DOI:
10.1039/C6RA13029A
(Paper)
RSC Adv., 2016,
6, 82824-82831
Orange- to green-emitting Li(Sr,Ca)4(BO3)3:Eu2+ phosphor: emission-tunable properties and white light emitting diode application†
Received
19th May 2016
, Accepted 31st July 2016
First published on 1st August 2016
Abstract
An emission-tunable phosphor, Eu2+-activated LiSr(4−x−y)Cax(BO3)3:yEu2+ phosphor, was synthesized by high temperature solid state reaction. By combining with Reitveld refinement analysis, the X-ray powder diffraction test confirmed the formation of a crystalline phase. The optical spectral properties, concentration quenching properties, decay curves and thermal stability of the LiSr(4−x−y)Cax(BO3)3:yEu2+ phosphor were investigated in detail. The results showed that with the Ca2+ doping concentration increasing, the emission maxima is shifted from 618 to 501 nm and the color hue can be tuned from orange red to blue green. The effects of Ca2+ substitutions into Sr2+ sites on the emission band change were investigated. A white LED was fabricated using a blue 450 nm chip pumped by LiSr3.92Ca0.06(BO3)3:0.02Eu2+ phosphor film. Under 100 mA driven currents, the white LED device has a Commission International de I'Eclairage (CIE) color coordinate of (0.3563, 0.3576) at a white light (correlated color temperature = 5083 K) and an excellent color rendering indices of 85.3.
1. Introduction
In recent years, the white light-emitting diode (w-LED) has received wide attention from all over the world as a new generation of light source.1–3 It has many advantages, such as, high luminous efficiency, small volume, saving energy and so on. At present, the major commercial white LEDs are phosphor-converted LEDs which consist of a blue InGaN chip and a yellow phosphor, (Y,Gd)3(Al,Ga)5O12:Ce3+. Unfortunately, the w-LEDs packaged with yellow phosphors exhibit high color temperature and poor color rendering index (CRI), due to the lack of red components.4 To improve the rendering index, w-LEDs that use a near ultraviolet (n-UV) LED chip with three primary color emission phosphors have been widely investigated. The w-LEDs have several benefits, including high color rendering index, high luminous efficiency, and a stable light color that are almost independent of the changed current. However, compared with 1-pc-LED (one phosphor-converted LED), mixing powders and finding high-efficiency compounds are more difficult for this type of device. Furthermore, low efficiency of the red phosphors and the need for complex coating technology, restrict their development.5
Recently, several w-LEDs modules fabricated using n-UV chip coupled with a blend of tunable emitting (blue-to-yellow) and red-emitting phosphor. This combination mode presented favourable properties, for example, excellent CRI values, tunable CIE chromaticity coordinates and tunable correlated color temperature.6–10 Phosphor with tunable colors from the same luminescent component are highly desirable. As we know, the emission colour tuning range of the phosphor with a single activated ion is only limited to blue to yellow. In order to broaden the tuning range, the realization of color tunable phosphors from red to yellow has been a challenge.
As orange-emitting borate phosphor, LiSr4(BO3)3:Eu2+ phosphor has been reported in the literatures.11–13 To the best of our knowledge, the tunable luminescence properties of LiSr4(BO3)3:Eu2+ phosphor has not yet been reported in literature. In this study, we investigated the tunable luminescent properties of Eu2+-doped LiSr4(BO3)3:Eu2+ by alkali-earth metal ions (Sr2+ and Ca2+) mutual substitution in detail. The phosphor can be excited by n-UV and blue LED chips and shows tunable orange red to blue green emission. In addition, the LED device using Li(Sr,Ca)4(BO3)3:Eu2+ phosphor with a n-UV chip were fabricated to demonstrate the potential applications of Li(Sr,Ca)4(BO3)3:Eu2+ phosphor as a novel color conversion material.
2. Experimental
2.1 Materials and synthesis
The LiSr4−x−yCax(BO3)3:yEu2+ phosphor samples were synthesized by a traditional high-temperature solid-state reaction method. The SrCO3 (99.99%), CaCO3 (99.99%), H3BO3 (99.99%), Li2CO3 (99.9%) and Eu2O3 (99.999%) were used as beginning materials. According to the stoichiometric ratio, the beginning materials were weighed, and the mixed thoroughly in an agate mortar. Finally the mixture was sintered at 950 °C with reducing atmosphere (5% H2/95% N2) for 5 h in a tube furnace.
2.2 Materials characterization
The X-ray diffraction (XRD) pattern was collected on an X-ray diffractometer (Bruker Axs D2 PHASER diffractometer). Rietveld refinenements on the X-ray diffraction data were performed using the software TOPAS. The excitation and emission spectrum of phosphors were recorded by a PL3-211-P spectrometer (HORIBA JOBIN YVON, America) and a 450 W xenon lamp was used as the excitation source. The diffuse reflectance spectrums of the undoped and doped Eu2+ samples were tested by a UV-3600 UV-Vis spectrometer (Shimadzu, Japan). The spectral, photometric and colorimetric quantities of the LEDs were measured by a HAAS-2000 Light & radiation measuring instrument for packaged LEDs (Everfine, China).
2.3 LED lamp fabrication
We mixed the LED packaging epoxy resin and LiSr3.92Ca0.06(BO3)3:0.02Eu2+ phosphor together, then spread the mixture on a substrate which can be used to control the shape and thickness of film. And then, the film was placed into an oven and baked at 60 °C for 4 h. Finally a luminescent thin film was obtained, and packaged on a 450 nm blue LED chip.
3. Results and discussion
3.1 Crystal structure
The phase purity and structure change of all phosphor samples were verified by XRD analysis. Fig. 1 shows XRD pattern for a Li(Sr3.98−xCax)(BO3)3:0.02Eu2+ sample as a function of substituted Ca2+ ions content (x). All the diffraction peaks of the obtained samples can be indexed to the standard data of LiSr4(BO3)3 with the ICSD file no. 170861, indicating that these samples are single-phase in our experimental doping-content ranges and that the doping ions Ca2+/Eu2+ are incorporated in the host lattice. Additionally, the diffraction peaks exhibit a slight shift toward higher angles with increasing content of the Ca2+ ion, which may be related to the substitution of the larger Sr2+ by the smaller Ca2+.
 |
| Fig. 1 XRD patterns of Li(Sr3.98−xCax)(BO3)3:0.02Eu2+ samples. The standard data for LiSr4(BO3)3 (ICSD-170861) is shown as a reference. | |
Fig. 2 showed the Rietveld refinement results of LiSr3.98(BO3)3:0.02Eu2+ phosphors. ICSD 170861 (LiSr4(BO3)3) was performed as the standard data to refine the experimental crystal structures. LiSr3.98(BO3)3:0.02Eu2+ crystallized in a cubic unit cell with the same space group Ia
d(230). For Li(Sr3.98Eu0.02)(BO3)3, the lattice constants (a, b, c) are 15.0710 (Å) and cell volume (V) is 3423.15 (Å3). The detailed refinement results were showed in Table 1.
 |
| Fig. 2 Observed (black), calculated (red), and difference (green) XRD profiles for the Rietveld refinement of LiSr3.98(BO3)3:0.02Eu2+. Bragg reflections are indicated with tick marks. | |
Table 1 Rietveld refinement and crystal data of LiSr3.98(BO3)3:0.02Eu2+ phosphors
Compound |
x = 0 |
Formula |
LiSr3.98(BO3)3:0.02Eu2+ |
Cryst. syst. |
Cubic |
Space group |
La d(230) |
Crystal density (g cm−3) |
6.642 |
Unit, Z |
16 |
a (Å) = b (Å) = c (Å) |
15.0710 |
V (Å3) |
3423.15 |
Rexp (%) |
4.61 |
Rwp (%) |
5.13 |
Rp (%) |
4.46 |
GOF |
1.57 |
The results of structure refinement show that there are two crystallographically different sites for Sr atoms in the crystal structure of Li(Sr3.98Eu0.02)(BO3)3. The Sr(1) atom occupies the center of octahedral SrO6 polyhedron (CN = 6) and Sr(2) in the center of hexagonal bipyramidal SrO8 polyhedral (CN = 8) (see the inset in Fig. 2). The ionic radius of Sr2+ was 1.18 Å (CN = 6) and 1.26 Å (CN = 8). The ionic radii of Eu2+ were 1.17 Å (CN = 6) and 1.25 Å (CN = 8). Because of the similarity in ionic radius, the Sr2+ ions sites would be occupied randomly by the Eu2+ ions in the LiSr4(BO3)3 host. Meanwhile, when Eu2+ ions occupied Sr2+ ions sites, two different types of emission spectra could be obtained.
The unit cell constants a and V of Li(Sr3.98−xCax)(BO3)3:0.02Eu2+ samples were calculated by refining the powder XRD data as a function of the Ca2+ content. It can be seen from Fig. 3 that the lattice constants a and V decrease linearly when x is gradually increased from 0 to 0.15, which is in good agreement with Vegard's law.14 The change in lattice constants also indicates that the Ca2+ ions are doped into the host and that the solid solution is well formed in the LiSr4(BO3)3 crystal.
 |
| Fig. 3 Unit cell parameters of LiSr3.98−xCax(BO3)3:0.02Eu2+ samples obtained from Rietveld refinement data present both a contraction in the (a) lattice parameter and (b) cell volume as the concentration of Ca2+ increases, reflecting the effects of the cations substitutions. | |
3.2 Diffuse reflection spectra
The diffuse reflection spectra of the undoped LiSr4−xCax(BO3)3 powders were illustrated in Fig. 4(a). In the visible range (400–700 nm) the sample of LiSr4−xCax(BO3)3 showed a high reflection. The undoped LiSr4−xCax(BO3)3 samples showed absorption band between about 200 and 425 nm range. The absorption edge shifted toward long wavelength with the increasing of Ca2+ ions content.
 |
| Fig. 4 (a) Diffuse reflection spectra of undoped LiSr4−xCax(BO3)3 samples, (b) absorption spectra of LiSr4−xCax(BO3)3 (x = 0 and 1.5) as calculated by the Kubelka–Munk formula. | |
The optical band gap of LiSr4−xCax(BO3)3 hosts had been calculated by following Kubelka–Munk function:15
|
F(R) = K/S = (1 − R)2/2R
| (1) |
here,
S,
K and
R are the scattering coefficient, absorption coefficient and reflection coefficient, respectively. The absorption spectra of LiSr
4−xCa
x(BO
3)
3 derived with the Kubelka–Munk function are shown in
Fig. 4(b). The value of the optical band gap is calculated to be from 5.35 to 5.85 eV with the increasing content of Ca
2+ ions in the LiSr
4−xCa
x(BO
3)
3 host by extrapolating the Kubelka–Munk function to
K/
S = 0.
The diffuse reflectance spectra of LiSr3.4−yCa0.6(BO3)3:yEu2+, y = 0–0.08 are given in Fig. 5, the strong absorption band is presents at 360 nm, and one more weak absorption band is observed at 310 and 400 nm; all the bands are attributed to the 4f7–4f65d1 electronic transitions of Eu2+. These bands are absent in the diffuse reflectance spectrum of the parent phase (unsubstituted LiSr3.4Ca0.6(BO3)3). The band at 260 nm observed for all compositions is due to host absorption. Also, the absorption edge gradually extends to longer wavelengths, and the absorption gets enhanced with higher Eu2+ concentration.
 |
| Fig. 5 Diffuse reflectance spectra of LiSr3.4−yCa0.6(BO3)3:yEu2+ (y = 0–0.10). | |
3.3 Excitation and emission properties
Fig. 6 depicts the excitation and emission spectra of LiSr3.38Ca0.6(BO3)3:0.02Eu2+ phosphor, and the deconvoluted Gaussian components. The excitation spectrum of LiSr3.38Ca0.6(BO3)3:0.02Eu2+ monitored by 534 and 585 nm presents a broad hump between 300 and 450 nm, respectively, except for the difference of peak shapes and intensities, which means the phosphor can be well excited by the n-UV LED chips. The excitation spectra contains two excitation bands peaking at 360 (380) and 410 nm, which can be attributed to the 4f7–4f65d1 transitions of Eu2+ from the ground state to the different splitting levels of the 5d state.
 |
| Fig. 6 Excitation (left) and emission (right) spectra of LiSr3.38Ca0.6(BO3)3:0.02Eu2+ phosphors, (a) λex = 515 nm, λem = 360 nm, (b) λex = 588 nm, λem = 380 nm, fitted curve (red dashed line) and deconvoluted Gaussian components (green solid lines). | |
Upon 360 nm (27
777 cm−1) and 380 nm (26
315 cm−1) excitation, the phosphor shows the different broad band emission centered at 534 nm (18
726 cm−1) and 585 nm (17
094 cm−1), respectively, which is attributed to the electric-dipole-allowed transition from the lowest level of the 5d excited state (4f65d) to the 4f ground state (4f7(8S7/2)) of the doped Eu2+ ions. The unsymmetrical emission spectrum can be deconvoluted into two Gaussian peaks centered at 515 nm (19
417 cm−1) and 588 nm (17
006 cm−1) (shown as two green dotted curves in Fig. 6), respectively.16 It is also observed that the excitation spectrum monitoring by 515 and 588 nm exhibits a different spectral profile, which means that the two emission peaks should be ascribed to two different Sr2+ sites occupied by Eu2+ ions, which is also in accordance with the above analysis on the crystal structure.
The decay curves of LiSr3.38Ca0.6(BO3)3:0.02Eu2+ phosphors are showed in Fig. 7, the corresponding luminescence decay time could be best fitted with a double exponential functions as I = A1
exp(−t/τ1) + A2
exp(−t/τ2), where I is the luminescence intensity, A1 and A2 are constants, t is the time, τ1 and τ2 are rapid and slow lifetimes for exponential components, respectively. The average decay times were calculated to be 411.36 ns (monitored at 517 nm) and 523.58 ns (monitored at 587 nm). The double exponential decay dynamic is in agreement with the fact that two types of cation sites exist in the host lattice (Fig. 8).
 |
| Fig. 7 Decay curves of Eu2+ emission in LiSr3.92Ca0.06(BO3)3:0.02Eu2+ phosphor monitored at (a) 517 nm and (b) 587 nm. | |
 |
| Fig. 8 The 5d energy levels of free Eu2+, Eu2+(1) and Eu2+(2) in LiSr4(BO3)3. | |
According to the report of Van Uitert, the energy (E) of the lower d-band edge for Eu2+ can be estimated using the following equation:17,18
|
 | (2) |
where
E represents the real position of the d band edge in energy for the Eu
2+ ion (cm
−1),
Q represents the position in energy for the lower d-band edge of the Eu
2+ free ions (34
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
000 cm
−1),
V represents the valence of the Eu
2+ ion,
n represents the number of anions in the immediate shell about this ion,
Ea represents the electron affinity of the atoms that form anions,
r is the radius of the host cation replaced by the Eu
2+ ion (in Å). There are two Sr
2+/Ca
2+ sites in LiSr
3.92Ca
0.06(BO
3)
3:0.02Eu
2+ phosphor. It is apparent that the value of
E (cm
−1) is directly proportional to the product of
n and
r. Thus, we can conclude that the band peaked at 515 nm (19
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
417 cm
−1) is ascribed to the 4f
65d
1 → 4f
7 transition of Eu
2+ ions occupying the eight-coordinated Sr1/Ca1O8 site, while the other emission peaks centered at 588 nm (17
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
006 cm
−1) attribute to Eu
2+ occupying six-coordinated Sr2/Ca2O6 ions site, respectively.
The energy of the lowest 5d excited level of Eu2+ is influenced by centroid shift (εc) and crystal field splitting (εcfs). The 5d centroid shift for Eu2+ can be expressed by the following equation19
|
 | (3) |
where
Ri is the distance (pm) between Eu
2+ and anion
i in the undistorted lattice, 0.6Δ
R is correction for lattice relaxation around Eu
2+, and Δ
R is the difference between the radii of Eu
2+ and cation sites that Eu
2+ ions occupy. The summation is over all
N anions that coordinate Eu
2+.
αspi (in units of 10
−30 m
−3) is the spectroscopic polarizability of anion
i. For oxides,

,
χav is the electro-negativity of the cations. For LiSr
4(BO
3)
3 χav = [
χ(Li) + 8
χ(Sr) + 9
χ(B)]/18 = 1.36, and one obtains
αOsp = 2.93 × 10
−30 m
−3. The
εc value of Eu
2+(1) and Eu
2+(2) are calculated to be 1.36 and 1.20 eV, respectively.
The crystal field splitting of the 5d levels can be expressed as20
where
βpolyQ is a constant that depends on the type of coordination polyhedron,
Q is 2+ for Eu
2+ ions. The ratio
βpoly(6)2+![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
:
βpoly(8)2+![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
:
βpoly(12)2+ equals 1
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
:
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
0.89
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
:
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
0.42, namely,
βpolyQ decreases with increasing coordination number. The value of
βpoly(6)2+ for Eu
2+ is 135
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
466 eV pm
2.
20 The value of
βpoly(8)2+ for 8 O-coordinated Sr
2+ polyhedron is estimated to be 0.89 times of
βpoly(6)2+ according to the above ratio.
Rav is close to the average distance between anions and cations that is replaced by Eu
2+, which can be expressed as follow:
|
 | (5) |
The
Rav values for Sr1–O and Sr2–O in LiSr
4(BO
3)
3 are 261.5 and 253.0 pm, respectively. Therefore, the crystal field splitting values for 5d levels of Eu
2+(1) and Eu
2+(2) are 1.76 and 2.12, respectively. That is to say, Eu
2+(1) have smaller
εcfs than that of Eu
2+(1). Therefore, the lowest excited 5d level of Eu
2+(1) is at a higher energy (corresponding to a shorter emission wavelength, 515 nm) than that of Eu
2+(2) (588 nm).
Fig. 7 shows the 5d energy levels of free Eu
2+ and Eu
2+ on Sr sites based the above analyses.
In addition, the quantum efficiency of the phosphor was recorded using an integrating sphere attached to the PL3-211-P spectrometer, the internal (ηi) and external (ηex) quantum efficiencies (QEs) of the LiSr3.92Ca0.06(BO3)3:0.02Eu2+ phosphor were calculated using following equations:21
|
 | (6) |
|
 | (7) |
where Ex(
λ) and
Re(
λ) are the spectra of the excitation light without and with the phosphor in the integrating sphere, respectively; Em(
λ) is the emission spectrum of the phosphor in the integrating sphere. The
ηi and
ηex values of LiSr
3.92Ca
0.06(BO
3)
3:0.02Eu
2+ under 360 nm excitation are calculated to be ∼53.7% and ∼36.5%, respectively.
Fig. 9 shows the concentration dependence of the integrated emission intensity of LiSr3.92−xCa0.06(BO3)3:xEu2+ with various concentration of Eu2+ ions under 360 and 380 nm excitation, respectively. For LiSr3.92−xCa0.06(BO3)3:xEu2+ phosphors, the emission intensity increases with the increase in Eu2+ content up to a maximum x value of about 0.02 mol, and then it decreases because of interaction between Eu2+ ions. According to the percolation model, concentration quenching of the compound can occur by two mechanisms: (1) interactions between the Eu2+ ions, which results in energy reabsorption among neighboring Eu2+ ions in the rare earth sublattice; (2) energy transfer from a percolating cluster of Eu2+ ions to killer centers.
 |
| Fig. 9 (a) Concentration dependence of the relative integrated emission intensity of LiSr3.92−xCa0.06(BO3)3:xEu2+ (x = 0.005–0.08) under 380 and 360 nm excitation, respectively. (b) Relationships of log(I/xEu2+) versus log(xEu2+) in LiSr3.92−xCa0.06(BO3)3:xEu2+ phosphor. | |
Hence we have calculated the critical distance between the Eu2+ ions for energy transfer by using the relation given by Blasse22
|
 | (8) |
where
V is the volume of the unit cell,
xc is the critical concentration of activator,
Z is the number of formula units per unit cell. For LiSr
4(BO
3)
3 host, using
Z = 16,
xc = 0.02 and
V = 3243.3 Å
3. The obtained
Rc is 26.84 Å. For the exchange interaction, the critical distance is about 5 Å, which is much smaller than the obtained result (26.84 Å), thus, the exchange interaction between the Eu
2+ ions may be ruled out in LiSr
4(BO
3)
3 host.
The non-radiative energy transfer between Eu2+ ions is caused by electric multipole–multipole interactions, which depend on distance according to Dexter's theory.23 The strength of the multipole–multipole interactions can be determined from the change in the emission intensity if the energy transfer occurs between the same types of activators. The emission intensity (I) per activator ion is expressed by the formula given in the following:24,25
|
 | (9) |
here
x is Eu
2+ ions concentration above the quenching contents;
k and
β are constants for a given host crystal; and
θ is the energy transfer parameter.
θ = 3 correspond to the energy transfer among the nearest-neighbor ions, in addition,
θ = 6, 8, or 10 correspond to dipole–dipole, dipole–quadrupole, or quadrupole–quadrupole interactions, respectively.
Fig. 9(b) presents that the relationships of log(
I/
xEu2+)
versus log(
xEu2+) is linear and the slope is −1.8. By using the
eqn (9), the value of
θ was 5.4, which is approximately equal to 6. The result indicates that the dipole–dipole is the main mechanism of concentration quenching of the central Eu
2+ emission in LiSr
3.92−yCa
0.06(BO
3)
3:
yEu
2+ phosphors.
Fig. 10 shows the normalized emission spectra of LiSr3.98−xCax(BO3)3:0.02Eu2+ under the 360 nm excitation. From the spectrum can be observed that there is an obvious blue shift from 618 to 501 nm. At the same time, the peak of excitation spectra is shifted to the short wavelength (Fig. S2†). The peaks of excitation and emission wavelength, Stokes shift, crystal field splitting and the full-width at half maximum (FWHM) for LiSr3.98−xCax(BO3)3:0.02Eu2+ are summarized in Table S1 (ESI†). From Fig. 6 and 10, two emission peaks should be ascribed to two different sites occupancy of Eu2+ ions. In order to analyze the occupancy of Eu2+ after entering LiSr4(BO3)3 host, Gauss imitating decomposition was undertaken successfully to the emission spectra of LiSr3.98−xCax(BO3)3:0.02Eu2+ phosphors (see Fig. S2†). The peak value and normalized intensity of the two emission bands are summarized in Table S2 (ESI†). The results indicated that the two emission bands are blue shifted with the content of Ca2+ increased in the LiSr3.98−xCax(BO3)3:0.02Eu2+. But the intensity changes of the two emission bands are not consistent: the intensity of the short wavelength emission band increases gradually, while the long wavelength emission band is just the opposite. The intensity and peak position of the sub emission bands were obviously changed, and the final performance was the blue shift of the whole emission.
 |
| Fig. 10 Normalized emission spectra of LiSr3.98−xCax(BO3)3:0.02Eu2+ (λex = 360 nm). | |
With the introduction of Ca2+ substituted Sr2+ ions, the crystal structure of the LiSr4(BO3)3 system would change to a certain extent, thus, the local environment of the Eu2+-substituted sites are significantly change due to bond-length changes between the activator cation and the ligand anion, according to the equation:
|
 | (10) |
where
Dq is the crystal field strength,
Z is the charge or valence of the anion,
e is the charge of the electron,
r is the radius of the
d wavefunction and
R is the distance between the central ion and its ligands. The gradually replace of Sr
2+ by the smaller Ca
2+ ions is expected to be followed by a shorter metal–ligand distance, leading to a stronger crystal field environment, which can be proved by the decreasing lattice parameters and volumes, as shown in
Fig. 3. Thus, there is theoretical causing a red shift in the emission, meanwhile, decreasing the emission energy. Out of our expectation, there are such abnormal blue shift emission properties of the sub emission bands which should be worth further study.
From the crystal structure of LiSr4(BO3)3, the centre Sr2+ ions are surrounded by coordinated O2− ions which also linked with other neighboring cations, such as B3−, and outer Sr2+ ions. When an activator ion was introduced, Eu2+ would randomly occupy the Sr2+ site in LiSr4(BO3)3, thus, causing the orange emitting. When Sr2+ are replaced by smaller Ca2+ at the nearest neighbour sites of Eu2+, blue shift is evidently shown in emission spectrum, suggesting that the introduction of Ca2+ ions could change the crystal field of Eu2+ ions. As for how to change, the ionic potential should be considered. The attractive force of the central cations towards the anions can be estimated using the following formula:26
|
 | (11) |
where
Z is the electric charge number of ion, and
r is the ion radius (pm). As is known, the ionic radius of Ca
2+ is smaller than Sr
2+. Thus; Ca
2+ has the higher attractive force towards O
2− than the Sr
2+ ions. When Sr
2+ is substituted by the smaller Ca
2+ ion, the coordinated O
2− ions could be attracted by strong attraction, then closer to Ca
2+ ions. As a result, the bond length between Eu
2+ and O
2− becomes longer, then the magnitude of the crystal field strength decreases, thus, the blue shift occurs in the sub emission bands with the increase of the content of Ca
2+ in LiSr
3.98−xCa
x(BO
3)
3:0.02Eu
2+.
Fig. 11 shows the temperature dependence of emission intensity for LiSr3.98−xCax(BO3)3:0.02Eu2+ phosphors. With the increasing of the Ca2+ contents in LiSr3.98−xCax(BO3)3:0.02Eu2+ phosphors, the thermal stability of the phosphors increase gradually. Furthermore, the activation energy ΔE of the activator ion, can be expressed by:27,28
|
 | (12) |
here
k is the Boltzmann constant (8.629 × 10
−5 eV K
−1),
A is a constant,
I0 and
IT are the luminescence intensity of at room temperature and the resting temperature. From
eqn (12), Δ
E was obtained to be 0.162 eV for LiSr
3.98(BO
3)
3:0.02Eu
2+ and 0.276 eV for LiSr
2.48Ca
1.5(BO
3)
3:0.02Eu
2+. The calculated value Δ
E of LiSr
2.48Ca
1.5(BO
3)
3:0.02Eu
2+ phosphor is bigger than that of LiSr
3.98(BO
3)
3:0.02Eu
2+ phosphor. Therefore, the thermal stability of LiSr
2.48Ca
1.5(BO
3)
3:0.02Eu
2+ phosphor is better than LiSr
3.98(BO
3)
3:0.02Eu
2+ phosphor.
 |
| Fig. 11 The relationship between temperature and emission intensity of the LiSr3.98−xCax(BO3)3:0.02Eu2+ phosphors. | |
The difference in thermal stability is related to the change in the Stokes shift, which can be explained by the configuration coordinate diagram in Fig. 12. R represents the Stokes shift, which is in close relation with the interatomic distance between the Eu2+ and the coordination anions. After absorption of the excitation energy, a non-radiative relaxation occurs at the bottom of the excited state, and then emission takes place by radiative transitions returning to ground state. However, under a high temperature, thermal activation can happen due to the electron–phonon coupling and the energy reaches the crossing point (C) between the excited and ground states. In this case, non-radiative relaxation occurs by heat dissipation rather than radiation emission, which could quench the luminescence. With the decreasing in the Stokes shift, the thermal activation energy increases, that means that a smaller Stokes shift results in a lower thermal quenching behaviour. Therefore, doping of Ca2+ is beneficial to improve the thermal stability of the LiSr3.98(BO3)3:0.02Eu2+ system.
 |
| Fig. 12 Configuration coordinate diagram of LiSr3.98−xCax(BO3)3:0.02Eu2+ phosphor. | |
Fig. 13 shows the CIE chromaticity coordinates for the LiSr3.98−xCax(BO3)3:0.02Eu2+ phosphors with different x values. The color hue can be tuned from orange red (point 1) to blue green (point 8), and the corresponding chromaticity coordinates (x, y) varying from (0.5524, 0.4270) to (0.1895, 0.4718) summarized in the inset table. The inset photos show luminescence of LiSr3.98−xCax(BO3)3:0.02Eu2+ (0 ≤ x≤ 1.5) phosphors under excitation at 360 nm.
 |
| Fig. 13 Variation in CIE chromaticity coordinates as function x in LiSr3.98−xCax(BO3)3:0.02Eu2+. | |
3.4 Application to white LEDs
Fig. 14 shows the electroluminescence (EL) spectrum of the white LED which is packaged from the LiSr3.92Ca0.06(BO3)3:0.02Eu2+ phosphor film and a 450 nm blue LED chip under a applied current of 100 mA. The inset shows the photograph of (a) LiSr3.92Ca0.06(BO3)3:0.02Eu2+ phosphor film, (b) the phosphor film excited at 365 nm in a UV box, (c) as-fabricated LED and (d) LED in operation, respectively. The white LED had CIE color coordinates of (0.3563, 0.3576) with a color-rendering index of 85.3 around the correlated color temperature of 5083 K which is higher than a YAG:Ce – based white LED (Ra = 75–78) at 20 mA,29 and β-Y0.95Ce0.05FS-based white LED (Ra = 66–77) at 50–250 mA,30 and is similar to γ-Ca2SiO4:Ce3+,Li+-based white LEDs (Ra = 86).31 The white LEDs exhibited a luminous efficacy (ηL) of 22.4 lm W−1. This result confirms that the LiSr3.92Ca0.06(BO3)3:0.02Eu2+ have potential for application in white LED.
 |
| Fig. 14 Electroluminescence spectrum of white LED using (Ca3.85Ba0.10)(PO4)2O:0.05Eu2+ under applied currents 100 mA. | |
4. Conclusions
In summary, we have synthesized the emission-tunable phosphor Eu2+-doped Li(Sr,Ca)4(BO3)3 by high temperature solid state method and investigated its tunable luminescence performance. The results show that there are the different broad band emission centered at 534 and 585 nm in Li(Sr,Ca)4(BO3)3:Eu2+ phosphors, respectively, and its emission can be adjusted from orange red to blue green by changing Ca2+ doping concentration. Applying a blend of LiSr3.92Ca0.06(BO3)3:0.02Eu2+ phosphor film on blue chip, we can obtain a white LED device with high CRI value of 85.3 and CCT value of 5083 K. With the interesting tunable emission property, Li(Sr,Ca)4(BO3)3:Eu2+ phosphor has great application potential as a good color conversion material for solid state lighting.
Acknowledgements
This research is supported by Zhejiang Provincial Natural Science Foundation of China under Grant No. LR14A040002, LQ13E020003 and LZ14F050001, the Project of the National Nature Science Foundation of China under Grant No. 51402077, 51272243, 61370049 and 61405185.
Notes and references
- E. Schubert and J. Kim, Science, 2005, 308, 1274 CrossRef CAS PubMed.
- Y. Narukawa, J. Narita, T. Sakamoto, T. Yamada, H. Narimatsu, M. Sano and T. Mukai, Phys. Status Solidi A, 2007, 204, 2087 CrossRef CAS.
- Z. Xia, Z. Xu, M. Chen and Q. Liu, Dalton Trans., 2016, 45, 11214 RSC.
- A. Setlur, W. Heward, M. Hannah and U. Happek, Chem. Mater., 2008, 20, 6277 CrossRef CAS.
- C. C. Lin and R.-S. Liu, J. Phys. Chem. Lett., 2011, 2, 1268 CrossRef CAS PubMed.
- A. Setlur, V. E. Radkov, S. C. Henderson, J. Her, M. A. Srivastava, N. Karkada, M. Kishore and N. Kumar, Chem. Mater., 2010, 22, 4076 CrossRef CAS.
- W. Im, S. Brinkley, J. Hu, A. Mikhailovsky, S. DenBaars and R. Seshadri, Chem. Mater., 2010, 22, 2842 CrossRef CAS.
- Y. Wu, D. Wang, T. Chen, C. Lee, K. Chen and H. Kuo, ACS Appl. Mater. Interfaces, 2011, 3, 3195 CAS.
- H. Yu, D. Deng, D. Zhou, W. Yuan, Q. Zhao, Y. Hua, S. Zhao, L. Huang and S. Xu, J. Mater. Chem. C, 2013, 1, 5577 RSC.
- X. Zhou, X. Yang, Y. Wang, X. Zhao, L. Li and Q. Li, J. Alloys Compd., 2014, 608, 25 CrossRef CAS.
- X. Zhang, L. Zhou and M. Gong, J. Biolumin. Chemilumin., 2014, 29, 104 CrossRef CAS PubMed.
- G. Li, Y. Wang, W. Zeng, S. Han, W. Chen, Y. Li and H. Li, Opt. Mater., 2014, 36, 1808 CrossRef CAS.
- Y. Jin, Y. Hu, H. Wu, L. Chen and X. Wang, J. Lumin., 2016, 172, 53 CrossRef CAS.
- W. D. Kingery, H. K. Bowen and D. R. Uhlmann, Introduction to Ceramics, Wiley, New York, 1976 Search PubMed.
- A. E. Morales, E. S. Mora and U. Pal, Rev. Mex. Fis., 2007, 53, 18 CAS.
- Z. Xia, C. Ma, M. S. Molokeev, Q. Liu, K. Rickert and K. R. Poeppelmeier, J. Am. Chem. Soc., 2015, 137, 12494 CrossRef CAS PubMed.
- M. Shang, G. Li, D. Geng, D. Yang, X. Kang, Y. Zhang, H. Lian and J. Lin, J. Phys. Chem. C, 2012, 116, 10222 CAS.
- L. V. Uitert, J. Lumin., 1984, 29, 1 CrossRef.
- P. Dorenbos, Phys. Rev. B: Condens. Matter Mater. Phys., 2000, 62, 15640 CrossRef CAS.
- P. Dorenbos, J. Phys.: Condens. Matter, 2003, 15, 4797 CrossRef CAS.
- N. Hirosaki, R.-J. Xie, K. Kimoto, T. Sekiguchi, Y. Yamamoto, T. Suehiro and M. Mitomo, Appl. Phys. Lett., 2005, 86, 211905 CrossRef.
- G. Blasse, Philips Res. Rep., 1969, 24, 131 CAS.
- D. L. Dexter, J. Chem. Phys., 1953, 21, 836 CrossRef CAS.
- L. G. Van Uitert, J. Electrochem. Soc., 1967, 114, 1048 CrossRef CAS.
- L. Ozawa and P. M. Jaffe, J. Electrochem. Soc., 1971, 118, 1678 CrossRef CAS.
- Y. Zhang, X. Li, K. Li, H. Lian, M. Shang and J. Lin, J. Mater. Chem. C, 2015, 3, 3294 RSC.
- Y. Huang, Opt. Express, 2011, 19, 6311 Search PubMed.
- W. Liu, C. Yeh, C. Huang, C. C. Lin, Y. Chiu, Y. Yeh and R. Liu, J. Mater. Chem., 2011, 21, 3740 RSC.
- H. S. Jang, Y. H. Won and D. Y. Jeon, Appl. Phys. B: Lasers Opt., 2009, 95, 715 CrossRef CAS.
- Y. Wu, D. Wang, T. Chen, C. Lee, K. Chen and H. Kuo, ACS Appl. Mater. Interfaces, 2011, 3, 3195 CAS.
- H. Seong Jang, H. You Kim, Y.-S. Kim, H. Mo Lee and D. Young Jeon, Opt. Express, 2012, 20, 2761 CrossRef PubMed.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra13029a |
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.