DOI:
10.1039/C4RA11516C
(Paper)
RSC Adv., 2014,
4, 64244-64251
Site occupancy and photoluminescence properties of Eu3+-activated Ba2ZnB2O6 phosphor†
Received
30th September 2014
, Accepted 18th November 2014
First published on 18th November 2014
Abstract
A series of Ba2ZnB2O6:Eu3+ phosphors with a red-emitting band centered at 616 nm were prepared by traditional high temperature solid-state reaction methods. The site-preferred occupancy of Eu3+ in Ba2ZnB2O6 and luminescence properties of Ba2ZnB2O6:Eu3+ were studied combined with X-ray diffraction (XRD), photoluminescence excitation (PLE) spectra and emission (PL) spectra as well as temperature-dependent PL and decay curves. The Rietveld refinements indicate that the Eu3+ ions prefer to occupy Zn (1) (4a) and Zn (2) (4a) sites simultaneously. The PL intensity is improved with increasing Eu3+ content and the optimal dopant content is 0.05. The temperature-dependent PL spectra indicate that the emission intensity decreases with the temperature because of the enhancement of the non-radiative transition. The PL emission intensities of Ba2ZnB2O6:0.05Eu3+ phosphors with Li+, Na+ and K+ as charge compensators are enhanced significantly, and the phosphor compensated by Li+ ions emits the strongest emission. The Commission Internationale de I'Eclairage (CIE) color coordinates of Ba2ZnB2O6:0.05Eu3+ are very close to the CIE of standard red light.
1. Introduction
The borate system comprises more than 500 kinds of structures due to the different combination modes of the B–O groups.1,2 Because of the relative low synthesis temperature, stable physical and chemical properties, and excellent luminescence performance,3,4 borates have been a hot research topic in the field of luminescent host materials recently.
Eu3+ is an important rare earth ion in synthesizing phosphors with good properties, for example Y2O3:Eu3+ is used as a red phosphor in wLEDs. Recently, many types of Eu3+-doped compounds have been reported as interesting candidates for potential red-emitting phosphors.5–11 Because of the proper ionic radii of Ba2+ and Zn2+, a series of barium containing compounds and zinc containing compounds have been used as host materials to be doped by Eu3+, such as Ba1−xB8O13:xEu3+,12 Ba5−2x(VO4)3Cl:xEu3+, xK+,13 Zn1−xAlO4:xEu3+14 and Zn1−x−yB2O4:xBi3+, yEu3+,15 and so on. The radius of Ba2+ ion is larger than that of Eu3+ ion, and Zn2+ ion is smaller than that of Eu3+. However, Eu3+ can occupy Ba2+ sites or Zn2+ sites in different hosts. For example, Eu3+ ions occupy Ba2+ site in Ba1−xB8O13 doped by Eu3+,10 but occupy Zn2+ site in Zn1−x−yB2O4 codoped by Bi3+and Eu3+.15
As one of alkaline earth borates, Ba2ZnB2O6 was firstly synthesized by Robert and Koliha in 1994.16 This compound crystallizes in Pca21 space group with a = 15.068(2), b = 8.720(2), c = 10.128(3) Å, and V = 1330.7 (9) Å3. There are four different Ba2+ ions sites, two different Zn2+ ions sites and three different B3+ ions sites in this compound. Two different groups, vertex-sharing ZnO4 tetrahedra and BO3 triangles, constitute two-dimensional [Zn3B3O6]∞ layers which are perpendicular to the [100] direction in the Ba2ZnB2O6 structure. The layers are linked by the additional BO3 groups, forming a 3D framework. The Ba2+ ions fill the space of the 3D framework to balance the charge. All the BO3 groups in the layers are parallel with each other throughout the structure. The structure of Ba2ZnB2O6 indicates that this compound can offer a proper host structure environment for the doped Eu3+ ions. As mentioned above, the doped Eu3+ can occupy Ba2+ sites or Zn2+ sites in the Ba-based borates or Zn-based borates. However, Ba2ZnB2O6 contains Ba2+ and Zn2+ ions. The site occupancy and the photoluminescence properties of Eu3+ ions in such Ba and Zn co-containing compounds are not studied before. In this study, Eu3+ activated Ba2ZnB2O6 phosphors are prepared and the site occupancy of Eu3+ in Ba2ZnB2O6 is studied as well as the photoluminescence properties of Eu3+ activated Ba2ZnB2O6 phosphors. In order to eliminate the charge unbalance because of Eu3+ doped into Ba2ZnB2O6, the effect of charge compensation is also studied in this paper.
2. Experimental
Polycrystalline samples of Ba2ZnB2O6:Eu3+ and Ba2ZnB2O6:Eu3+, M+ (M = Li, Na and K) were prepared by a solid-state reaction at high temperature starting from analytical purity BaCO3, ZnO, H3BO3 and Eu2O3, Li2CO3, Na2CO3, and K2CO3 (99.99%). The raw materials were weighed out, mixed and ground together in an agate mortar, and then sintered at 600 °C for 24 h in air to remove the H2O and CO2 in a muffle furnace. After cooled down to room temperature, the samples were ground again and further sintered at 800 °C for 72 h in air.
The X-ray diffraction (XRD) data for phase identification of the as-prepared powders were collected in the range of 10° to 80° on a PANalytical X'Pert Pro powder X-ray diffractometer with Cu Kα radiation (40 kV, 40 mA). The XRD for structure refinement were collected over a 2θ range from 10° to 140° at intervals of 0.017° with a counting time of 1 s per step. The photoluminescence (PL) and photoluminescence excitation (PLE) spectra of the samples were measured by a spectrofluorometer (Edinburgh Instruments, FLS920) equipped with a Xe light source and double excitation monochromators. The powder samples were placed into circular sample cells with quartz plates and excited under 45° incidences. The emitted fluorescence was detected by a photomultiplier (R928P) that is perpendicular to the excitation beam. A cutoff filter was used to avoid the influence of the second-order emission of the source radiation. Diffuse reflectance spectra of the phosphors were measured by a UV-visible spectrophotometer (Hitachi U-4100). The lifetimes were recorded using a μF900 lamp (100 W) as a light source and a photomultiplier (R928P) was used as detector. The temperature-dependent luminescence properties were measured on a fluorescence spectrophotometer (F-4600, HITACHI, Japan) with a photomultiplier tube operating at 400 V, and a 150 W Xe lamp used as the excitation lamp.
3. Results and discussion
3.1 Site-preferred occupancy of the Ba2ZnB2O6:xEu3+
The phase purities of Ba2ZnB2O6:xEu3+ (x = 0–0.08) were confirmed by X-ray diffraction (XRD) at room temperature, as shown in Fig. 1. The XRD patterns are found to be in good agreement with that was reported by Smith et al. in 1994,16 which indicates that the structure of host is not changed by doping Eu3+ ions regardless of Eu3+ contents.
 |
| | Fig. 1 XRD patterns of Ba2ZnB2O6:xEu3+ (x = 0.01–0.08). | |
As introduced above, Ba2ZnB2O6 crystallizes in an orthorhombic system with space group Pca21, and offers two types of ions for Eu3+ to replace: four-coordinated Zn2+ ions and six- or seven-coordinated Ba2+ ions. Considering the ionic radius (r) of different coordination number (CN) reported by Shannon,17 the ionic radius changes with the coordination number. The radius of Zn2+ is r = 0.60 Å as CN = 4, and the radius of Ba2+ is r = 1.35 Å as CN = 6 and r = 1.38 Å as CN = 7, respectively. However, the radius of the doped Eu3+ is r = 0.89 Å as CN = 4, r = 0.947 Å as CN = 6, and r = 1.01 Å as CN = 7. Compared with the radii of Zn2+ and Ba2+, the doped Eu3+ will prefer to occupy Zn2+ site. The radius of Ba2+ is much larger than that of Eu3+ (the difference between of these two ions is 0.403 Å as CN = 6 and 0.37 Å as CN = 7), which will result in the crystal structure distortion of Ba2ZnB2O6 too much to be stable in crystallography if the Ba2+ site is occupied by Eu3+. Therefore, it is expected that Eu3+ will not prefer to occupy Ba2+ sites in this compound.
Rietveld refinement was proposed by H. Rietveld in 1967. Refinement of the structure parameters from diffraction data can obtain the crystal structure properties, such as the lattice parameters, the atomic positions of the doped ions, and occupancies.18,19 In order to prove the Eu3+ ions occupy Zn2+ sites in Ba2ZnB2O6:Eu3+, the refinement of the XRD patterns of Ba2ZnB2O6:Eu3+ by Rietveld method20,21 within the Fullprof Program22 were performed. It is found that the Ba2+ sites cannot be occupied because of the final agreement factors are very high and the occupancy of Eu3+ on this site is far away from the nominal doping content. As for Zn2+ sites, it is found that all the doped Eu3+ ions are preferred to occupy Zn (1) (4a) site and Zn (2) (4a) sites during the refinement. Table 1 summarizes the lattice parameters and agreement factors for ZnBi2B2O7:xEu3+ (x = 0.01–0.08) refined by Rietveld method. Fig. 2(a) shows the selected Rietveld refinement plot of Ba2ZnB2O6:0.05Eu3+. Fig. 2(b) shows the structure of Ba2ZnB2O6 and Fig. 2(c) shows two different Zn2+/Eu3+ coordination environment. The final agreement is converged to Rp = 5.97%, Rwp = 7.95%, and Rexp = 3.73%. The refinement results indicate that Ba2ZnB2O6:0.05Eu3+ crystallizes in orthorhombic system with a space group Pca21 and the lattice parameters a = 15.1104(4) Å, b = 8.7218(2) Å, c = 10.1341(3) Å, and cell volume = 1336.010(2) Å3. The longer lattice parameters of Ba2ZnB2O6 doped by Eu3+ is because of the fact that rEu3+ > rZn2+. The atomic positions in the doped Ba2ZnB2O6:0.05Eu3+ unit cell are shown in Table S1.† The refined concentration of Eu3+ is 5.3%, which is in good agreement with the original doping concentration of 5%. The refinement results confirm that Eu3+ ions prefer to occupy Zn 1 (4a) site and Zn 2 (4a) site, which means Eu3+ is coordinated by four O2− with a tetrahedral crystal field environment.
Table 1 Lattice parameters and agreement factors for ZnBi2B2O7:xEu3+ (x = 0.01–0.08) refined by Rietveld method
| Lattice parameter |
X = 0.01 |
X = 0.02 |
X = 0.03 |
X = 0.04 |
X = 0.05 |
X = 0.06 |
X = 0.07 |
X = 0.08 |
| a (Å) |
15.0762(2) |
15.0842(2) |
15.0941(1) |
15.1000(3) |
15.1095(4) |
15.1172(1) |
15.1260(3) |
15.1337(4) |
| b (Å) |
8.7203(2) |
8.7207(1) |
8.7211(1) |
8.7214(2) |
8.7213(1) |
8.7223(2) |
8.7225(1) |
8.7231(2) |
| c (Å) |
10.1290(3) |
10.1305(1) |
10.1317(1) |
10.1324(2) |
10.1333(3) |
10.1354(2) |
10.1360(2) |
10.1377(3) |
| V (Å3) |
1331.61(3) |
1332.67(2) |
1333.69(1) |
1334.35(3) |
1335.57(4) |
1336.42(2) |
1337.31(3) |
1338.30(4) |
| Rp (%) |
5.43 |
5.68 |
5.88 |
5.84 |
5.97 |
6.12 |
6.22 |
6.21 |
| Rwp (%) |
7.57 |
7.67 |
7.80 |
7.76 |
7.95 |
8.22 |
8.37 |
8.40 |
| Rexp (%) |
3.70 |
3.70 |
3.73 |
3.73 |
3.73 |
3.73 |
3.74 |
3.74 |
 |
| | Fig. 2 (a) Rietveld refinement plot on the XRD pattern of Ba2ZnB2O6:0.05Eu3+, small black circles and the red continuous lines represent the experimental and the calculated values respectively; vertical bars (|) indicate the position of Bragg peaks. The blue bottom trace depicts the corresponding residuals between the experimental and the calculated intensity values. (b) Crystal structure projection of Ba2ZnB2O6 along c axis. (c) The coordination environments of two different Zn2+/Eu3+ ions. | |
3.2 Diffuse reflectance spectra
Because the powder of Ba2ZnB2O6:Eu3+ is opaque, diffuse reflectance spectra (DRS) of the parent Ba2ZnB2O6 and doped Ba2ZnB2O6:0.05Eu3+ phosphor were measured instead of absorption spectra. As shown in Fig. 3, the diffuse reflectance spectrum of undoped sample shows obvious decrease from 240 nm to 370 nm near UV-region, which is due to the boron to oxygen charge transition (CT) in Ba2ZnB2O6. When introducing the Eu3+ ions, the Ba2ZnB2O6:0.05Eu3+ sample reveals stronger absorption than the undoped sample in this range. The stronger absorption is possibly caused by the energy transfer from borate groups to Eu3+ ions and CT transition of O2−–Eu3+ in Ba2ZnB2O6:0.05Eu3+ phosphor. In careful observation, the absorption in the range of 390–470 nm has a slight decrease compared with the undoped sample (upper inset in Fig. 3), which is ascribed to the typical 4f–4f transition absorption of Eu3+. The characteristic absorption bands of Eu3+ are not observed besides of the weak band centered at 466 nm (7F0–5D2 transition).
 |
| | Fig. 3 Diffuse reflectance spectra of the parent and doped Ba2ZnB2O6:0.05Eu3+ phosphors. Inset in the upper is the enlarged drawing on the peak centered at 465 nm. Inset in the below shows the absorption spectra (K/S) of Ba2ZnB2O6 and Ba2ZnB2O6:0.05Eu3+ derived with the Kubelka–Munk function. | |
The absorption spectra of undoped and Ba2ZnB2O6:0.05Eu3+ samples are also calculated by the Kubelka–Munk function:23
| | |
F(R) = (1 − R)2/2R = K/S
| (1) |
where
R,
K and
S represent the reflection, absorption, and scattering coefficient, respectively. The absorption spectra are displayed in the below inset of
Fig. 3. The Ba
2ZnB
2O
6:0.05Eu
3+ sample shows strong absorption from 240 nm to 370 nm, which is consistent with the diffuse reflectance spectra.
3.3 PLE and PL spectra
The PLE and PL spectra of Ba2ZnB2O6:0.05Eu3+ are shown in Fig. 4(a). As can be seen in this figure, the PLE spectrum (recorded at 616 nm emission, black line) contains a broad band and a series of narrow bands locating in the range of ∼240–480 nm, which is associated with the absorption spectrum (Fig. 3). The broad band at 240–320 nm originates from the charge transfer (CT) transition of O2−–Eu3+.24 Those narrow bands belong to the typical 4f–4f transition absorption bands of Eu3+, in which the two intense bands centered at 393 nm and 466 nm are attributed to 7F0–5L6 and 7F0–5D2 transition. This demonstrates that the phosphor can be excited efficiently by near-ultraviolet and blue light. The other excitation peaks centered at 319 nm, 362 nm, 376 nm, 383 nm, 398 nm, and 415 nm correspond to the transitions from 7F0 to 5H6, 5D4, 5G3, 5L7, 5L6, and 5D3 levels, respectively.
 |
| | Fig. 4 (a) PLE/PL spectra of the Ba2ZnB2O6:0.05Eu3+ sample. (b) The energy level diagram for Eu3+ in Ba2ZnB2O6. | |
The PL spectrum (excited at 393 nm, red line) in Fig. 4(a) shows a series of sharp bands from 579 nm to 621 nm corresponding to the 5D0–7FJ (J = 0, 1, 2) transitions of Eu3+. According to the magnetic dipole (MD) transition selection rule (ΔJ = 0, ±1), the emission band belonging to 5D0–7F1 transition (ΔJ = 1) is attributed to the magnetic dipole transition and the emission band belonging to 5D0–7F2 is attributed to the electric dipole (ED) transition on the basis of the selection rule of electric dipole transition (ΔJ ≤ 6, when J or J′ = 0, ΔJ = 2, 4, 6).15 The 5D0–7F0 transition is forbidden and sensitive to the crystal field.25 The weak emission peak centered at 579 nm belongs to the 5D0–7F0 transition that is induced by role of crystal field in odd. The 5D0–7F0 transition exists only when Eu3+ occupies sites with local symmetries of Cn, Cnv or Cs. So we can deduce that Eu3+ occupies one site of Cn, Cnv or Cs symmetries.26 The emission bands of Eu3+ in Fig. 4(a) are broadened and split into several lines, indicating that the 7FJ levels appear Stark levels because of crystal field effect.27,28 The 5D0–7F1 transition splits into two emission peaks centered at 590 nm and 594 nm, the 5D0–7F2 transition splits into three emission peaks centered at 610 nm, 616 nm and 621 nm, and the 5D0–7F3 transition splits into three emission peaks centered at 649 nm, 656 nm and 666 nm, respectively. Fig. 4(b) illustrates the partial energy diagram of Eu3+ ions in Ba2ZnB2O6 host. For the excited state of Eu3+, a fast non-radiation (NR) transition happened to the 5D0 level, resulting the 5D3,2,1–7FJ transitions are restrained, so the emission of Eu3+ ions is thought to come from the 5D0–7FJ transitions.29
3.4 Concentration quenching
As seen in Fig. 5(a), the quenching concentration of Ba2ZnB2O6:xEu3+ phosphors is 5 mol%. The emission intensity increases with increasing Eu3+ dopant content when x < 0.05, while it decreases after that doping concentration. Based on the Dexter and Schulman theory,30 concentration quenching is due to energy transfer from one activator to another until an energy sink is reached in the lattice. The concentration quenching in solid systems is because of the role of electric multipole or magnetic dipole interaction. On the basis of energy transfer formula:| | |
I/x = k[1 + β(x)θ/3]−1
| (2) |
where x is the mole fraction of activator ions; k and β are constants; I is the luminous intensity, and θ = 6, 8, 10 corresponding to dipole–dipole, dipole–quadrupole, and quadrupole–quadrupole interactions, respectively. As illustrated in the inset of Fig. 5(b), the relationship of lg(I/x) versus lg(x) is linear and the slope of the line is about −2. So the value of θ is approximately equal to 6, which clearly indicates that the concentration quenching mechanism of Eu3+ in Ba2ZnB2O6:0.05Eu3+ is the dipole–dipole interaction.
 |
| | Fig. 5 (a) PL spectrum of the Ba2ZnB2O6:xEu3+ sample. (b) The emission intensity as a function of Eu3+ concentration in Ba2ZnB2O6:xEu3+. The inset is the dependence of lg(I/x) on lg(x). | |
Blasse31 suggested that the critical distance (Rc) of energy transfer can be expressed by eqn (3):
where
xc is the critical concentration of dopants,
V is the volume of the unit cell,
Z is the number of formula units per unit cell. As for Ba
2ZnB
2O
6 host,
V = 1336.010(2) Å
3,
Z = 8, and
xc = 0.05, the calculated value of
Rc is about 18.55 Å.
3.5 Temperature-dependent PL properties
Temperature-dependent PL spectra of Ba2ZnB2O6:0.05Eu3+ are measured and displayed in Fig. 6(a). Because of the resolution of the fluorescence spectrophotometer used for temperature-dependent PL measurement, the splits of the emission peak 5D0–7F1 and the 5D0–7F2 cannot be distinguished clearly. With the temperature increasing, the emission intensities decrease fast, and the intensity at 150 °C only remains 30% of that at room temperature. The fast decrease of emission intensity with the increasing of temperature is possibly because of the non-radiative process. The non-radiative transition probability is strongly dependent on temperature. With the temperature increase, the non-radiative transition probability enhances greatly, which will result in the decrease of the emission intensity. It can also be found that the main emission peak wavelength shifts from 616 nm to 623 nm. To better understand the temperature dependence of photoluminescence and to determine the activation energy for thermal quenching, the Arrhenius equation is fitted to the thermal quenching data:32| | |
IT = I0/[1 + exp(−Ea/kT)]
| (4) |
where I0 and IT are the luminescence intensities of Ba2ZnB2O6:0.05Eu3+ at room temperature and the testing temperature, respectively. Ea is activation energy and k is the Boltzmann constant (8.617 × 10−5 eV K−1). Fig. 6(b) plots ln(I0/IT − 1) vs. 1/kT and the slope of the line is calculated to be −0.27. Therefore, Ea is obtained to be 0.27 eV.
 |
| | Fig. 6 (a) Temperature-dependent PL spectra of Ba2ZnB2O6:0.05Eu3+ phosphor (λex = 393 nm). (b) The ln(I0/IT − 1) vs. 1/kT activation energy graph for thermal quenching of Ba2ZnB2O6:0.05Eu3+. | |
3.6 Charge compensation
In Ba2ZnB2O6:Eu3+, the charge is not balanced because of the trivalent Eu3+ ions occupy the sites of the bivalent Zn2+ ions, which will introduce Zn vacancies or extra O2− ions at nearby interstitial position. Each of these defects leads to the distortion of local environment symmetries of optical centers.33 Therefore, alkali metal ions Li+, Na+ and K+ are co-doped with Eu3+ in the matrix to balance the charge and reduce the variety of the activator symmetry sites. The PL spectra of Ba2ZnB2O6:0.05Eu3+ and Ba2ZnB2O6:0.05Eu3+, 0.05M (M = Li+, Na+ and K+) under 393 nm excitation are shown in Fig. 7. No significant difference is observed between the emission peaks of Ba2ZnB2O6:0.05Eu3+ and Ba2ZnB2O6:0.05M, 0.05Eu3+ (M = Li+, Na+ and K+) except the enhanced emission intensity. As shown in Fig. 7, the emission intensity compensated by Li+ is strongest. The intensities of the emission peak centered at 616 nm for the phosphors doped by Li+ and Na+ as charge compensators are enhanced 1.61 and 1.41 times compared with the corresponding peak intensity for the phosphor without charge compensators, respectively. However, the emission intensity of the peak centered at 616 nm is almost unchangeable when the K+ ions are added as charge compensation, which is considered to be caused by the huge difference between the K+ and Zn2+ ionic radii. In Ba2ZnB2O6:Eu3+, Eu3+ ions have been proved to occupy the two four-coordinated Zn2+ sites. The ionic radii of Li+, Na+ and K+ (CN = 4) reported by Shannon17 are 0.59 Å, 0.99 Å and 1.37 Å, respectively. Obviously, the ionic radius of Li+ is very close to the radius of Zn2+ (r = 0.60 Å, CN = 4), while K+ is too large, which will be difficult to occupy Zn2+ sites. Therefore Ba2ZnB2O6:0.05Eu3+ singly doped Li+ as compensator shows the strongest emission intensity.
 |
| | Fig. 7 PL spectra of Ba2ZnB2O6:0.05Eu3+ and Ba2ZnB2O6:0.05M, 0.05Eu3+ (M = Li+, Na+ and K+) (λex = 393 nm). | |
3.7 Decay properties and CIE coordinate
The decay curves of Ba2ZnB2O6:0.05Eu3+ and Ba2ZnB2O6:0.05Li+, 0.05Eu3+ were measured and shown in Fig. 8(a) and (b). The corresponding luminescence decay times of Eu3+ emission at 616 nm can be best expressed by the double exponential equation:25| |
I = I0 + A1 exp(−t/τ1) + A2 exp(−t/τ2)
| (5) |
where I0 and I represent the luminescence intensities when time is 0 and t, τ1 and τ2 are the fast and slow components of luminescent lifetime respectively, and A1 and A2 are the fitting parameters respectively. When there is no interaction between rare earth ions, the decay curve is usually a single exponential function. But both the decay curves of Ba2ZnB2O6:0.05Eu3+ and Ba2ZnB2O6:0.05Eu3+, 0.05Li+ exhibit obvious deviations from the single exponential decay. As displayed in Fig. 8(a) and (b), the decay curves of Ba2ZnB2O6:0.05Eu3+ and Ba2ZnB2O6:0.05Eu3+, 0.05Li+ are fitted well by the double exponential equation, which is one of the evidences that Eu3+ ions occupy the sites of both Zn (1) and Zn (2) and charge compensator does not change the site occupancy of Eu3+ in Ba2ZnB2O6. The related fitting parameters are listed in Table 2.
 |
| | Fig. 8 (a) Decay curve of Eu3+ fluorescence in Ba2ZnB2O6:0.05Eu3+; (b) decay curve of Eu3+ fluorescence Ba2ZnB2O6:0.05Li+, 0.05Eu3+ (excited at 393 nm, monitored at 616 nm). | |
Table 2 Constants (A) and decay times (τ) of Ba2ZnB2O6:0.05Eu3+ and Ba2ZnB2O6:0.05Li+, 0.05Eu3+
| |
A1 |
A2 |
τ1 (ms) |
τ2 (ms) |
τ (ms) |
| Ba2ZnB2O6:0.05Eu3+ |
0.963 |
2.148 |
0.619 |
1.077 |
0.98 |
| Ba2ZnB2O6:0.05Li+, 0.05Eu3+ |
1.506 |
2.130 |
1.565 |
1.360 |
1.45 |
The effective decay lifetime is calculated by the following equation:
| | |
τ = (A1τ21 + A2τ22)/(A1τ1 + A2τ2)
| (6) |
Based on eqn (6), the average lifetimes of Ba2ZnB2O6:0.05Eu3+ and Ba2ZnB2O6:0.05Li+ are 0.98 ms and 1.45 ms, respectively. The results indicate the decay times of Ba2ZnB2O6:0.05Eu3+ and Ba2ZnB2O6:0.05Li+, 0.05Eu3+ are in the order of milliseconds.
The CIE coordinates of Ba2ZnB2O6:0.05Eu3+ phosphors are calculated to be x = 0.655, y = 0.345, as shown in Fig. 9. The CIE coordinate of Ba2ZnB2O6:0.05Eu3+ is very close to the CIE coordinate of standard red light (x = 0.67, y = 0.33), and is considered to be better than that of the commercial Y2O2S:Eu3+ (x = 0.622, y = 0.351).34
 |
| | Fig. 9 CIE chromaticity diagram of Ba2ZnB2O6:0.05Eu3+ phosphor under 393 nm excitation. | |
Table 3 summarizes the photoluminescence properties of some Ba-based and Zn-based hosts doped by Eu3+ as well as the data reported in this study. For those phosphors listed in Table 3, there is no big difference for their photoluminescence properties. The reported critical distance is around 15 nm for all the phosphors and the decay time is the magnitude of millisecond. The quenching mechanism are all dipole–dipole interaction for the phosphors listed in Table 3, which indicate that the different site occupancy for Eu3+ will not bring a big effect on the quenching mechanism of phosphors. However, emission color is different for these two different group phosphors. Compared with the red color emitted by Zn-based phosphor, the Ba-based phosphors emit the light from orange to red color (5D0–7F1 and 5D0–7F2 transitions), which indicate that crystal environment will bring effect on the emission wavelength. When the large Eu3+ ion occupy Zn site, the volume of the crystal structure will be expanded and the polyhedron will be distorted, which will result in the increase the crystal field strength. As a consequence, a red shift will be observed in Zn-based phosphors.
Table 3 Luminescence properties including the position of emission peak, optimal concentration, critical distance and quenching mechanism for Eu3+ in some Ba-based and Zn-based hosts as well as that in Ba2ZnB2O6
| Host |
Strongest excited peak (nm) |
Main transition |
Optimal concentration |
Critical distance (Å) |
Quenching mechanism |
Life time (ms) |
Ref. |
| Ba5−2x(VO4)3Cl |
466 |
5D0–7F2 (614 nm) |
0.22 |
— |
Dipole–dipole |
0.56 |
13 |
| BaB2O4 |
394 |
5D0–7F2 (615 nm) |
0.06 |
14.56 |
Dipole–dipole |
— |
35 |
| KBaBP2O8 |
394 |
5D0–7F1 (594 nm) |
0.1 |
15.2 |
Dipole–dipole |
2.54 |
36 |
| ZnB2O4 |
393 |
5D0–7F2 (621 nm) |
0.1 |
— |
— |
— |
15 |
| Ba2ZnB2O6 |
393 |
5D0–7F2 (616 nm) |
0.05 |
18.55 |
Dipole–dipole |
1.45 |
This work |
4. Conclusion
In summary, the Ba2ZnB2O6:Eu3+ phosphors were synthesized and investigated. The site-preferred occupancy of Eu3+ in Ba2ZnB2O6 is studied by the Rietveld refinement. The doped Eu3+ ions prefer to occupy two Zn2+ sites. These phosphors can be excited efficiently by near-ultraviolet and blue light. The concentration quenching of the Ba2ZnB2O6:xEu3+ phosphors in the PL emission spectra optimized at 616 nm is x = 0.05. The PL emission intensities of Ba2ZnB2O6:0.05Eu3+ phosphors compensated by Li+, Na+ and K+ as charge compensators are enhanced, and the phosphors compensated by Li+ ions show the strongest emission. The critical distance (Rc) and the lifetime value are determined to be 18.55 Å and 0.98 ms, respectively. The CIE coordinate is calculated to be x = 0.655, y = 0.345.
Acknowledgements
This work was financially supported by National Natural Science Foundation of China (51372121, 61274053, U146020005), Natural Science Foundation of Tianjin (14JCYBJC17800), the Program for New Century Excellent Talents in University of China (NCET-11-0258), and the Program for Changjiang Scholars and Innovative Research Team in University (IRT0149). We thank Mrs Xu of N01 group, Institute of Physics, Chinese Academy of Science for her great help in collecting the powder X-ray diffraction data and Prof. Dr Z. Xia for the great help of in the temperature-dependent PL spectra measurement.
References
- M. Touboul, N. Penin and G. Nowogrocki, Solid State Sci., 2003, 5, 1327 CrossRef CAS.
- FindIt, Version 1.3.3, ICSD database, FIZ Karlsruhe, Germany, 2004–02 Search PubMed.
- L. Li, H. Liang, Z. Tian, H. Lin, Q. Su and G. Zhang, J. Phys. Chem. C, 2008, 112, 13763 CAS.
- A. Shyichuk and S. Lis, J. Rare Earths, 2011, 29, 1161 CrossRef CAS.
- L. K. Bharat, B. V. Rao and J. S. Yu, Chem. Eng. J., 2014, 255, 205 CrossRef.
- L. K. Bharat and J. S. Yu, J. Nanosci. Nanotechnol., 2013, 13, 8239 CrossRef CAS.
- Y. Huang, Y. Nakai, T. Tsuboi and H. J. Seo, Opt. Express, 2011, 19, 6303 CrossRef CAS.
- F. Lei and B. Yan, J. Solid State Chem., 2008, 181, 855 CrossRef CAS.
- S.-F. Wang, K. Koteswara Rao, Y.-R. Wang, Y.-F. Hsu, S.-H. Chen and Y.-C. Lu, J. Am. Ceram. Soc., 2009, 92, 1732 CrossRef CAS.
- L. Wu, Y. Zhang, M. Gui, P. Lu, L. Zhao, S. Tian, Y. Kong and J. Xu, J. Mater. Chem., 2012, 22, 6463 RSC.
- Y. Zhang, L. Wu, M. Ji, B. Wang, Y. Kong and J. Xu, Opt. Mater. Express, 2012, 2, 92 CrossRef CAS.
- L. He, Y. Wang and W. Sun, J. Rare Earths, 2009, 27, 385 CrossRef.
- Z. Xia, Y. Liang, W. Huang, M. Zhang, D. Yu, J. Wu, J. Zhao, M. Tong and Q. Wang, J. Mater. Sci.: Mater. Electron., 2013, 24, 5111 CrossRef CAS.
- V. Singh, V. Natarajan and J.-J. Zhu, Opt. Mater., 2007, 29, 1447 CrossRef CAS.
- W.-R. Liu, C. C. Lin, Y.-C. Chiu, Y.-T. Yeh, S.-M. Jang and R.-S. Liu, Opt. Express, 2010, 18, 2946 CrossRef CAS.
- R. W. Smith and L. J. Koliha, Mater. Res. Bull., 1994, 29, 1203 CrossRef CAS.
- R. Shannon, Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr., 1976, 32, 751 CrossRef.
- L. Wu, M. Y. Ji, H. R. Wang, Y. F. Kong and Y. Zhang, Opt. Mater. Express, 2014, 4, 1535 CrossRef.
- L. Wu, B. Wang, Y. Zhang, L. Li, H. R. Wang, H. Yi, Y. F. Kong and J. J. Xu, Dalton Trans., 2014, 43, 13845 RSC.
- H. Rietveld, Acta Crystallogr., 1967, 22, 151 CrossRef CAS.
- H. Rietveld, J. Appl. Crystallogr., 1969, 2, 65 CrossRef CAS.
- J. Rodriguez-Carvajal, FULLPROF2000, LLB, France, 2001 Search PubMed.
- Y. Nobuhiko, J. Phys. Soc. Jpn., 1973, 35, 1089 CrossRef.
- C. A. Kodaira, H. F. Brito and M. C. F. Felinto, J. Solid State Chem., 2003, 171, 401 CrossRef CAS.
- G. Blasse and B. Grabmaier, Luminescent materials, Springer, 1994 Search PubMed.
- X. Bai, G. Zhang and P. Fu, J. Solid State Chem., 2007, 180, 1792 CrossRef CAS.
- Y.-C. Li, Y.-H. Chang, Y.-F. Lin, Y.-S. Chang and Y.-J. Lin, J. Alloys Compd., 2007, 439, 367 CrossRef CAS.
- B. Liu, C. Shi and Z. Qi, Appl. Phys. Lett., 2005, 86, 191111 CrossRef.
- G. V. L. Reddy, L. R. Moorthy, T. Chengaiah and B. C. Jamalaiah, Ceram. Int., 2014, 40, 3399 CrossRef CAS.
- D. Dexter and J. Schulman, J. Chem. Phys., 1954, 22, 1063 CrossRef CAS.
- G. Blasse, Philips Res. Rep., 1969, 24, 131 CAS.
- R.-J. Xie, N. Hirosaki, N. Kimura, K. Sakuma and M. Mitomo, Appl. Phys. Lett., 2007, 90, 191101 CrossRef.
- M. Puchalska, E. Zych, M. Sobczyk, A. Watras and P. Deren, Mater. Chem. Phys., 2014, 147, 304 CrossRef CAS.
- C. Guo, L. Luan, C. Chen, D. Huang and Q. Su, Mater. Lett., 2008, 62, 600 CrossRef CAS.
- J. Liu, X. D. Wang, Z. C. Wu and S. P. Kuang, Spectrochim. Acta, Part A, 2011, 79, 1520 CrossRef CAS.
- B. Han, P. Li, J. Zhang and H. Shi, J. Lumin., 2014, 155, 15 CrossRef CAS.
Footnote |
| † Electronic supplementary information (ESI) available: Rietveld refinement results for Ba2ZnB2O6:0.05Eu3+ phosphor. See DOI: 10.1039/c4ra11516c |
|
| This journal is © The Royal Society of Chemistry 2014 |
Click here to see how this site uses Cookies. View our privacy policy here.