Zaifa Pan*a,
Yu Xua,
Qingsong Hua,
Weiqiang Lia,
Huan Zhoub and
Yifan Zheng*a
aCollege of Chemical Engineering, Zhejiang University of Technology, Hangzhou 310014, P.R. China. E-mail: panzaifa@zjut.edu.cn
bResearch Center of Analysis and Measurement, Zhejiang University of Technology, Hangzhou 310014, P.R. China
First published on 19th December 2014
Cation substitution is a valuable approach to tune the crystal field splitting of garnet phosphor and hence enhances the red shift of the Ce3+ emission. A novel aluminate silicate garnet phosphor Mg2Y2Al2Si2O12:Ce3+ is designed with the aim of creating a large distortion of the dodecahedron and an external pressure from the large octahedron through the combination of the substitution of both dodecahedral Y3+ and octahedral Al3+ sites with Mg2+. The garnet structure and elemental composition of this phosphor were confirmed by XRD, SEM and TEM. The particular coordination environment of each element of the phosphor was illustrated by XPS and Rietveld refinement. Mg2Y1.94Al2Si2O12:0.06Ce3+ exhibits a strong and broad yellow-orange emission band with a CIE coordinate of (x = 0.519, y = 0.472) and shifts to the red side of 57 nm, compared with the commercial phosphor YAG:Ce3+. This red-shift could be mainly ascribed to the distortion and shrinking of the dodecahedron, and subsequently, the larger crystal field splitting of the Ce3+ 5d levels. A white LED was fabricated and showed a high colour rending index of up to 84. These results reveal that Ce3+-doped Mg2Y2Al2Si2O12 phosphor is a promising blue light converted yellow-orange light emitting phosphor for white LED.
Besides the co-doping of other activator ions, including the Pr3+,7 Sm3+,8 Cr3+,9 and Mn2+–Si4+ pairs,10 cation substitution is a useful approach to tune the crystal field splitting of the 5d orbit of Ce3+, and hence makes the emission considerably shift to the red side. These modifications involve the direct substitution of the dodecahedral activator ion sites with lager ion substituted Tb3Al5O12:Ce3+
11 and Gd3Al5O12:Ce3+.12 The octahedral sites could also be substituted by the widely accepted strategy of replacing Al3+ (0.51 Å) by the Mg2+ (0.66 Å)–Si4+ (0.42 Å) pair, which results in a stronger crystal field splitting and a further shift of the Ce3+ emission to the red side.13,14 Recently, a double substitution of Mg2+–Si4+/Ge4+ to replace Al(1)3+–Al(2)3+ on the octahedral and tetrahedral sites was reported and the red shift of the emission was interpreted via the bond distance increasing in the octahedron.15 However, the relationship between the luminescence shifts with the crystallography of these substitutions is still unclear, except for the phenomenological design rules of the cation size effects.14 On the other hand, besides this cation size effect on the redshift, Cheetham et al. preferred the distortion factor of the dodecahedral doping site.16,17 Furthermore, Seshadri's group revealed that the compression of the Ce3+ local environments could be ascribed to the unusually large crystal-field splitting of garnet, giving insights into the local environment of the activator ions.18
Considering the factors mentioned above, one could hypothesize that the part substitution of the dodecahedral site with small ion radii might directly result in a compression environment for the activator ion Ce3+ and hence a large distortion of the dodecahedron, and hence should result in strong crystal field splitting. Meanwhile, the large octahedron may provide external pressure on the activator centre dodecahedron and could, furthermore, enhance the crystal field splitting. Based on the above design rule, the combination cation substitution tuning garnet phosphor Mg2Y2Al2Si2O12:Ce3+ is proposed. Herein, Mg2+ is predicted to incorporate both dodecahedral Y3+ and octahedral Al3+ sites. Subsequently, a large distortion dodecahedron and a large octahedron could be built up, and hence a longer wavelength emission could be expected.
In this article, a novel Ce3+-doped garnet phosphor Mg2Y2Al2Si2O12 was synthesized by a traditional solid state reaction at a relatively low temperature. The crystal structure and the particular coordination environment of each element were illustrated by XRD, TEM and XPS. An intense yellow-orange emission was observed for the synthesized phosphor, and the effect of the crystal structure parameters of Mg2Y2Al2Si2O12 on such a red shift is discussed. A white LED (WLED) was fabricated and the colorimetric properties were evaluated.
The phase composition of the obtained powder samples were characterized by X-ray diffraction (XRD) analysis on a PNAlytical X'Pert Pro diffractometer with Cu Kα radiation at 40 mA and 40 kV in the 2θ range from 5° to 125°, with a scanning step of 0.0167°. A Rietveld structural refinement was carried out by General Structure Analysis System (GSAS) software package. The particle size, morphology and the content of each element were observed with a scanning electron microscope (SEM, Hitachi S-4700II) with an accelerating voltage of 15 kV, which also had an energy-dispersive X-ray spectrometer (EDX) attachment. High-resolution and elemental mapping analyses of the samples were operated under a high-angle annular dark field mode by scanning transmission electron microscope (STEM, FEI Tecnai G2 F30 S-TWIN) with an accelerating voltage of 300 kV. X-ray photoelectron spectroscopy (XPS) measurements were performed using an Axis Ultra DLD spectrometer (Kratos Analytical). The instrument was equipped with a monochromatized aluminium X-ray source powered at 15 kV and 3 mA, which delivered an X-ray beam of 300 μm × 700 μm. Charge compensation was obtained with the built-in charge neutralization system. The pass energy was set to 160 eV for the survey spectra and for the high-resolution spectra, and the pass energy of Al 2p, Si 2p and Y 3d are 40 eV, 40 eV and 20 eV, respectively. The binding energies were calculated with respect to the C component (BE = 284.8 eV) of the C 1s peak, and a Shirley background subtraction was used. For the spectrum fitting, an XPS analysis program XPS peak 4.1 was used, and a Lorentz–Gauss algorithm was applied. The fitted peak positions were compared with those reported in the literature and the full width half maximum (FWHM) of all the peaks was also used to evaluate the reasonableness of the fitting results. The photoluminescence (PL) and photoluminescence excitation (PLE) spectra were recorded on a Fluoromax-4 (JOBIN YVON) equipped with a 150 W xenon lamp as the excitation light source. A cut-off filter was used in the measurements to eliminate the second-order bands of the source radiation. The temperature-dependent emission spectra were recorded using a Fluoromax-4 spectrophotometer equipped with a heater (TAP-02, Orient Koji Scientific) in the sample compartment. The PL spectra of all the samples were recorded three times and averaged to reduce the error. The quantum efficiency was measured using the same fluorescent spectrometer (Fluoromax-4) with an integrating sphere (F3018) accessory. The electroluminescence spectra of phosphor-converted WLED were measured using a UV-VIS-near IR Spectrophoto Colormeter (PMS-80, EVERFINE PHOTO). All the measurements were performed at room temperature. The commercial phosphor YAG:Ce3+ was provided by HangZhou Ying-he Optoelectronic Technology Co., LTD (YH-Y552M). And the Blue LED chip (455 nm) was purchased from JiangXi Yuan-en Optoelectronic Technology Co., LTD. The chip size was 45 mil × 45 mil and the nominal voltage range was 3.4–3.6 V.
d. The strong diffraction peaks at 2θ = 33.4° and 18.2° are attributed to the (420) and (210) crystal planes, respectively. The peak marked with an asterisk belongs to the competition second phase of Y4.67(SiO4)3O (no. 30-1457), which is also observed as the second phase for the phosphor Lu2CaMg2(Si,Ge)3O12:Ce3+.14 Similar to the conclusions of Setlur,14 the phase of Y4.67(SiO4)3O with an apatite structure does not influence the photoluminescence properties of the phosphor Mg2Y2Al2Si2O12:Ce3+. In addition, the diffraction peaks of the host Mg2Y2Al2Si2O12 are very sharp and strong, suggesting that a good crystallinity of the phosphor can be obtained through this method. Also, this is very important for phosphor, because better crystallinity means less traps and a stronger luminescence.
![]() | ||
| Fig. 1 XRD patterns of the host Mg2Y2Al2Si2O12, standard card no. 088-2048 (Y3Al5O12) and no. 30-1457(Y4.67(SiO4)3O). | ||
The morphology was characterized by SEM, and it was found that the phosphors consist of many irregular granular microcrystals with particle sizes of 10–20 μm, as shown in Fig. S1.† EDX was used to determine the stoichiometric ratio of each element of the synthesized samples. The measurement results are shown in Table 1. The sample is mainly composed of Mg, Y, Al, Si and O, and the atomic ratio was in accordance with the theoretical composition 1
:
1
:
1
:
1
:
6 of the sample. A similar composition was detected for the different surface sites of the powder, indicating that each element is distributed uniformly in the sample. From the picture of the high resolution transmission electron microscopy (HRTEM), the interplanar distance between the adjacent lattice fringes were determined to be 0.27 nm and 0.5 nm (Fig. 2a). The corresponding 2θ value could be calculated by the Bragg's equation nλ = 2d
sin
θ, where λ is the X-ray wavelength, d is the distance between the two crystal planes, and θ is the diffraction angle of an observed peak. According to the equation, the calculated 2θ results were 33.2° and 17.7°, respectively, which are close to the 2θ value of the (420) and (210) crystal planes (2θ = 33.4° and 18.2°) obtained by XRD. These indicate that the lattice fringes shown in Fig. 2a belong to the (420) and (210) crystal planes. The energy dispersive X-ray mapping performed by the scanning transmission electron microscope was carried out to further confirm the elemental composition and the uniform distribution of the elements. As shown in Fig. 2b, the phosphor is composed of Mg, Y, Al, Si and O, and the distribution of Mg, Y, Al, Si and O elements is homogeneous in the phosphor, both of which agree well with the results measured by EDX.
| Element | Wt% | At% |
|---|---|---|
| Mg (K) | 6.75 | 7.56 |
| Al (K) | 10.03 | 10.13 |
| Si (K) | 11.06 | 10.73 |
| Y (L) | 34.32 | 10.51 |
| O (K) | 35.61 | 61.63 |
To illustrate the particular coordination sites of each element occupied in the garnet structure for the present novel host, XPS was used to confirm the chemical states of each element. The XPS survey scan is shown in Fig. 3a, suggesting that the host contains Mg, Y, Al, Si and O elements. Fig. 3b shows the high resolution XPS analysis of the variation in the Al 2p peak. The Al 2p peak is broad and can be roughly de-convoluted into two peaks, one at a lower binding energy, 73.7 eV, and the other at a slightly higher binding energy, 75.3 eV, which belong to [AlO4] tetrahedral and [AlO6] octahedral coordination, respectively.19–21 The Si 2p XPS peak is displayed in Fig. 3c with two peaks, which could be assigned to the Si 2p1/2 and Si 2p3/2 of [SiO4], proving that Si exists in the form of [SiO4] in the structure of Mg2Y2Al2Si2O12.22,23 In addition, the two peaks of Y 3d can be well fitted to the 3d orbit of the Y splitting of 3d3/2 and 3d5/2, indicating that the Y exists as one type of dodecahedral coordination in Mg2Y2Al2Si2O12.24 The XPS data of Mg2+ could be fitted into two peaks, i.e., two kinds of the chemical states, which occupy the left dodecahedral and the octahedral sites of the garnet structure. From the aforementioned, Mg2+ was incorporated into both dodecahedral Y3+ and octahedral Al3+ sites successfully, and the chemical formula of Mg2Y2Al2Si2O12 could be represented as {Y2Mg}[MgAl](AlSi2)O12, where the {}, [] and () denote dodecahedral, octahedral and tetrahedral coordination, respectively. In conclusion, the XPS results provide additional evidence for the formation of the garnet structure Mg2Y2Al2Si2O12.
![]() | ||
| Fig. 3 XPS survey scan (a) and high-resolution Al 2p (b), Si 2p (c), Y 3d (d) and Mg 2p (e) XPS spectra of the host Mg2Y2Al2Si2O12. | ||
Crystal structure refinement is a powerful tool to obtain detailed crystal structure information on Mg2Y2Al2Si2O12 and to ensure the phase purity of the sample. Based on the XPS results, the atom positions are set as following: Y3+ and half of Mg2+ occupy the dodecahedral sites, while the octahedral sites are occupied by the rest of Mg2+ and half of Al3+, and the tetrahedral sites are occupied by Si4+ and the rest of Al3+. Moreover, as no structural detail of Y4.67(SiO4)3O was reported, the crystallographic data of Gd4.67(SiO4)3O, which has the same structure as Y4.67(SiO4)3O, was used as the starting model.25
Fig. 4 shows the Rietveld refinement result of Mg2Y2Al2Si2O12, with a final converged weighted-profile of Rwp = 4.00%, by indexing to two phases of YAG and Gd4.67(SiO4)3O. The refinement results show that the main phase of the prepared sample is garnet, with a phase content of 95.6%. The crystallographic data, atomic coordinates and the isotropic parameters of the final refinement of Mg2Y2Al2Si2O12 and Y4.67(SiO4)3O are summarized in Table 2. The well-converged Rietveld refinement is further proof of the designed element occupancy and the formation of the garnet structure {Y2Mg}[MgAl](AlSi2)O12.
| Atom | x/a | y/b | z/c | S.O.F. | U (Å2) |
|---|---|---|---|---|---|
a Cubic; space group: Ia d; lattice parameters: a = 11.972 Å; V = 1715.76 Å3; α = β = γ = 90°; T = 298 K; Z = 8.b Hexagonal; space group: P63/m; lattice parameters: a = b = 9.3294 Å, c = 6.661 Å, V = 579.51 Å3; α = β = 90°, γ = 120°; T = 298 K; Z = 2. Cu Kα, λ = 1.5418; total reflections = 7171. |
|||||
| Mg2Y2Al2Si2O12a | |||||
| Al | 0 | 0 | 0 | 0.50 | 0.03418 |
| Mg | 0.25 | 0.125 | 0 | 0.33 | 0.01919 |
| Y | 0.25 | 0.125 | 0 | 0.67 | 0.01919 |
| Si | 0.25 | 0.375 | 0 | 0.67 | 0.03619 |
| O | 0.032 | 0.053 | 0.65 | 1.00 | 0.03563 |
| Al | 0.25 | 0.375 | 0 | 0.33 | 0.03619 |
| Mg | 0 | 0 | 0 | 0.5 | 0.03418 |
![]() |
|||||
| Y4.67(SiO4)Ob | |||||
| Y | 0.2270 | 0.2289 | 0.75 | 1 | 0.03233 |
| Y | 0.6667 | 0.3333 | 0 | 1 | 0.03458 |
| Si | 0.4714 | 0.3884 | 0.25 | 1 | 0.06273 |
| O | 0.3178 | 0.4872 | 0.25 | 1 | 0.06742 |
| O | 0.6002 | 0.4740 | 0.25 | 1 | 0.02195 |
| O | 0.3418 | 0.2497 | 0.0575 | 1 | 0.07056 |
| O | 0 | 0 | 0.25 | 1 | 0.04281 |
The garnet crystal structure of Mg2Y2Al2Si2O12 is shown in Fig. 5, which can be depicted as a network of octahedrons and tetrahedrons connected by sharing oxygen ions at the corners of the polyhedral. These polyhedrons are arranged in chains along the three crystallographic directions and form dodecahedral cavities, which are occupied by the Y3+ and Mg2+ ions. On the basis of the ionic radii in the Mg2Y2Al2Si2O12 lattice, the Ce3+ ions are considered to mainly occupy the distorted dodecahedral sites by substituting the Y3+ ions and as being coordinated by eight O2− ions. The interatomic distances in Mg2Y2Al2Si2O12 are listed in Table 3, and the changes versus that in YAG are in accordance with the ionic radii of the substitute element in each polyhedron. The decreasing Y–O bond distance could be attributing to the substitution of the dodecahedral sites with the smaller Mg2+ ion, as well as the compression effect from the bigger neighbouring octahedron. Furthermore, according to the interatomic distance, the difference of the bond length between the two kinds of Y–O in the host Mg2Y2Al2Si2O12 is greater than that in YAG, suggesting that the dodecahedron becomes more distorted and a red-shift of the Ce3+ emission hence is expected.
| Mg2Y2Al2Si2O12 | Y3Al5O12 | |
|---|---|---|
| a Bond distances in Å. | ||
| Y–O | 2.294 | 2.317 |
| Y–O | 2.417 | 2.437 |
| Al–O(6) | 1.992 | 1.938 |
| Al–O(4) | 1.715 | 1.754 |
| Mg–O(8) | 2.294 | |
| Mg–O(8) | 2.417 | |
| Mg–O(6) | 1.992 | |
| Si–O | 1.715 | |
Fig. 6 illustrates the XRD pattern of the phosphor Mg2Y2Al2Si2O12 with different doping concentrations of Ce3+ ion. It is obvious that the diffraction peaks of the samples concur with the peaks of Y3Al5O12. However, all the diffraction peaks slightly shift to low angles, manifesting that Ce3+ ion has been effectively doped into the matrix lattice of Mg2Y2Al2Si2O12. Because the replacement of Y3+ by the lager radius Ce3+ results in a bigger d-spacing, the diffraction peaks shift to the low angle according to Bragg's equation. Meanwhile, when the doping concentration of the Ce3+ ion increases, the diffraction peaks consistently shift to low angles (see the inset of Fig. 6), demonstrating that the more the Ce3+ ion is doped into the matrix lattice, the larger the cell parameters will become. Table 4 reveals such a relationship, as the lattice parameters and cell volumes increase gradually with increasing Ce3+ ions for all the samples, which were calculated by cell refinement fitting with GSAS on the basis of the XRD data.26
| Z = 0.01 | Z = 0.03 | Z = 0.06 | Z = 0.09 | Z = 0.12 | |
|---|---|---|---|---|---|
a Crystal system: cubic, space group: Ia d (no. 088-2048). |
|||||
| a = b = c (Å) | 11.973 | 11.975 | 11.976 | 11.980 | 11.982 |
| Cell volume (Å3) | 1716.36 | 1717.22 | 1717.65 | 1719.37 | 1720.24 |
| χ2 | 6.23 | 6.45 | 4.69 | 6.58 | 6.76 |
| Rwp (%) | 5.19 | 5.33 | 4.75 | 5.53 | 5.68 |
| Rp (%) | 3.47 | 3.61 | 3.30 | 3.70 | 3.81 |
The dependence of the emission intensities on the Ce3+ doping concentration is shown in the Fig. 8. The emission intensity increases with the increasing Ce3+ content, until it reaches a maximum at x = 0.06, and then the emission intensity declines when the concentration of Ce3+ goes beyond 0.06 (also see the inset of Fig. 8), because of concentration quenching. Thus, the optimal doping concentration of Ce3+ is 0.06. Furthermore, the inset of Fig. 8 also shows the relationship of the peak wavelength with the Ce3+ doping concentration. With the increasing of the doping concentration of Ce3+, the emission spectra gradually shift to the red side, with the wavelength shift from 584 nm for x = 0.01 to 617 nm for x = 0.12. This indicates that the crystal field splitting of the 5d energy level of the Ce3+ ions increases with the increasing Ce3+ concentration, which can be supported by the unit cell parameters in Table 4.
![]() | ||
| Fig. 8 Emission spectra of phosphors with different doping concentration. The inset displays the dependence of wavelength and the intensity of the emission on the concentration of Ce3+. | ||
The concentration quenching could be ascribed to the energy transfer via one Ce3+ ion to another Ce3+ ion, which often occurs as a result of a radiation reabsorption, exchange interaction, or electric multipolar interaction. The mechanism of the radiation reabsorption comes into effect only when there is a broad overlap of the spectra of the excitation and emission. As shown in Fig. 7, the overlap between the spectra of the excitation and emission is less than 2%, suggesting that the radiation reabsorption can be neglected. In this phosphor, when in low concentration, the average distance between the identical Ce3+ ions is so large that energy migration does not occur. This critical distance is of great importance, and can be calculated with the following equation:
![]() | ||
| Fig. 9 Temperature-dependent photoluminescence of Mg2Y1.94Al2Si2O12:0.06Ce3+ under excitation of 460 nm. The inset displays the dependence of the emission intensity on the increasing temperature. | ||
Also, the emission intensity of YAG:Ce3+ declines to 89% when the temperature increases to 160 °C. The thermal stability of YAG:Ce3+ is better than that of Mg2Y2Al2Si2O12:Ce3+. The quenching activation energy (Ea – energy barrier for thermal quenching) reduces when the Stokes shifts between 4f–5d absorption and then the 5d–4f emission gradually increases.15 Because the Stokes shift of Mg2Y2Al2Si2O12:Ce3+ is larger than that of YAG:Ce3+, thermal quenching is more likely to occur.
The CIE chromaticity diagram for the sample Mg2Y1.94Al2Si2O12:0.06Ce3+ under excitation at 460 nm is shown in Fig. 10. The colour coordinate of the present phosphor is (x = 0.519, y = 0.472), which is located in the yellow-orange region, indicating that warm light with low CCT could be expected when coupling this yellow-orange light emitting phosphor with a blue LED chip. The quantum efficiency (QE) of phosphor is considered to be another important parameter for practical application, and the main two factors leading to low quantum efficiency are concentration quenching and thermal quenching. The quantum efficiency data were measured for the phosphors with different Ce3+ doping concentrations. Also, the quantum efficiency of phosphor Mg2Y1.94Al2Si2O12:0.06Ce3+ was measured to be 57% at room temperature. As a reference, the quantum efficiency value of the commercial yellow phosphor YAG:Ce3+ was measured with the same instruments, and the value was found to be 73% under 460 nm excitation. Many factors will affect the QE, such as the excitation wavelength, the synthesis conditions and particle size, and also further optimizing the synthesis conditions is necessary to improve the QE.
![]() | ||
| Fig. 10 CIE coordinates of the phosphor Mg2Y1.94Al2Si2O12:0.06Ce3+ and the commercial phosphor YAG:Ce3+. | ||
![]() | ||
| Fig. 11 Electroluminescent spectra of Mg2Y1.94Al2Si2O12:0.06Ce3+ phosphor-converted WLEDs under various drive currents from 100 mA to 500 mA. | ||
| Current (mA) | Coordinate | CCT (K) | CRI (Ra) | Luminous efficacy (lm W−1) |
|---|---|---|---|---|
| 100 | (0.3493, 0.3129) | 4620 | 81.0 | 27.54 |
| 150 | (0.3466, 0.3097) | 4730 | 81.2 | 26.31 |
| 200 | (0.3436, 0.3061) | 4860 | 81.6 | 23.42 |
| 250 | (0.3414, 0.3035) | 4967 | 82.0 | 20.47 |
| 300 | (0.3389, 0.3001) | 5091 | 82.4 | 18.30 |
| 350 | (0.3368, 0.2975) | 5206 | 82.7 | 17.68 |
| 400 | (0.3343, 0.2946) | 5356 | 83.3 | 15.87 |
| 450 | (0.3306, 0.2896) | 5599 | 84.2 | 15.20 |
| 500 | (0.3279, 0.2863) | 5796 | 84.7 | 13.56 |
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c4ra14425b |
| This journal is © The Royal Society of Chemistry 2015 |