis of the X-ray diffraction (XRD) patterns showed that even when the synthesizing temperature was 1100 °C, secondary or unknown phases were observed in Sr2−xEuxZnMoO6 ceramic powders. The effect of the concentration of Eu3+ ions on the luminescence properties of the Sr2−xEuxZnMoO6 phosphors was readily observable because no characteristic emission peak was observed in the Sr2ZnMoO6 phosphor. Two characteristic emission peaks at 597 and 616 nm were observed, which correspond to the 5D07F1 and 5D07F2 transitions of Eu3+ ions, respectively. The two characteristic emission peaks of the Sr2−xEuxZnMoO6 phosphors were apparently influenced by the synthesizing temperature and the concentration of Eu3+ ions. When x was larger than 0.08, a concentration quenching effect of Eu3+ ions in the Sr2−xEuxZnMoO6 phosphors could be observed. The lifetime of the Sr2−xEuxZnMoO6 phosphors decreased as the synthesizing temperature increased. A linear relation between temperature and lifetime was obtained by using a fitting curve of t = −0.0016 × T + 3.543, where t was lifetime and T was synthesizing temperature.

1. Introduction

UV light can be generated as a consequence of electronic transitions of light sources through an Hg discharge. In low-pressure Hg discharge, the main emission line is located at a wavelength of 254 nm. This light is invisible and harmful to human bodies, so it has to be converted into visible light, which can be done with a combination of luminescent materials. These luminescent materials can strongly absorb light of that wavelength and efficiently convert it into visible light. Recently, white light emitting diodes (LEDs) have become popular because they have several advantages, including high efficiency, long lifetime, and low power consumption. Red phosphors are also helpful for generating white light when they are excited by blue or near-UV lights. To obtain phosphors with highly efficient emissions, it is important to choose the right compound materials and ensure they have outstanding physical and chemical stability. Numerous studies have explored different luminescent materials to enable the development of suitable phosphors. When lanthanide contraction ions are introduced into host materials, unfilled 4fN electron orbitals result, and these have attracted considerable attention. The resulting phosphors emit very luminescent emissions with specific light wavelengths because of variations in the energy level of some free electrons [1,2,3,4]. These phosphors have been the most promising candidates for applications in fluorescent lamps and flat panel display devices, such as electroluminescence panels, plasma display panels, and field emission displays.
A large number of isotropic compounds with perovskite structures, such as double perovskite oxides, have the general formula A2BB’O6, in which BO6 and B’O6 octahedra are corner-shared, alternately. The great flexibility of A and B(B’) sites in A2BB’O6 allows very rich substitutions, and this framework forms cube-octahedral cavities filled by A-site cations [5,6]. Double perovskites with the formula A2BB’O6, where A uses an alkaline earth, B and B’ are metal transition magnetic and nonmagnetic ions, and O is oxygen, have been investigated as magnetic materials for many years. For example, Sr2CrMoO6 has been studied as a half-metallic system [7]. Kobayashi et al. recently reported room-temperature low-field magnetic resistance in the ordered double perovskite Sr2FeMoO6 [8]. Complete ordering of Fe and Mo on the B and B’ sites of this metallic A2BB’O6 double perovskite is predicted to give half-metallic ferromagnetism with localized majority-spin electrons on the Fe atoms [9,10]. Recently, the study of A2BB’O6-based materials has increased due to various technological applications, such as inorganic oxide luminescent materials.
The emitting materials are usually composed of activators and a host lattice. Some host lattice materials can produce light themselves, and some can produce light when doped with rare-earth activators (ions). Rare-earth ions are known to exist in various valence states, although the trivalent state is the most prevalent. Rare-earth ions can be applied in lighting devices and display panels due to their abundant energy levels across a wide spectrum range, from ultraviolet to near infrared. Sm- and Eu-based ions are the most commonly used dopants because they are stable in trivalent (Sm3+ and Eu3+), as well as divalent (Eu2+), states. The luminescence of rare-earth ions doped in perovskite-type ceramics was actively investigated in the 1960s and 1970s because of interest in their ferroelectricity, phase transitions, and semiconducting properties [11]. Recently, many studies have shown that the double perovskite structure with a composition of A2BMO6 (A = Ba, Sr; B = Ca, Zn; M = Mo, W) is activated by trivalent europium ions (Eu3+) [12,13,14,15]. Phosphors activated by Eu3+ are considered ideal red sources because of their sharp emission lines in the red region [12,13,14,15,16]. Eu3+-doped double-perovskite materials have a broad excitation band ranging from UV to visible light, and they also show highly efficient red luminescence. For that, Eu3+-doped double molybdenum (Mo)-based double perovskite oxides have attracted significant attention for their possible application as luminescent materials, such as Sr2MgMoxW1−xO6: Eu3+ [17], Sr2Ca(Mo/W)O6: Eu3+ [18], Sr2CaMoO6: Eu3+ [19], (Ba,Sr)2CaMoO6: Eu3+, Yb3+ [20], Ca2LaMO6: Eu3+ (M = Sb, Nb, Ta) [21], and A2CaMoO6 (A = Sr, Ba) [6], respectively.
To the best of our knowledge, the luminescent characteristics of concentrated Eu3+ ions in Sr2−xEuxZnMoO6 phosphors, which are molybdenum-based double-perovskite oxides, have not been reported. In this study, the red-emitting phosphors of Eu3+-doped Mo-based double-perovskite Sr2−xEuxZnMoO6 oxides were synthesized by the conventional high-temperature solid state reaction method. We found that Eu3+ concentration and the synthesizing temperature of the Sr2−xEuxZnMoO6 phosphors had a large effect on their luminescent characteristics. The concentration quenching effect of Eu3+ ions in Sr2−xEuxZnMoO6 phosphors was found and would be well discussed [22]. In this study, the first important novelty is that we have not found any similar studies about Sr2−xEuxZnMoO6 phosphors. The second important novelty is that we found a linear relationship between the synthesizing temperature and lifetime of Sr2−xEuxZnMoO6 phosphors. We also investigated the relationship between the synthesized temperature and lifetime of Sr2−xEuxZnMoO6 phosphors.

2. Materials and Methods

Sr2−xEuxZnMoO6 powders were synthesized through the solid-state reaction method. Stoichiometric amounts of SrCO3, ZnO, MoO3, and Eu2O3 were weighed according to the composition formula of (2 − x) SrCO3 + ZnO + MoO3 + (0.5 x) Eu2O3, where x = 0, 0.04, 0.06, 0.08, 0.10, and 0.12, respectively. After being mixed in acetone, dried, and ground, the solid-state reaction method was used to heat the Sr2−xEuxZnMoO6 compositions in an air atmosphere. The Sr2−xEuxZnMoO6 powders were heated to 900 °C, 1000 °C, 1100 °C, and 1200 °C for 4 h. When the synthesizing temperature was equal to 1300 °C, the Sr2−xEuxZnMoO6 was melted and gathered together to not be used as part of the phosphors. The crystalline structures of the synthesized Sr2−xEuxZnMoO6 powders were measured using X-ray diffraction (XRD) (Bruker, Boston, MA, USA ) patterns with Cu Kα radiation (λ = 1.5418 Å) and with a scanning speed of 2° per minute. Photoluminescence (PL) properties were recorded at room temperature in the wavelength range of 450–800 nm on a Hitachi F-4500 fluorescence spectrophotometer (Hitachi, Tokyo, Japan). In the past, we had found that 271 nm had a better excitation effect on BaZr1−xEuxO3 powders [16]. This result suggested that we would also need to find the optimum optical wavelength for exciting the Sr2−xEuxZnMoO6 powders. In this study, the three-dimensional (3D) scanning process using a spectrophotometer (Hitach, Tokyo, Japan) was used to find the optimum optical wavelength, and this value was dependent on the compositions of the Sr2−xEuxZnMoO6 phosphors and the synthesizing temperature. We found that the optimum exciting optical wavelength for all of the Sr2−xEuxZnMoO6 powders was 350 nm, and the Sr2−xEuxZnMoO6 powders excited by other wavelengths had the weaker PL intensities.

3. Results and Discussion

To achieve high PL properties, the preparation of Sr2−xEuxZnMoO6 powders forming the double-perovskite phase is very important, because crystallization of Sr2−xEuxZnMoO6 powders influences their photoluminescent properties. Figure 1 shows the XRD patterns of our Sr2−xEuxZnMoO6 powders as a function of the synthesizing temperature. The strong peaks occurred at around 31.9° for the (220) diffraction peak of the six host lattices. Those results suggest that the XRD patterns showed stable double-perovskite features regardless of the synthesizing temperature and Eu3+ concentration.
As Figure 1a shows, when the synthesizing temperature of the Sr2−xEuxZnMoO6 powders was 900 °C and as the x value increased from 0 to 0.12, the 2θ value of the (220) diffraction peak shifted from 31.85 to 31.87 and the full width at half maximum (FWHM) values for the (220) diffraction peak were in the range of 2θ = 0.25°–0.27°. When the synthesizing temperature was 1200 °C, the 2θ value of the (220) diffraction peak shifted from 31.88 to 31.90, and the FWHM values for the (220) diffraction peak of the Sr2−xEuxZnMoO6 powders were in the range of 2θ = 0.19–0.20°, as the x value of the Sr2−xEuxZnMoO6 powders increased from 0 to 0.12. The results in Figure 1a–d show that the 2θ of the (220) diffraction peak shifted to a higher value and the FWHM values for the (220) diffraction peaks of the Sr2−xEuxZnMoO6 powders decreased as the synthesizing temperature increased.
The ideal cubic double-perovskite structure (with the same space group Pm 3 ¯ m (221)) can be described by a faced-centered cubic (fcc) lattice with lattice constant 2a [23]. The B(B’) ion is coordinated by the B’(B) ion using an O ion as an intermediate in the middle, and the lengths of B–O and B’–O are considered to be equal. After relaxation, both lattice constants and atomic positions reduce the ideal cubic structure (space group Fm 3 ¯ m) to a tetragonal structure (space group I4/mmm). There are two O1 atoms located on the z-axis with B and B’ atoms sitting between, and the four O2 atoms are located on the xy-plane; the same as the B and B’ atoms. The angle of the B–O–B’ remains at 180° during structural optimization, whereas the lattice constant and bond length change. The lattice a can be calculated by using (1) the reflection peaks (011), (111), (200), and (220) from the XRD patterns in Figure 1; and (2) the closeness of the c/a ratio to the ideal value of 2 . As the synthesizing temperature was 900 °C, the Sr2−xEuxZnMoO6 powders exhibited a cubic crystal structure with the cell parameters changing from a = b = c/ 2 = 0.3968 nm to a = b = c/ 2 = 0.3971 nm as the concentration of Eu3+ ions increased from 0 to 0.12. The cell parameters for Sr2−xEuxZnMoO6 powders synthesized at 1200 °C were also calculated from the XRD patterns shown in Figure 1, and the cell parameters changed from a = b = c/ 2 = 0.3971 nm to a = b = c/ 2 = 0.3974 nm as the concentration of Eu3+ ions increased from 0 to 0.12. These results were in good agreement with those from the Joint Committee on Powder Diffraction Standards (JCPDS) file number 742474.
As the results of the X-ray diffraction (XRD) patterns showed in Figure 1 were compared, these results indicated that the diffraction intensities of unknown phases for peaks located at around 2θ = 28.44° and 33.09° increased as the concentration of Eu3+ ions increased and decreased as the synthesizing temperature increased. Those peaks are in good agreement with the (222) and (400) peaks of JCPDS file number 120393 for cubic Eu2O3. The decrease in the diffraction intensities of peaks located at around 2θ = 28.44° and 33.09° prove that more Eu3+ ions will substitute the sites of Sr2+ ions as the synthesizing temperature is raised. Even when the synthesizing temperature was 1100 °C, secondary or unknown phases were observed in the Sr2−xEuxZnMoO6 ceramic powders. These Eu2O3 and secondary or unknown phases were not observed when the synthesizing temperature was 1200 °C. These results suggest that when the same synthesizing temperature is used, the concentration of Eu3+ ions has no apparent effect on the crystallization of Sr2−xEuxZnMoO6 powders; hence, the synthesizing temperature is an important factor in determining the crystalline properties of Sr2−xEuxZnMoO6 powders. Additionally, the concentration of Eu3+ ions and the synthesizing temperature affect the photoluminescent properties of Sr2−xEuxZnMoO6 phosphors.
XRD patterns for Sr2−xEuxZnMoO6 phosphors synthesized at 1200 °C for 4 h in the narrow range of 29–35° are shown in Figure 2. These results are significant. Initially, the splitting of the (220) diffraction peak was observed in the Sr2ZnMoO6 and Sr1.98Eu0.02ZnMoO6 phosphors, but it was not observed in other Sr2−xEuxZnMoO6 phosphors. The two split peaks of the Sr2ZnMoO6 phosphors were located at 2θ = 31.83° and 31.88°, and the two split peaks of the Sr1.98Eu0.02ZnMoO6 phosphors were located at 2θ = 31.81° and 31.89°. These results suggest that the Sr2ZnMoO6 and Sr1.98Eu0.02ZnMoO6 phosphors revealed a perovskite structure with the tetragonal phase. As the concentration of Eu3+ was more than 0.02, the Sr2−xEuxZnMoO6 phosphors would have transformed from the tetragonal phase to the (pseudo-)cubic phase, since the splitting of the (220) diffraction peak was not observed. Since the Sr2−xEuxZnMoO6 phosphors were tetragonal phase or (pseudo-)cubic phase, even when the concentration of Eu3+ ions was increased to 0.12, this result also proves that the Eu3+ ions would have substituted into the sites of Ba2+ ions. If the Eu3+ ions had substituted into the sites of Zn2+ (or Mo) ions, the Sr2−xEuxZnMoO6 phosphors would have revealed other crystalline phases, or more secondary phases would have been revealed in the Sr2−xEuxZnMoO6 phosphors rather than in the double-perovskite features.
Previously, Lin et al. successfully synthesized novel near-UV and blue-excited Eu3+, Tb3+-co-doped one-dimensional strontium germanate full-color nanophosphors by a simple sol-hydrothermal method. They found that incorporation of the Eu3+ and Tb3+ ions into strontium germanate resulted in a slight shrinkage of the lattice constants and the unit cell volume because Eu3+ and Tb3+ had smaller radii than Sr2+, indicating the Eu3+ and Tb3+ ions had been incorporated into the host lattice of SrGe4O9 and did not change the crystal structure [24]. Those results also suggested that the Eu3+ ions substituted into the sites of Sr2+ ions. The 2θ value of the (220) diffraction peak was shifted to a higher value, as the concentration of Eu3+ was equal to, or greater than, 0.02. The ion radius of Sr2+ is 0.130 nm and the ion radius of Eu3+ is 0.1087 nm. Hence, as more Sr2+ would have been substituted by Eu3+, the ionic radius of the Sr2−xEuxZnMoO6 phosphors would have decreased, increasing the 2θ value of the (220) diffraction peak. The ion radius of Eu2+ is 0.131 nm, which is thought to be the same as the ion radius of Sr2+. If Eu2O3 exists as Eu2+ ions, the cell parameters of the Sr2−xEuxZnMoO6 phosphors would not have been changed. The shift of the (220) diffraction peak to a higher 2θ value or the decrease in the cell parameters can prove that Eu2O3 existed in the Sr2−xEuxZnMoO6 phosphors and in the Eu3+ state.
In order to find the best doping concentration of Eu3+, we synthesized a series of Sr2−xEuxZnMoO6 phosphors (x = 0 to 0.12) and measured their emission spectra. The PL emission spectra of Sr2−xEuxZnMoO6 powders excited at a wavelength of 350 nm are shown in Figure 3 for the light wavelength range of 400–650 nm. The spectra in Figure 3 show that that all the excitation spectra of Sr2−xEuxZnMoO6 phosphors (except the Sr2ZnMoO6 powder) consisted of two parts: one broad band in the 400–575 nm region and two sharp peaks located at 597 nm and 616 nm. However, for the Sr2ZnMoO6 powder, even when the synthesizing temperature was 1200 °C, the two sharp peaks of Sr2−xEuxZnMoO6 phosphors located at 597 nm and 616 nm were not found in the emission spectra. This means that no characteristic peaks were observed in Sr2ZnMoO6 powder even when the synthesizing temperature was 1200 °C. Obviously, the broad band in the 400–575 nm region is assignable to the well-known O2–Mo6+ charge transfer band (CTB) [25]. As Eu2O3 was substituted for SrO, the Sr2−xEuxZnMoO6 phosphors had the same spectral profile, but with a different concentration of Eu3+ ions, the characteristic peaks were observed.
When the synthesizing temperature was changed from 900 to 1200 °C, the emission spectra of the Sr2−xEuxZnMoO6 phosphors consisted of sharp peaks in two strong bands at 597 and 616 nm, corresponding to the 5D07F1 (597 nm) and 5D07F2 (616 nm) transitions of Eu3+ ions. These results prove that the two emission peaks of the Sr2−xEuxZnMoO6 phosphors, 5D07F1 at 597 nm and 5D07F2 at 616 nm, were excited by the addition of Eu3+ ions. When the synthesizing temperature was raised to 1100 and 1200 °C, the intensity of the broadened emission went from 400 nm to 575 nm, increasing with the increase in the Eu3+ ion concentration. However, the emission intensities of the Sr2−xEuxZnMoO6 phosphors were influenced by the synthesizing temperature and the concentration of the Eu3+ ions.
In a previous study, the spectra of BaZrO3 doped with Eu3+ powders consisted of a series of resolved emission peaks located at 576 nm, 597 nm, 616 nm, 623 nm, 651 nm, 673 nm, 696 nm, and 704 nm, which are assignable to the 5D07FJ (J = 0, 1, 2, 3, 4) transitions of Eu3+ ions, namely, 5D07F0 (576 nm), 5D07F1 (597 nm), 5D07F2 (616 nm, 623 nm), 5D07F3 (651 nm), and 5D07F4 (673 nm, 696 nm, and 704 nm) [16]. Liu and Wang’s research also showed that the emission intensities of 5D07F0 (576 nm), 5D07F1 (597 nm), and 5D07F2 (616, 623 nm) had almost the same values, even if the Eu concentration in the BaZr1−xEuxO3 phosphor powders was different [26]. These previous results [6,16], and the results this study, suggest that the transitions of Eu3+ ions between different energy bands are affected by the host materials of the prepared phosphors.
As the synthesizing temperature was changed from 900 °C to 1100 °C, the intensities of the two emission peaks of the Sr2−xEuxZnMoO6 phosphors was enhanced by increasing the Eu3+ doping concentration and reached a maximum value at x = 0.08. In contrast, the intensities of the two emission peaks of the Sr2−xEuxZnMoO6 phosphors decreased when the Eu3+ doping ratio was more than 0.08. This proves that the concentration-quenching effect of Eu3+ doping happened in the Sr2−xEuxZnMoO6 phosphors. When 1200 °C was used as the sintering temperature, the intensities of the two emission peaks of the Sr2−xEuxZnMoO6 phosphors increased as the concentration of Eu3+ ions rose. The results in Figure 3 also show that when the temperature changed from 1100 °C to 1200 °C, the emission peak varied significantly. For the Sr2−xEuxZnMoO6 phosphors, when the synthesizing temperature was changed from 1100 °C to 1300 °C, the emission peak with the maximum intensity shifted from 5D07F1 (597 nm) to 5D07F2 (616 nm). Nevertheless, the strong band at 576 nm corresponding to the 5D07F0 transition was not observed in the emission spectra.
The maximum emission intensities (PLmax values) of the Sr2−xEuxZnMoO6 phosphors in the transitions of the 5D07F1 (597 nm) and 5D07F2 (616 nm) peaks are presented in Figure 4 as a function of the synthesizing temperature and Eu3+ ion concentration. Those results suggest again that the PL characteristics of the Sr2−xEuxZnMoO6 phosphors were strongly affected by the synthesizing temperature and the concentration of Eu3+ ions. As the x value of the Sr2−xEuxZnMoO6 phosphors was smaller than 0.10, the emission intensities of the 5D07F1 (597 nm) and 5D07F2 (616 nm) peaks first increased, reached a maximum, and then decreased as the synthesizing temperature increased. When the x values of Sr2−xEuxZnMoO6 phosphors were 0.10 and 0.12, the emission intensity of the 5D07F1 (597 nm) first increased, then decreased at 1100 °C, and then increased at 1200 °C. The emission intensity of the 5D07F2 (616 nm) peaks increased with the increase in synthesizing temperature. These results suggest that 1100 °C is an important synthesizing temperature for Sr2−xEuxZnMoO6 phosphors because the transition of PL properties happens at this temperature, but the reasons for this are not known.
The results in Figure 3 and Figure 4 present an important result regarding Sr2−xEuxZnMoO6 phosphors: the PLmax values of the transition of the 5D07F1 (597 nm) critically increased as the synthesizing temperature increased from 900 °C to 1000 °C, then they did not apparently increase as the synthesizing temperature increased from 1000 °C to 1100 °C, as Figure 4a shows. However, for Sr2−xEuxZnMoO6 phosphors with x = 0.04, 0.06, and 0.08, the PLmax values of the transition of the 5D07F2 (616 nm) linearly increased as the synthesizing temperature increased from 900 °C to 1100 °C. For Sr2−xEuxZnMoO6 phosphors with x = 0.10 and 0.12, the PLmax values of the transition of the 5D07F2 (616 nm) linearly increased as the synthesizing temperature increased from 900 °C to 1200 °C, as Figure 4b shows. When the synthesizing temperature was 1200 °C, the emission intensity of the transition of 5D07F2 (616 nm) for Sr1.9Eu0.1ZnMoO6 and Sr1.88Eu0.12ZnMoO6 phosphors was higher than that of 5D07F1 (597 nm), as Figure 4 shows.
The transitions of 5D07F1 (588 nm), 5D07F2 (612 nm), 5D07F3 (649 nm), and 5D07F4 (695 nm) for the Eu3+ ions can be simultaneously observed in the emission spectrum of the Eu3+, Tb3+-co-doped one-dimensional strontium germinate, full-color nano-phosphors [24]. However, introduction of the Eu3+ ions in the Sr2−xEuxZnMoO6 lattice results in disorder. The disorder in the structure will cause transitions in 5D07F2 and point defects in the lattice because of differences in the chemical valence and in the ionic radius between Eu3+ ions and Sr2+ ions. The Eu3+ ions could occupy either of the sites of Sr2+ or the sites of Zn2+ (Mo2+) in the Sr2−xEuxZnMoO6 phosphor’s double-perovskite structure. If Eu3+ ions occupy a lattice site with a strict center of symmetry, the odd terms of the static crystal field vanish. This will lead to electric dipole transitions being strictly forbidden for purely electric transitions and the transitions of the 5D07F1 (597 nm) will predominantly produce the emission.
The Eu3+ ions could occupy either the tetrahedral sites of Sr2+ or the octahedral sites of Zn2+ or Mo6+ in the Sr2−xEuxZnMoO6 double-perovskite crystal structure. Consequently, as both of these sites are centrosymmetric, electric dipole transitions (5D07Fj=0,2, only the 5D07F2 transition is observed in this study) are forbidden and, as such, should not appear in the spectra. As the synthesizing temperature is 900 °C, the fact that the 5D07F2 electric dipole transition at 615 nm prevails over the 5D07F1 magnetic dipole transition at 595 nm in the case of Sr2−xEuxZnMoO6 phosphors leads us to the conclusion that the Eu3+ ions are generally located in a disordered manner in the Sr2−xEuxZnMoO6 powders [27]. The charge-compensating oxygen vacancies surrounding the Eu3+ ions will lead to the deviation from the point symmetry and relaxation of electric dipole transitions selection rules, with the appearance of the 5D07F2 transition lines in the spectra [27]. Additionally, the Eu3+ luminescent centers contributing to the 5D07F2 transition line are probably located at the Sr2−xEuxZnMoO6 particle surface or subsurface and will dominate the emission of Sr2−xEuxZnMoO6 phosphors when the synthesizing temperature is 1200 °C [27]. If Eu3+ ions occupy a non-centrosymmetric lattice site, both electric and magnetic transitions are possible, and the non-symmetric site occupancy will be a dominant reason for causing the emission of 5D07F2 (612 nm) transition [28].
The definition of fluorescence (or phosphorescence) lifetime is the time for the intensity of a single emission peak to decrease to 1/e (approximately 37%) of its original intensity. The decay curves of Sr2−xEuxZnMoO6 phosphors were obtained at different concentrations of Eu3+ (x = 0.04, 0.06, 0.08, 0.010, and 0.12), and Figure 5 shows the intensity decay of the luminescence of Sr2−xEuxZnMoO6 phosphors as a function of the synthesizing temperature, which can be used to calculate the lifetime. The wavelength of the exciting light was 271 nm, and the measured wavelength for the intensity decay was 597 nm (5D07F1), because it had the maximum emission intensity, as the synthesizing temperature was 900–1100 °C. Typically, the standard form of the single exponential decay function is:
A(t) = A0 exp[-(t/τ)]
where A0 is the initial population, τ is constant for the decay time , and t is time. Figure 5 shows that the all curves of decay time can be fitted well by a single-exponential function. Figure 5 also shows that if the same synthesizing temperature is used, the observed decay time is almost unchanged even the x value in Sr2−xEuxZnMoO6 phosphor increases from 0.04 to 0.12. Moreover, the decay time is a single exponential function for Sr2−xEuxZnMoO6 phosphors with different concentrations of Eu3+ ions and synthesized at different temperatures, because the activator lies in the same coordination environment [29].
From the results shown in Figure 5, the lifetimes of Sr2−xEuxZnMoO6 phosphors, which are compared in Table 1, were measured at 2.04–2.10 ms, 1.93–2.03 ms, 1.78–1.82 ms, and 1.65–1.67 ms, when the synthesizing temperature was 900 °C, 1000 °C, 1100 °C, and 1200 °C, respectively. The lifetimes of Sr2−xEuxZnMoO6 phosphors apparently decreased with the increase in the synthesizing temperature. When the synthesizing temperature was 900 °C and 1000 °C, the lifetimes of the Sr2−xEuxZnMoO6 phosphors varied more. When the synthesizing temperature was 1100 °C and 1200 °C, the Sr2−xEuxZnMoO6 phosphors had less variation in their lifetimes. In Figure 6 we use a fitting curve to find the relationship between lifetime and synthesizing temperature; the equation is:
t = −0.0016 × T + 3.543
where t represents the emitting lifetime of the luminescent material and T represents the synthesizing temperature (°C). In other words, the luminescence lifetime of Sr2−xEuxZnMoO6 phosphors can be obtained by using Equation (2), as the synthesizing temperature is changed from 900 °C to 1200 °C. These results suggest that the luminescence lifetime of Sr2−xEuxZnMoO6 phosphors can be a function of synthesizing temperature, and that the concentration of Eu3+ ions not only can affect their emitting intensity but also can affect their luminescence lifetime.

4. Conclusions

For Sr2−xEuxZnMoO6 powders, the 2θ value of the (220) diffraction peak shifted to a higher value and the FWHM value for the (220) diffraction peak decreased as the synthesizing temperature increased. When the x value (the concentration of Eu3+ ions) increased from 0 to 0.12 and the synthesizing temperatures rose from 900 °C to 1200 °C, the cell parameters of the Sr2−xEuxZnMoO6 powders were changed from a = b = c/ 2 = 0.3968 nm to a = b = c/ 2 = 0.3971 nm and from a = b = c/ 2 = 0.3971 nm to a = b = c/ 2 = 0.3974 nm (cubic crystal structure). When the synthesizing temperature was 900–1100 °C, the emission intensities of the Sr2−xEuxZnMoO6 phosphors peaked at x = 0.08 and decreased as the concentration of Eu+3 ions increased. There was an observable concentration-quenching effect of Eu3+ ions in the Sr2−xEuxZnMoO6 phosphors. The emission spectrum of Sr2ZnMoO6 powder consisted of one broad band in the 400–575 nm region, and the emission spectra of the Sr2−xEuxZnMoO6 phosphors (except for the Sr2ZnMoO6 powder) consisted of two parts: one broad band in the 400–575 nm region and two sharp peaks located at 597 nm and 616 nm. The lifetimes of the Sr2−xEuxZnMoO6 phosphors, which depended on synthesizing temperature and were independent of the concentration of Eu3+ ions, were 2.04–2.10 ms, 1.93–2.03 ms, 1.78–1.82 ms, and 1.65–1.67 ms when the synthesizing temperature was 900 °C, 1000 °C, 1100 °C, and 1200 °C, respectively. We found that when the synthesizing temperature was in the range of 900–1200 °C, the relationship between the lifetime and the synthesizing temperature could be determined by the following equation: t = −0.0016 × T + 3.543, where t and T represent the emitting lifetime of the luminescent material and the synthesizing temperature (°C).

Acknowledgments

The authors would like to acknowledge the financial supports of MOST 104-2221-E-390-013-MY2, MOST 105-2622-E-390-003-CC3, and MOST 105-2221-E-344-002.

Author Contributions

S.-H.Y. and J.-L.L. helped proceeding the experimental processes, measurements, and data analysis; C.-Y.L. and C.-F.Y. organized the paper and encouraged in paper writing; Also, C.-Y.L. and C.-F.Y. helped proceeding the experimental processes and measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Binnemans, K.; Görller-Walrand, C. On the color of the trivalent lanthanide ions. Chem. Phys. Lett. 1995, 235, 163–174. [Google Scholar] [CrossRef]
  2. Dorenbos, P. The 5d level of positions of the trivalent lanthanides in inorganic compounds. J. Lumin. 2000, 91, 155–176. [Google Scholar] [CrossRef]
  3. Reid, M.F.; Pieterson, L.V.; Meijerink, A. Trends in parameters for 4fN ↔ 4fN−15d spectra of lanthanide ions in crystals. J. Alloys Compd. 2002, 344, 240–245. [Google Scholar] [CrossRef]
  4. Dorenbos, P. Energy of the first 4f7 ↔ 4f65d transition of Eu2+ in inorganic compounds. J. Lumin. 2003, 104, 239–260. [Google Scholar] [CrossRef]
  5. Anderson, M.T.; Greenwood, K.B.; Taylor, G.A.; Poeppelmeier, K.R. B-cation arrangements in double perovskites. Prog. Solid State Chem. 1993, 22, 197–233. [Google Scholar] [CrossRef]
  6. Sivakumar, V.; Varadaraju, U.V. A Promising Orange-Red Phosphor Under Near UV Excitation. Electrochem. Solid-State Lett. 2006, 9, H35–H38. [Google Scholar] [CrossRef]
  7. Bonilla, C.M.; Landinez Tellez, D.A.; Arbey Rodriguez, J.; Vera Lopez, E.; Roa-Rojas, J. Half-metallic behavior and electronic structure of Sr2CrMoO6 magnetic system. Physica B 2007, 398, 208–211. [Google Scholar] [CrossRef]
  8. Kobayashi, K.-I.; Kimura, T.; Sawada, H.; Terakura, K.; Tokura, Y. Room-temperature magnetoresistance in an oxide material with an ordered double-perovskite structure. Nature 1998, 395, 677–680. [Google Scholar] [CrossRef]
  9. Yin, H.Q.; Zhou, J.-S.; Zhou, J.-P.; Dass, R.; McDevitt, J.T.; Goodenough, J.B. Intra-versus intergranular low-field magnetoresistance of Sr2FeMoO6 thin films. Appl. Phys. Lett. 1999, 75, 2812–2814. [Google Scholar] [CrossRef]
  10. Jalili, H.; Heinig, N.F.; Leung, K.T. Growth evolution of laser-ablated Sr2FeMoO6 nanostructured films: Effects of substrate-induced strain on the surface morphology and film quality. J. Chem. Phys. 2010, 132, 204701. [Google Scholar] [CrossRef] [PubMed]
  11. Makishima, S.; Yamamoto, H.; Tomotsu, T.; Shionoya, S.J. Luminescence spectra of Sm3+ BaTiO3 host lattice. J. Phys. Soc. Jpn. 1965, 20, 2147–2151. [Google Scholar] [CrossRef]
  12. Li, Y.; Liu, X. Sol–gel synthesis, structure and luminescence properties of Ba2ZnMoO6:Eu3+ phosphors. Mater. Res. Bull. 2015, 64, 88–92. [Google Scholar] [CrossRef]
  13. Sivakumar, V.; Varadaraju, U.V. Synthesis, phase transition and photoluminescence studies on Eu3+-substituted double perovskites—A novel orange-red phosphor for solid-state lighting. J. Solid State Chem. 2008, 181, 3344–3351. [Google Scholar] [CrossRef]
  14. Sivakumar, V.; Varadaraju, U.V. An orange-red phosphor under near-UV excitation for white light emitting diodes. J. Electrochem. Soc. 2007, 154, J28–J31. [Google Scholar] [CrossRef]
  15. Ye, S.; Wang, C.H.; Liu, Z.S.; Lu, J.; Jing, X.P. Photoluminescence and energy transfer of phosphor series Ba2−zSrzCaMo1−yWyO6:Eu,Li, Li for white light UV LED applications. Appl. Phys. B 2008, 91, 551–557. [Google Scholar] [CrossRef]
  16. Chen, K.N.; Hsu, C.M.; Liu, J.; Chiu, Y.T.; Yang, C.F. Effect of Different Heating Process on the Photoluminescence Properties of Perovskite Eu-Doped BaZrO3 Powder. Appl. Sci. 2016, 6, 22. [Google Scholar] [CrossRef]
  17. Li, S.; Wei, X.; Deng, K.; Tian, X.; Qin, Y.; Chen, Y.; Yin, M. A new red-emitting phosphor of Eu3+-doped Sr2MgMoxW1−xO6 for solid state lighting. Curr. Appl. Phys. 2013, 13, 1288–1291. [Google Scholar] [CrossRef]
  18. Ye, S.; Wang, C.H.; Jing, X.P. Photoluminescence and Raman Spectra of Double-Perovskite Sr2Ca(Mo/W)O6 with A- and B-Site Substitutions of Eu3+. J. Electrochem. Soc. 2008, 155, J148–J151. [Google Scholar] [CrossRef]
  19. Zhang, L.; Lu, J.J.; Liu, J.Q.; Li, Y.; Wang, L.X.; Zhang, Q.T. Structure and luminescent properties of Eu3+ doped double perovskite Sr2CaMoO6 orange-red phosphors. Chin. J. Inorg. Chem. 2012, 28, 2036–2042. [Google Scholar]
  20. Ye, S.; Li, Y.; Yu, D.; Yang, Z.; Zhang, Q. Structural effects on Stokes and anti-Stokes luminescence of double-perovskite (Ba,Sr)2CaMoO6:Yb3+,Eu3+. J. Appl. Phys. 2011, 110, 013517. [Google Scholar] [CrossRef]
  21. Zhang, L.; Han, P.; Han, Y.; Lu, Z.; Yang, H.; Wang, L.; Zhang, Q. Structure evolution and tunable luminescence of (Sr0.98−mBamEu0.02)2Ca(Mo1−nWn)O6 phosphor with ultraviolet excitation for white LEDs. J. Alloys Compd. 2013, 558, 229–235. [Google Scholar] [CrossRef]
  22. Wang, Z.J.; Li, P.L.; Yang, Z.P.; Guo, Q.L. Luminescence and concentration quenching of Tb3+ in Sr2ZnMoO6. Opt. Commun. 2014, 321, 100–103. [Google Scholar] [CrossRef]
  23. Liu, Y.P.; Chen, S.H.; Fuh, H.R.; Wang, Y.K. First-Principle Calculations of Half-Metallic Double Perovskite La2BB’O6 (B,B’ = 3d transition metal). Commun. Comput. Phys. 2013, 14, 174–185. [Google Scholar] [CrossRef]
  24. Lin, L.; Sun, X.; Jiang, Y.; He, Y. Sol-hydrothermal synthesis and optical properties of Eu3+, Tb3+-codoped one-dimensional strontium germanate full color nano-phosphors. Nanoscale 2013, 5, 12518–12531. [Google Scholar] [CrossRef] [PubMed]
  25. Li, L.; Shen, J.; Pan, Y.; Zhou, X.; Huang, H.; Chang, W.; Ha, Q.; Wei, X. Enhancing luminescence of Lu2MoO6:Eu3+ phosphors by doping with Li+ ions for near ultraviolet based solid state lighting. Mater. Res. Bull. 2016, 78, 26–30. [Google Scholar] [CrossRef]
  26. Liu, X.; Wang, X. Preparation and luminescence properties of BaZrO3:Eu phosphor powders. Opt. Mater. 2007, 30, 626–629. [Google Scholar] [CrossRef]
  27. Dimitrievska, M.; Ivetic, T.B.; Litvinchuk, A.P.; Fairbrother, A.; Bojan, B.; Miljevic, B.B.; Strbac, G.R.; Rodríguez, A.P.; Lukic-Petrovic, S.R. Eu3+-Doped Wide Band Gap Zn2SnO4 Semiconductor Nanoparticles: Structure and Luminescence. J. Phys. Chem. C 2016, 120, 18887–18894. [Google Scholar] [CrossRef]
  28. Blasse, G.; Bril, A.; Nieuwpoort, W.C. On the Eu3+ Fluorescence in Mixed Metal Oxides. J. Phys. Chem. Solids 1966, 27, 1587–1592. [Google Scholar] [CrossRef]
  29. Xie, A.; Yuan, X.M.; Wang, F.X.; Shi, Y.; Li, J.; Liu, L.; Mu, Z.F. Synthesis and luminescent properties of Eu3+-activated molybdate-based novel red-emitting phosphors for white LEDs. J. Alloys Compd. 2010, 501, 124–129. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Sr2−xEuxZnMoO6 phosphors synthesized at (a) 900 °C; (b) 1000 °C; (c) 1100 °C; and (d) 1200 °C for 4 h, respectively.
Figure 1. XRD patterns of Sr2−xEuxZnMoO6 phosphors synthesized at (a) 900 °C; (b) 1000 °C; (c) 1100 °C; and (d) 1200 °C for 4 h, respectively.
Applsci 07 00030 g001
Figure 2. XRD patterns of Sr2−xEuxZnMoO6 phosphors synthesized at 1200 °C for 4 h.
Figure 2. XRD patterns of Sr2−xEuxZnMoO6 phosphors synthesized at 1200 °C for 4 h.
Applsci 07 00030 g002
Figure 3. Emission spectra of Sr2−xEuxZnMoO6 phosphors synthesized at (a) 900 °C; (b) 1000 °C; (c) 1100 °C; and (d) 1200 °C, respectively.
Figure 3. Emission spectra of Sr2−xEuxZnMoO6 phosphors synthesized at (a) 900 °C; (b) 1000 °C; (c) 1100 °C; and (d) 1200 °C, respectively.
Applsci 07 00030 g003
Figure 4. Emission intensity of Sr2−xEuxZnMoO6 phosphors as a function of synthesizing temperature. (a) Emission peak of 597 nm; and (b) emission peak of 616 nm.
Figure 4. Emission intensity of Sr2−xEuxZnMoO6 phosphors as a function of synthesizing temperature. (a) Emission peak of 597 nm; and (b) emission peak of 616 nm.
Applsci 07 00030 g004
Figure 5. Luminescence lifetime of Sr2−xEuxZnMoO6 phosphors synthesized at (a) 900 °C; (b) 1000 °C; (c) 1100 °C; and (d) 1200 °C, respectively.
Figure 5. Luminescence lifetime of Sr2−xEuxZnMoO6 phosphors synthesized at (a) 900 °C; (b) 1000 °C; (c) 1100 °C; and (d) 1200 °C, respectively.
Applsci 07 00030 g005aApplsci 07 00030 g005b
Figure 6. Fitting curve of lifetime for Sr2−xEuxZnMoO6 phosphors as a function of the synthesizing temperature and concentration of Eu3+ ions.
Figure 6. Fitting curve of lifetime for Sr2−xEuxZnMoO6 phosphors as a function of the synthesizing temperature and concentration of Eu3+ ions.
Applsci 07 00030 g006
Table 1. Luminescence lifetimes of Sr2−xEuxZnMoO6 phosphors as a function of the synthesizing temperature and concentration of Eu3+ ions.
Table 1. Luminescence lifetimes of Sr2−xEuxZnMoO6 phosphors as a function of the synthesizing temperature and concentration of Eu3+ ions.
Synthesizing TemperatureX = 0.04X = 0.06X = 0.08X = 0.10X = 0.12
900 °C2.052.12.042.042.05
1000 °C2.032.031.931.971.99
1100 °C1.811.831.781.821.82
1200 °C1.651.651.661.661.67

Share and Cite

MDPI and ACS Style

Lin, C.-Y.; Yang, S.-H.; Lin, J.-L.; Yang, C.-F. Effects of the Concentration of Eu3+ Ions and Synthesizing Temperature on the Luminescence Properties of Sr2−xEuxZnMoO6 Phosphors. Appl. Sci. 2017, 7, 30. https://doi.org/10.3390/app7010030

AMA Style

Lin C-Y, Yang S-H, Lin J-L, Yang C-F. Effects of the Concentration of Eu3+ Ions and Synthesizing Temperature on the Luminescence Properties of Sr2−xEuxZnMoO6 Phosphors. Applied Sciences. 2017; 7(1):30. https://doi.org/10.3390/app7010030

Chicago/Turabian Style

Lin, Chi-Yu, Su-Hua Yang, Jih-Lung Lin, and Cheng-Fu Yang. 2017. "Effects of the Concentration of Eu3+ Ions and Synthesizing Temperature on the Luminescence Properties of Sr2−xEuxZnMoO6 Phosphors" Applied Sciences 7, no. 1: 30. https://doi.org/10.3390/app7010030

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop