DOI:
10.1039/C6RA11258G
(Paper)
RSC Adv., 2016,
6, 68852-68859
Synthesis, structure and luminescent properties of a new blue-green-emitting garnet phosphor Ca2LuScZrAl2GeO12:Ce3+†
Received
1st May 2016
, Accepted 13th July 2016
First published on 13th July 2016
Abstract
A series of new garnet phosphors, Ca2Lu1−xZrScAl2GeO12:xCe3+ (0.005 ≤ x ≤ 0.1), have been successfully synthesized through a conventional high-temperature solid-state reaction method. The crystal structure and electronic structure of the host, and the morphology, luminescence property, diffuse reflectance spectra, fluorescence lifetime and thermal stability of the phosphors are investigated in this article. The crystal structure was characterized by X-ray diffraction Rietveld refinement and it belongs to the Ia
d(230) space-group. The band gap of this matrix is about 5.15 eV, which is much narrower than that of YAG, enhancing the possibility of photo-ionization. The optimized phosphor can be excited by UV or near-UV light in the range of 320–420 nm and exhibits a broad blue-green emission centering at 472 nm under 405 nm excitation. In addition, the concentration quenching mechanism and differentiation of dodecahedral sites were discussed in detail.
1 Introduction
White light-emitting diodes (w-LEDs) have been widely used in the illumination and display fields due to their excellent properties, such as energy saving, long lasting life, fast start up etc.1–3 Nowadays, the combination of blue InGaN LED chips with yellow garnet phosphor YAG:Ce3+ is the main route to fabricate w-LEDs.4,5 However, these lamps are inherently limited to low color-rendering index (CRI) and high correlated color temperature (CCT) due to the red emission deficiency in the visible spectrum.6 Meanwhile, the application of w-LEDs has an intense blue emission, which easily causes sleep disorders and harms human health.7 In order to solve the above problems, combining tricolor phosphors with near ultraviolet (n-UV, 380–420 nm) LED chips to fabricate w-LEDs is considered as a promising route to cover indoor lighting needs for their better color-rendering ability, lower correlated color temperature, and tunable blue emission intensity.8–10
It is well known that Ce3+ activated garnet phosphors always have strong and broad absorption in the blue or n-UV light region ascribed to transition of 4f to 5d parity allowed electric dipole. Moreover, the photoluminescence properties of Ce3+ strongly depend on crystal-field splitting of the 5d state and the nephelauxetic effect, which lead to the emission peak of Ce3+ varying from blue to red region.11 Furthermore, the garnet compound has an outstanding physicochemical stability and an opened framework, for which, it is extensively investigated as the matrix material for phosphor. It can be demonstrated by a general formula {A}3[B]2(X)3O12, where A, B and X indicate cations which are coordinated by eight, six and four O2− ions forming dodecahedron, octahedron and tetrahedron, respectively.12–14 Moreover, A, B and X sites can be substituted by different cations, thus a variety of new matrix materials for phosphors have been designed by single- or multi-ions substitution, for instance, substituting Y3+ by other rare earth cations (Lu3+, Gd3+, Tb3+, etc.),15–19 Mg2+–Si4+/Ge4+ replacing Al3+–Al3+,13,20,21 or Si4+–N3− superseding Al3+–O2− in the YAG:Ce3+,22 etc. Thereupon, more and more efficient phosphors have been successively developed, such as YAG:Ce3+,5 Ca3Sc2Si3O12:Ce3+,12 Lu2CaMg2(Si,Ge)3O12:Ce3+,13 etc.
Motivated by the above investigations, some Ce3+ doped new garnet phosphors, which can be excited by near-UV LED chip, have been successively designed and synthesized.23,24 Typically, Zhong et al. replaced Sn4+ with Zr4+ in garnet host Ca3Sn2SiGa2O12 and obtained a new blue-green-emitting phosphor Ca3Zr2SiGa2O12:Ce3+,25 and Wang et al. reported a new cyan-green garnet phosphor Ca2YZr2Al3O12:Ce3+,26 etc.
Zhong et al. summarized the B site tends to be filled by a single kind of atom, which is helpful to maintain the cubic structure with minimum lattice stress.11 However, in this work, we extend another approach, applying the Diagonal Relationship theory, to design a new garnet matrix. As we all known, Sc/Zr are diagonal elements in periodic table of the elements, respectively, which have the similar chemical properties and ionic radius based on the Diagonal Relationship theory.27 Herein, the A site is introduced with Ca and Y; B site with Sc and Zr, and X sites with Al and Ge. However, A site with larger cations generally destabilize the garnet structure.21 So we replaced Y3+ with smaller Lu3+ ions. Finally, a new blue-green-emitting garnet phosphor Ca2LuScZrAl2GeO12:Ce3+ was successfully synthesized. The structures, electronic structure, photoluminescence properties and related mechanisms of the phosphor were investigated in detail.
2 Experimental
2.1 Materials and synthesis
The powder samples Ca2Lu1−xScZrAl2GeO12:xCe3+ (0 ≤ x ≤ 0.1) were synthesized by high-temperature solid-state reaction method. The raw materials CaCO3 (99.9%), Lu2O3 (99.99%), Sc2O3 (99.99%), ZrO2 (99.99%), Al2O3 (99.99%), GeO2 (99.99%) and CeO2 (99.99%), were all weighed out as the stoichiometry, and thoroughly mixed in an agate mortar. The powder mixtures were put in an alumina crucible and continually fired at 1400 °C in a reducing atmosphere (CO) for 4 h. Finally, the samples were cooled to room temperature, and ground in an agate mortar.
2.2 Calculation methods
In our work, density functional theory (DFT) method is used for structural relaxation and electronic structure calculation. The ion–electron interaction is treated by the projector augmented-wave28 technique as implemented in the Vienna ab initio simulation package (VASP).29 The exchange-correlation potential is treated by PBE.30 For structural relaxation, all the atoms are allowed to relax until atomic forces are smaller than 0.01 eV Å−1. The k-mesh is generated by the Monkhorst–Pack scheme, which depends on the lattice constant.
2.3 Characterization methods
The phase formation and crystal structure of the as-prepared samples were analyzed by the X-ray powder diffractometer (Rigaku, Japan) with Co-Ka radiation (λ = 0.17889 nm). The data were collected over a 2θ range from 10° to 90° with scanning speed of 6° 2θ min−1 for the series Ca2Lu1−xScZrAl2GeO12:xCe3+ (0 ≤ x ≤ 0.1) samples and 10–90° at intervals of 0.02° with a count time of 5 s per step for the structural refinement data (with Cu-Ka radiation (λ = 0.15418 nm)). The structural parameters of Ca2LuScZrAl2GeO12 garnet was refined by the Rietveld method using the Fullprof software.31 The morphology of the samples were obtained using scanning electron microscopy (SEM; Hitachi SUI 510).
The photoluminescence spectra were measured by using a FLS 920 spectrofluorometer of Edinburgh Instruments with a monochromated Xe lamp as an excitation source. The thermal quenching were measured by a spectrofluorometer (Fluoromax-4, Edison, U.S.A.), which are composed of a Xe high-pressure arc lamp, a photomultiplier tube, and a heating apparatus. The reflectance spectra were measured by a Shimadzu UV-3600 Plus UV-VIS-NIR Spectrophotometer with an integrated sphere attachment, and BaSO4 was used as a white standard. Quantum efficiency was measured using the integrating sphere on the QE-2100 quantum yield measurement system (Otsuka Electronics Co., Ltd., Japan), and a Xe lamp was used as an excitation source and white BaSO4 powder as a reference. The fluorescent lifetime was conducted on a LifeSpec Red Spectrometer with excitation wavelength at 405 nm.
3 Results and discussion
3.1 Phase and crystal structure analysis
Fig. 1 shows the X-ray diffraction (XRD) patterns of as-prepared Ca2Lu1−xScZrAl2GeO12:xCe3+ (0 ≤ x ≤ 0.1) phosphors. The XRD results show that all samples are in agreement with the garnet structure, which indicates that Ce3+ have been doped into the crystal lattices of the cubic Ca2LuScZrAl2GeO12 instead of forming the impurity phase or generating significant structure changes in the host. Furthermore, the peaks of the synthesized material exhibit a slight shift towards lower 2θ angles with the increasing of Ce3+ content. This reflects the expansion of lattices when bigger Ce3+ ions substitute for smaller Lu3+ ions on dodecahedral sites.11 Rietveld refinement of XRD pattern of the Ca2LuScZrAl2GeO12 garnet is displayed in Fig. 2. The refined residual factors, atoms coordinates and unit cell parameters are summarized in Table 1. The refinement is convergent well with residual factors Rwp = 9.93, Rp = 9.16 and χ2 = 3.35. And the cell parameters are obtained to be a = b = c = 12.43660(4) Å and V = 1923.55(9) Å3.
 |
| Fig. 1 Powder XRD patterns of Ca2Lu1−xZrScAl2GeO12:xCe3+ (0 ≤ x ≤ 0.10). The panel on the right details the zoomed diffraction peak with varied Ce3+ contents near 2θ = 37.6°. | |
 |
| Fig. 2 Rietveld refinement XRD patterns of the Ca2LuZrScAl2GeO12 host; inset shows the crystal structure of Ca2LuZrScAl2GeO12 (a) and the coordination environment of cations (b). | |
Table 1 Results of structure refinement of Ca2LuZrScAl2GeO12
Formula |
Ca2LuZrScAl2GeO12 |
Symmetry |
Cubic |
Space group |
Ia d |
a = b = c (Å) |
12.43660(4) |
α = β = γ |
90° |
V (Å3) |
1923.55(9) |
Z |
8 |
Rp |
9.16 |
Rwp |
9.93 |
χ2 |
3.35 |
Atom |
Site |
x |
y |
z |
Occu. |
Ca |
24c |
0.12500 |
0.00000 |
0.25000 |
0.1351(4) |
Lu |
24c |
0.12500 |
0.00000 |
0.25000 |
0.0869(7) |
Zr |
16a |
0.00000 |
0.00000 |
0.00000 |
0.0969(9) |
Sc |
16a |
0.00000 |
0.00000 |
0.00000 |
0.0500(2) |
Al |
24d |
0.37500 |
0.00000 |
0.25000 |
0.3307(7) |
Ge |
24d |
0.37500 |
0.00000 |
0.25000 |
0.0157(6) |
O |
96h |
0.28449 |
0.09539 |
0.19698 |
1.0003(6) |
The inset of Fig. 2 presents a schematic diagram of the Ca2LuScZrAl2GeO12 garnet structure according to the Rietveld refinement. Ca2+/Lu3+, Sc3+/Zr4+, Al3+/Ge4+ cations are randomly distributed on the dodecahedral, tetrahedral, octahedral sites, connecting with eight, six and four surrounding O2−, respectively.32 Furthermore, the [(Sc/Zr)O6] octahedra are shared edges with [(Ca/Lu)O8] dodecahedra and [(Al/Ge)O4] tetrahedra, respectively. The [(Ca/Lu)O8] dodecahedra are connected with part of [(Al/Ge)O4] tetrahedra by O2− points, and the other tetrahedra by edges.33 That is to say, different ratio of Zr4+/Sc3+ or Al3+/Ge4+ surrounding dodecahedron can lead differentiated coordination environment of the dodecahedral sites, namely, differentiation of the dodecahedral sites, thereby affecting the luminescent properties of Ce3+.25
The Ca2LuZrScAl2GeO12, band structure (BS) and density of states (DOS) of the pure Ca2LuZrScAl2GeO12 host are investigated by first-principles calculation, presented in Fig. 3 and 4, respectively. The calculated band gap of the pure Ca2LuZrScAl2GeO12 host corresponding to the direct transition is 3.297 eV, which is much narrower than that of YAG 5.0 eV calculated by same method.34 Many studies reported that Ce3+-activated Ge4+-containing garnet phosphors always have a fatal drawback of relatively low luminescence intensity, which is most likely attributed to Ce3+ photo-ionization in Ge4+ series garnet matrices.13,35,36 This phenomenon can be explained by the fact that 5d levels at higher energy of Ce3+ ion is already in the conductive band of matrix which have a small band gap, and then the excited 5d electron of Ce3+ ion can enter this band easily.37 From our studies, the relatively low band gap may enhance the photo-ionization effect of this Ce3+-activated phosphors.
 |
| Fig. 3 Band structures of Ca2LuScZrAl2GeO12. | |
 |
| Fig. 4 Density of states of Ca2LuScZrAl2GeO12. | |
In Fig. 4, the DOS shows that the top edge of the valence band mainly consists of the O 2p orbital. And the 4s orbital of Ge and the 4d orbital of Zr are the major contributors to the bottom edge of the conductive band, which indicate that the dominant role in photo-ionization process is not only played by the Ge 4s orbital, but also by the Zr 4d orbital. Perhaps that is the reason for Ce3+-activated Ge-containing and Zr-containing series garnet phosphors always have relatively low luminescence efficiency.23–25,35,36
The Fig. 5 exhibits the SEM image of the Ca2Lu0.99ScZrAl2GeO12:0.01Ce3+ phosphor. It reveales that the grains trend to form spherical shape and smooth morphology, and the size is in the range of 2–10 μm, which also indicates that the well-crystallized powders have been obtained.
 |
| Fig. 5 SEM image of the Ca1.99LuZrScAl2GeO12:0.01Ce3+. | |
3.2 Photoluminescence properties of the Ca2LuScZrAl2GeO12:Ce3+ phosphor
The emission (PL) and excitation (PLE) spectra of the Ca2Lu0.99ZrScAl2GeO12:0.01Ce3+ at room temperature are depicted in the Fig. 6. The PLE spectrum monitored at 472 nm, ranges from 300 to 450 nm, and exhibits two main broad bands at 348 nm and 405 nm. Moreover, the position of the lowest Ce3+ 4f1–5d1 absorption transition in Ca2Lu0.99ZrScAl2GeO12:0.01Ce3+ is at a much higher energy level than that of classical garnet phosphor YAG:Ce3+, which indicates much weaker crystal field strength in this sample. Additionally, the emission spectrum of Ca2Lu0.99ZrScAl2GeO12:0.01Ce3+ under excitation at 405 nm, exhibits a blue-green emission band with a typical asymmetrical spectra from 425 nm to 600 nm peaking at 472 nm. We can decompose the emission band into two Gaussian profiles (the dashed lines in Fig. 6) with maximum at 464 nm (21
552 cm−1) and 494 nm (20
243 cm−1), respectively. The energy difference Δk is about 1310 cm−1, which is different with the theoretical energy difference between 2F5/2 and 2F7/2 levels of the Ce3+ ground state (∼2000 cm−1).38 It indicates that Ce3+ ions not only enter one specific site, which shows agreement with the structure analysis that a differentiation appears to the dodecahedral site of Ca2LuZrScAl2GeO12 host.25 So the emission band should be decomposed into multiple Gaussian profiles rather than two for this phosphor.
 |
| Fig. 6 PLE (λem = 472 nm) spectra and PL (λex = 405 nm) spectra of the Ca2Lu0.99ZrScAl2GeO12:0.01Ce3+ phosphor. | |
Fig. 7 illustrates the dependence of the emission spectra with Ce3+ doping concentration for Ca2Lu1−xZrScAl2GeO12:xCe3+ (0.005 ≤ x ≤ 0.1) samples under 405 nm excitation. Additionally, the inset of Fig. 7 shows the Ce3+ concentration dependent emission intensity and peak position. With the increasing of Ce3+ concentration, the emission peak position has a red shift from 469 nm to 493 nm. This phenomenon can be explained by the change of crystal field strength. The crystal field splitting (Dq) can be determined by the following eqn (1):39
|
 | (1) |
where
Dq relates a measure of the energy level separation,
Z is the anion charge,
e is the electron charge,
r represents the radius of the
d wavefunction, and
R is the bond length. With the increasing concentration of bigger Ce
3+ entering into smaller dodecahedral sites, it is possible that the Ce
3+ will suffer intensive compression from other neighboring atoms in the rigid garnet structure, which results in a closer band length between Ce
3+ and O
2−. Moreover, the crystal field splitting is proportional to 1/
R5, shorter Ce
3+–O
2− distance can lead to the enhancement of crystal field surrounding the Ce
3+ ion, thus results in a larger crystal field splitting of the 5d level of Ce
3+ with a red shift in the emission spectra.
 |
| Fig. 7 PL spectra (λex = 405 nm) of Ca2−xLuZrScAl2GeO12:xCe3+ with varying Ce3+ concentration. Inset shows the emission intensity and emission peak wavelength as a function of the Ce3+ content. | |
In the Fig. 7, the emission intensity of samples first increases with Ce3+ content reaching the maximum intensity at x = 0.01, and then decreases gradually with x beyond the critical concentration. Obviously, the concentration quenching occurs when the Ce3+ concentration exceeds 1 mol%. The critical distance (Rc) of energy transfer among Ce3+ ions in Ca2LuZrScAl2GeO12 can be demonstrated by the flowing eqn (2):40
|
 | (2) |
here,
V corresponds to the unit cell volume;
Xc equals the critical concentration of dopant ions Ce
3+ here; and
N represents the number of host cations in the unit cell. Here, the values
V = 1923.55(9) Å
3,
N = 24,
Xc = 0.01, then we can obtain the critical distance
Rc of energy transfer is about 24.8 Å. Apparently, this value is much longer than 5 Å,
41 indicating little possibility of the nonradiative energy transfer between different Ce
3+ ions in the Ca
2LuZrScAl
2GeO
12 host
via the exchange interaction mechanism. Generally, the other two mechanisms of the nonradiative energy transfer are multipolar interactions and radiation reabsorption.
42 The multipolar interactions between Ce
3+ ions can be identified by the following formula
(3) according to the Dexter's theory,
43 |
 | (3) |
where
I is the emission intensity,
x is the concentration of the Ce
3+ ion,
β and
k are constants for the given host under the same excitation conditions,
θ is a function of the multipole–multipole interaction. According to the Van Uitert's report,
θ = 6, 8, and 10 represents dipole–dipole (d–d), dipole–quadrupole (d–q), and quadrupole–quadrupole (q–q) interactions, respectively. In order to get the value of
θ, the relationship between lg(
I/
x) and lg(
x) with a slope of −1.39 by linear fitting illustrated in
Fig. 8. So the calculated
θ = 4.2, which is roughly close to 6, suggesting that the dipole–dipole (d–d) interaction is not the only reason of the nonradiative energy transfer in Ca
2LuZrScAl
2GeO
12:Ce
3+ phosphor. On the other hand, we can't ignore the overlap between the excitation band and emission band in
Fig. 6. In fact, Ce
3+ excitation garnet phosphors always have such a reabsorption problem in nonradiative energy transfer process.
11,44 So that the reabsorption effect should be taken into consideration in the nonradiative energy transfer process for Ca
2LuZrScAl
2GeO
12:Ce
3+ phosphor.
 |
| Fig. 8 The linear fitting of the relationship of lg(I/x) vs. lg(x). | |
Fig. 9 presents the temperature dependent emission spectra of the optimized Ca2Lu0.99ZrScAl2GeO12:0.01Ce3+ phosphor under excitation at 405 nm and the inset shows the relative emission intensities as a function of temperature. It reveales that the relative PL intensity decreases with the increase of temperature, and remains only 40% of the corresponding initial value (at room temperature) when the temperature was raised up to 100 °C, which is far lower than YAG:Ce3+ (no less than 90%).11 Generally, the thermal quenching process can be explained by the enhanced phonon–electron interactions with the increase of temperature and the thermally activated photo-ionization of the luminescent process.39 To verify the origin of temperature-dependent emission intensity IT, the activation energy ΔE for the electrons excited from the ground-state level 4f to 5d1 of the Ce3+ ions can be ascribed in the following equation:39
|
 | (4) |
where
I0 represent the initial emission intensity at room temperature and
IT is the intensity at different temperature,
c is a constant for a certain host, and
k is the Boltzmann constant (8.629 × 10
−5 eV). By linear fitting the relationship of In[(
I0/
IT) − 1]
vs. 1000/
T for the Ca
2Lu
0.99ZrScAl
2GeO
12:0.01Ce
3+ phosphor with a slop of −3.535 (as shown in
Fig. 10), the activation energy is calculated to be 0.305 eV, which is quite lower than that of YAG:Ce
3+ (0.81 eV).
45 Thus, the thermally activated photo-ionization of the luminescent process should be accounted for the thermal quenching of Ca
2Lu
0.99ZrScAl
2GeO
12:0.01Ce
3+. As the electronic structure study of matrix in
Fig. 3 and
4, photo-ionization effect may be easy to occur when the matrix contains Ge
4+ or Zr
4+ in garnet phosphor.
 |
| Fig. 9 Temperature-dependent PL spectra of Ca2Lu0.99ZrScAl2GeO12:0.01Ce3+ phosphor. Inset shows the emission intensity as a function of temperature. | |
 |
| Fig. 10 Arrhenius fitting of the emission intensity of Ca2Lu0.99ZrScAl2GeO12:0.01Ce3+ phosphor and the calculated activation energy for thermal quenching. | |
Fig. 11 exhibits the diffuse reflectance spectra (DRS) of Ce3+-doped and undoped Ca2LuZrScAl2GeO12. The undoped Ca2LuZrScAl2GeO12 shows a drop in reflection in the range about 450 nm, corresponding to the transition from the valence to conduction band of the host lattice. The intense reflection in the visible spectral range is in agreement with the grey-white daylight colour of undoped Ca2LuZrScAl2GeO12. Compared with the undoped sample, the Ce3+-doped Ca2LuZrScAl2GeO12 displays a pronounced absorption band peaking at about 405 nm which is attributed to the absorption by the Ce3+ ions. The band gap energy can be determined by extrapolating the absorption edge onto the energy axis, as shown in the inset of Fig. 11.46 And the conversion of the reflectance to absorbance data can be obtained by the Kubelka–Munk function (K–M)46,47
|
 | (5) |
where
R is the reflectance;
F(
R) is proportional to the extinction coefficient. The band gap energy of undoped Ca
2LuZrScAl
2GeO
12 by reflection spectrum analysis is estimated to be 5.15 eV (as shown in
Fig. 6 inset), which is much lower than YAG (6.5 eV).
11 Compared with calculated band gap energy of the host, the result indicates that the first principle calculation method could qualitatively analyze and predict the luminescence properties of phosphor. It is conceivable that the photo-ionization effect is more likely to occur in these Ge
4+- or Zr
4+-containing garnet phosphors.
 |
| Fig. 11 Diffuse reflectance spectra of the Ca1.99LuZrScAl2GeO12:0.01Ce3+ and Ca2LuZrScAl2GeO12 host at room temperature. The inset is the calculated energy of host. | |
Fig. 12 illustrates the florescence decay curves of different Ca2Lu1−xZrScAl2GeO12:xCe3+ phosphors under 405 nm excitation and monitored at 475 nm. Normally the decay curves of luminescent materials are represented as a single exponential function, i.e., lg(t) vs. t presents a linear relation, for the phosphor which has only one type of luminescent centre, such as YAG:Ce.48 However, the single exponential function is not able to deal with the decay curves of Ca2Lu1−xZrScAl2GeO12:xCe3+ phosphors. But a three exponential function (6) can be well fitted.49
|
 | (6) |
where
I is emission intensity,
A,
A1,
A2 and
A3 are constants,
t is time,
τ1,
τ2 and
τ3 are decay time for the exponential components, respectively. The average lifetime of the Ca
2Lu
1−xZrScAl
2GeO
12:
xCe
3+ phosphors were determined to be 9.84 ns, 8.81 ns, 7.14 ns, 5.34 ns, 4.08 ns, 3.40 ns and 3.13 ns by the following formula
(7),
50 |
 | (7) |
 |
| Fig. 12 Decay curves and calculated lifetimes of Ca2−xLuZrScAl2GeO12:xCe3+ (0.005 ≤ x ≤ 0.1) phosphor with varying Ce3+ concentration. | |
The decay time decreases gradually with increasing of Ce3+ concentration, which indicates that the nonradiative transition has been enhanced.51 Nevertheless, the decay curves fit well with a three exponential function and the relationship of lg
I(t) vs. t gradually deviates from a straight line with the increase of Ce3+ concentration, which reveals that more and more Ce3+ ions occupy the differentiated dodecahedral sites.
The Fig. 13 illustrates the CIE value of the Ca2Lu1−xZrScAl2GeO12:xCe3+ phosphors excited at 405 nm. The emission color of Ca2Lu1−xZrScAl2GeO12:xCe3+ phosphors varies from blue to green and the chromaticity index can be tuned from (0.1386, 0.2287) to (0.1600, 0.3644) by adjusting the concentration of Ce3+ ions from 0.5% to 1%. Additionally, the internal quantum efficiency (QE) of Ca1.99LuZrScAl2GeO12:0.01Ce3+ phosphor is measured to be 18.9% under 405 nm excitation at room temperature. The relatively low QE value corroborates the opinion that photo-ionization effect may be easy to occur in luminescence process of these Ce3+-activated phosphors. And this can be further improved by synthesis and composition optimization, especially, making further substitutions, such as Si4+ replacing Ge4+, which is conducive to reduce the photo-ionization process.13
 |
| Fig. 13 The CIE coordinates of Ca2Lu1−xZrScAl2GeO12:xCe3+ (0.005 ≤ x ≤ 0.1) phosphors under excitation at 405 nm. | |
4 Conclusions
In summary, we applied the Diagonal Relationship theory to phosphor design and successfully synthesized a group of new garnet Ca2Lu1−xZrScAl2GeO12:xCe3+ (0.005 ≤ x ≤ 0.1) phosphors by high-temperature solid-state reaction. The XRD and Rietveld refinement results confirm that this phosphor belong to garnet structure. The band gap between conductive band and valance band is about 5.15 eV, which is quite narrow that may cause a photo-ionization process. Both the 4s orbital of Ge and the 4d orbital of Zr play a dominant role in photo-ionization process which conjectured by the PDOS of the matrix. The SEM image reveals that the sample trends to form spherical shape and smooth morphology with the size at the range of 2–10 μm. In addition, the Ca2LuZrScAl2GeO12:Ce3+ shows two main broad excitation bands with the peaks at 348 nm and 405 nm. As the Ce3+ content increases, the emission spectra of Ca2LuZrScAl2GeO12:Ce3+ phosphor varies from 469 nm to 493 nm under 405 nm excitation, which mainly due to the increasing crystal field splitting. The optimal Ce3+ doping concentration is x = 0.01. Moreover, the critical distance of energy transfer among Ce3+ ions in Ca2LuZrScAl2GeO12 is 24.8 Å, and the corresponding concentration quenching mechanism is d–d interaction and radiation reabsorption.
Acknowledgements
This present work was financially supported by the National Basic Research Program of China (No. 2014CB643801).
References
- S. Nakamura, T. Mukai and M. Senoh, Appl. Phys. Lett., 1994, 64, 1687–1689 CrossRef CAS.
- E. Gibney, Nature, 2014, 514, 152–153 CrossRef CAS PubMed.
- E. F. Schubert and K. J. Kim, Science, 2005, 308, 1274–1278 CrossRef CAS PubMed.
- A. Setlur, Electrochem. Soc. Interface, 2009, 16, 32 Search PubMed.
- V. Bachmann, C. Ronda and A. Meijerink, Chem. Mater., 2009, 21(10), 2077–2084 CrossRef CAS.
- Y. Jia, Y. Huang, Y. Zheng, N. Guo, H. Qiao, Q. Zhao, W. Lv and H. You, J. Mater. Chem., 2012, 22(30), 15146–15152 RSC.
- C. Cajochen, M. Münch, S. Kobialka, K. Kräuchi, R. Steiner, P. Oelhafen, S. Orgül and A. Wirz-Justice, J. Clin. Endocrinol. Metab., 2005, 90(3), 1311–1316 CrossRef CAS PubMed.
- W. Lü, Z. Hao, X. Zhang, X. Liu, X. Wang and J. Zhang, Opt. Mater., 2011, 33(8), 1262–1265 CrossRef.
- J. K. Park, K. J. Choi, H. G. Kang, J. M. Kim and C. H. Kim, Electrochem. Solid-State Lett., 2007, 10(2), J15–J18 CrossRef CAS.
- W. R. Liu, C. H. Huang, C. P. Wu, Y. C. Chiu, Y. T. Yeh and T. M. Chen, J. Mater. Chem., 2011, 21(19), 6869–6874 RSC.
- J. Zhong, W. Zhuang, X. Xing, R. Liu, Y. Liu and Y. Hu, J. Phys. Chem. C, 2015, 119, 5562–5569 CAS.
- Y. Shimomura, T. Honma, M. Shigeiwa, T. Akai, K. Okamoto and N. Kijima, J. Electrochem. Soc., 2007, 154(1), J35–J38 CrossRef CAS.
- A. A. Setlur, W. J. Heward, Y. Gao, A. M. Srivastava, R. G. Chandran and M. V. Shankar, Chem. Mater., 2006, 18(14), 3314–3322 CrossRef CAS.
- Y. Jia, Y. Huang, N. Guo, H. Qiao, Y. Zheng, W. Lv, Q. Zhao and H. You, RSC Adv., 2012, 2, 2678–2681 RSC.
- H. L. Li, X. J. Liu and L. P. Huang, J. Am. Ceram. Soc., 2005, 88(11), 3226–3228 CrossRef CAS.
- J. K. Li, J. G. Li, X. L. Wu, S. H. Liu, X. D. Li and X. D. Sun, Key Eng. Mater., 2013, 544, 245–251 CrossRef.
- Y. S. Lin, R. S. Liu and B. M. Cheng, J. Electrochem. Soc., 2005, 152(6), J41–J45 CrossRef CAS.
- Y. S. Lin and R. S. Liu, J. Lumin., 2007, 122, 580–582 CrossRef.
- H. Yang and Y. S. Kim, J. Lumin., 2008, 128(10), 1570–1576 CrossRef CAS.
- M. Shang, J. Fan, H. Lian, Y. Zhang, D. Geng and J. Lin, Inorg. Chem., 2014, 53(14), 7748–7755 CrossRef CAS PubMed.
- Y. Shi, G. Zhu, M. Mikami, Y. Shimomura and Y. Wang, Dalton Trans., 2015, 44(4), 1775–1781 RSC.
- A. A. Setlur, W. J. Heward, M. E. Hannah and U. Happek, Chem. Mater., 2008, 20(19), 6277–6283 CrossRef CAS.
- X. Gong, J. Huang, Y. Chen, Y. Lin, Z. Luo and Y. Huang, Inorg. Chem., 2014, 53(13), 6607–6614 CrossRef CAS PubMed.
- X. Ding, W. Geng, Q. Wang and Y. Wang, RSC Adv., 2015, 5(119), 98709–98716 RSC.
- J. Zhong, W. Zhuang, X. Xing, R. Liu, Y. Li, Y. Zheng, Y. Hu and H. Xu, RSC Adv., 2016, 6(3), 2155–2161 RSC.
- X. Wang and Y. Wang, J. Phys. Chem. C, 2015, 119(28), 16208–16214 CAS.
- G. Restrepo, H. Mesa, E. J. Llanos and J. Villaveces, J. Chem. Inf. Comput. Sci., 2004, 44(1), 68–75 CrossRef CAS PubMed.
- P. E. Blöchl, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50(24), 17953 CrossRef.
- G. Kresse and J. Hafner, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 49(20), 14251 CrossRef CAS.
- J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77(18), 3865 CrossRef CAS PubMed.
- J. Rodríguez-Carvajal, Newsletter, 2001, 26, 12 Search PubMed.
- W. Y. Huang, F. Yoshimura, K. Ueda, W. K. Pang, B. J. Su, L. Y. Jang, C. Y. Chiang, W. Zhou, N. H. Duy and R. S. Liu, Inorg. Chem., 2014, 53, 12822 CrossRef CAS PubMed.
- C. Milanese, V. Buscaglia, F. Maglia and U. Anselmi-Tamburini, Chem. Mater., 2004, 16(7), 1232–1239 CrossRef CAS.
- A. B. Muñoz-García, E. Anglada and L. Seijo, Int. J. Quantum Chem., 2009, 109(9), 1991–1998 CrossRef.
- D. Pasiński, E. Zych and J. Sokolnicki, J. Lumin., 2016, 169, 862–867 CrossRef.
- S. Pinelli, S. Bigotta, A. Toncelli, M. Tonelli, E. Cavalli and E. Bovero, Opt. Mater., 2004, 25(1), 91–99 CrossRef CAS.
- G. Blasse, W. Schipper and J. J. Hamelink, Inorg. Chim. Acta, 1991, 189(1), 77–80 CrossRef CAS.
- G. Blasse and B. C. Grabmaier, Energy Transfer, Springer, Berlin, 1994 Search PubMed.
- J. Sun, Z. Lian, G. Shen and D. Shen, RSC Adv., 2013, 3(40), 18395–18405 RSC.
- G. Blasse, Philips Res. Rep., 1969, 24(2), 131–144 CAS.
- J. Zhou, Z. Xia, M. Yang and K. Shen, J. Mater. Chem., 2012, 22(41), 21935–21941 RSC.
- X. Chen, Z. Xia and Q. Liu, Dalton Trans., 2014, 43(35), 13370–13376 RSC.
- D. L. Dexter, J. Chem. Phys., 1953, 21(5), 836–850 CrossRef CAS.
- Z. Pan, W. Li, Y. Xu, Q. Hu and Y. Zheng, RSC Adv., 2016, 6(25), 20458–20466 RSC.
- P. Dorenbos, J. Lumin., 2013, 134, 310–318 CrossRef CAS.
- E. A. Davis and N. F. Mott, Philos. Mag., 1970, 22(179), 0903–0922 CrossRef CAS.
- P. Kubelka and F. Munk, Z. Tech. Phys., 1931, 12, 593–601 Search PubMed.
- J. Zhong, W. Zhuang, X. Xing, L. Wang, Y. Li, Y. Zheng, R. Liu, Y. Liu and Y. Hu, J. Alloys Compd., 2016, 674, 93–97 CrossRef CAS.
- T. Katsumata, T. Nabae, K. Sasajima, S. Komuro and T. Morikawa, J. Electrochem. Soc., 1997, 144(9), L243–L245 CrossRef CAS.
- Y. Li, Y. Wang, X. Xu, G. Yu and N. Wang, J. Am. Ceram. Soc., 2011, 94(2), 496–500 CrossRef CAS.
- Z. G. Xia, R. S. Liu, K. W. Huang and V. Drozd, J. Mater. Chem., 2012, 22, 15183 RSC.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra11258g |
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.