Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Tuning of luminescence properties by controlling an aid-sintering additive and composition in Na(Ba/Sr/Ca)PO4:Eu2+ for white LEDs

Qiongyu Bai, Panlai Li*, Zhijun Wang*, Shuchao Xu, Ting Li and Zhiping Yang
Hebei Key Lab of Optic-electronic Information and Materials, College of Physics Science & Technology, Hebei University, Baoding 071002, China. E-mail: li_panlai@126.com; wangzj1998@126.com

Received 27th March 2017 , Accepted 5th April 2017

First published on 21st April 2017


Abstract

A series of Na(Ba/Sr/Ca)PO4:Eu2+ phosphors were prepared via a high-temperature solid-state reaction method. When the phase of NaCaPO4:Eu2+ was pure, the luminescence of Eu2+ was enhanced by doping the sintering-aid additive NaCl (t), and it showed a maxima at t = 0.03. For NaCaPO4:Eu2+ with 0.03NaCl, the XRD patterns, and emission and decay spectra demonstrated that Eu2+ ions occupied two different Ca sites with different coordination (i.e., eight and seven coordination). Therefore, two green emission bands at 510 and 542 nm were observed, and the emission band ranging from 480 to 510 nm had a weaker intensity. To obtain a cyan-broad emission, Ba2+ or Sr2+ were introduced into NaCaPO4:0.01Eu2+. The substitution of small Ca2+ ions by large Ba2+ or Sr2+ ions induced a decreased crystal field splitting of Eu2+ ions, which resulted in various full width at half maximum and a blue shift. The colors varied from green (0.1996, 0.4380) to blue (0.1578, 0.0978) under the same excitation. Overall, the phosphor has promising applications for use in white LEDs.


1 Introduction

Nowadays, white light-emitting diodes (LED) have attracted significant attention as next generation lighting devices due to their advantages of energy saving and long lifetime.1–7 To date, most commercial white LEDs have been fabricated via a combination of blue LED chips and yellow emitting Y3Al5O12:Ce3+ phosphors. However, these white LEDs exhibit a low color rendering index (<75) and a high correlated color temperature (∼7750 K) due to the lack of a red component.8–12 Today, the most effective method to obtain white light is to combine an ultraviolet (UV) LED chip with tricolor phosphors.13–16 However, it usually shows lower luminescence efficiency due to strong reabsorption. Moreover, it shows lower color vividness because the cyan component in the range of 480–520 nm is weaker.17 Therefore, it is necessary to develop a cyan phosphor with a broad emission band that has strong absorption in the near-UV but at the same time has no absorption in the visible region.18

Phosphate hosts with the chemical formula ABPO4 (where A and B are the alkali and alkaline earth cations, respectively) are good candidates as hosts due to their several advantages such as they require low synthesis temperature and have high chemical and physical stability; moreover, they exhibit excellent luminescence properties when Eu2+ is doped as an activator, such as LiSrPO4 and KSrPO4.19–25 In our previous study, a green-emitting NaCaPO4:Eu2+ phosphor was synthesized, which can be used in a white LED.26 After this, green emission was enhanced in NaCaPO4:Eu2+, Tb3+ via energy transfer, according to Wang et al.27 In our study, to avoid a decreased luminescence efficiency due to energy transfer and save energy, the green emission of Eu2+ ions was enhanced by doping NaCl as an aid-sintering additive in NaCaPO4. Hence, a series of NaCaPO4:Eu2+ phosphors was synthesized by a high temperature solid-state method. They showed a green emission band at 515 nm; however, the emission band ranging from 480 to 510 nm was weaker. Therefore, to adjust the emission band and enhance the emission band in the range of 480–510 nm, Ba2+ or Sr2+ ions with larger radii were introduced in NaCaPO4:Eu2+ with 0.03NaCl. The emission band showed an obvious blue shift, and the color ranged from green to blue with the increasing concentrations of Ba2+ or Sr2+ ions; thus, these phosphors can be used in white LEDs.

2 Experimental

2.1 Sample preparation

A series of phosphors with the composition NaCaPO4:xEu2+ with tNaCl (t = 0.01–0.05 mol and x = 0.001–0.10 mol), Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl and Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl were synthesized by a high temperature solid state method using reagent grade CaCO3, BaCO3, SrCO3, NaCl, Na2CO3, NH4H2PO4, and Eu2O3 (99.99%) as raw materials. Starting materials were weighed according to stoichiometric proportion, thoroughly mixed, and ground using an agate mortar and pestle for more than 30 min until they were uniformly distributed. The obtained mixtures were heated at 1200 °C for 2 h in the reduction environment (95% H2 + 5% N2) and then naturally cooled down to room temperature to obtain the final phosphors.

2.2 Characterization

A Bruker D8 X-ray diffractometer was utilized to examine the X-ray powder diffraction (XRD) with Ni-filtered Cu Kα radiation (λ = 0.15418 nm), operating at 40 mA and 40 kV. The step length and diffraction range were 0.05° and 10–80°, respectively. To analyze the chemical composition of the phosphors, scanning electron microscopy (SEM) images and electron-dispersive X-ray (EDX) data were obtained by a Nova NanoSEM 650 with the accelerating voltage of 10 kV. Room temperature photoluminescence spectra of the samples were obtained via a Hitachi F-4600 fluorescence spectrophotometer using a 450 W Xe lamp as the excitation source, with the scanning wavelength ranging from 200 to 700 nm, scanning at 240 nm min−1. The temperature-dependent luminescent properties were obtained using a computer-controlled electric furnace and a self-made heating attachment. Luminescence decay curves and quantum efficiencies of the samples were obtained using a Horiba FL-4600 fluorescence spectrophotometer. Herein, to obtain the luminescence decay curves, a nano-LED (370 nm) was used as the excitation source. The Commission International de I'Eclairage (CIE) coordinates for all samples were measured by a PMS-80 UV-VIS NEAR IR spectra analysis system.

3 Results and discussion

3.1 NaCaPO4:0.01Eu2+ with tNaCl

NaCaPO4:Eu2+ showed a broad emission band at 505 nm, according to our previous study.26 However, it showed weaker luminescence intensity. Hence, we introduced NaCl as an aid-sintering additive to enhance the emission intensity. To verify the influence of NaCl on the emission intensity, the concentration of Eu2+ was randomly fixed at 0.01 mol. Fig. 1 shows the XRD patterns of NaCaPO4:0.01Eu2+ and NaCaPO4:0.01Eu2+ with tNaCl (t = 0.01, 0.02, 0.03, 0.04, and 0.05). It can be seen that the XRD patterns of NaCaPO4:0.01Eu2+ with tNaCl are indexed to NaCaPO4 (JCPDS no. 29-1193) and no impurity phase was detected, indicating that NaCl has no influence on the phase of the phosphor. To further investigate the influence of NaCl on the phase of the phosphor, the crystal structure of NaCaPO4:0.01Eu2+ with 0.03NaCl sample was refined using the General Structure Analysis System (GSAS) program.28 Fig. 2 shows the XRD pattern for Rietveld structure analysis of NaCaPO4:0.01Eu2+ with 0.03NaCl based on the NaCaPO4 phase model, and no impurity phase was detected. The final profile R-factors, Rwp, Rp, and χ2 obtained were 8.62%, 6.19% and 1.539, respectively, which suggested that the pure NaCaPO4:0.01Eu2+ with 0.03NaCl was successfully synthesized. NaCaPO4 was indexed to an orthorhombic crystal system and space group Pnam (a = 0.6797 nm, b = 0.9165 nm, c = 0.5406 nm, and V = 0.3368 nm3). The cell parameters and volume of NaCaPO4:0.01Eu2+ with 0.03NaCl were a = 0.680 nm, b = 0.916 nm, c = 0.542 nm, and V = 0.3376 nm3, which are larger than those of NaCaPO4, indicating that Eu2+ was doped into the host. The crystal structure of NaCaPO4 viewed down the b-axis and the coordination polyhedron was composed of a Na/Ca/P atomic site, as shown in Fig. 2. A total of three independent Na and Ca cation polyhedral sites were available with the coordination number ranging from 6 to 8 in the unit cell of NaCaPO4. Fig. 2 shows the presence of three different crystallographic Na environments: seven-coordinated Na(1) atom with an average Na(1)–O distance of 0.260 nm, Na(2) atom surrounded by eight O atoms at the average distances of 0.264 nm, and a Na(3) atom that was nine-coordinated by nine O atoms at the average Na(3)–O distance of 0.268 nm. There were three crystallographic positions of Ca in the unit cell: seven-coordinated Ca2+ sites (Ca(1)) with the average Ca–O distance of 0.244 nm and two eight-coordinated Ca2+ sites (Ca(2)/Ca(3)) with the average Ca–O distance of 0.246 nm and 0.249 nm, respectively. P atoms with four-coordination modes could also be observed.
image file: c7ra03557h-f1.tif
Fig. 1 XRD patterns of NaCaPO4:0.01Eu2+ and NaCaPO4:0.01Eu2+ with tNaCl (t = 0.01, 0.02, 0.03, 0.04, and 0.05). The standard data for NaCaPO4 (JCPDS no. 29-1193) are shown as a reference.

image file: c7ra03557h-f2.tif
Fig. 2 XRD Rietveld refinement result of NaCaPO4:0.01Eu2+ with 0.03NaCl, and the crystal structure of NaCaPO4.

Fig. 3 shows the emission spectra of NaCaPO4:0.01Eu2+ and NaCaPO4:0.01Eu2+ with tNaCl (t = 0.01, 0.02, 0.03, 0.04, and 0.05) excited at 365 nm. They exhibited a green emission band ranging from 400 to 600 nm. The intensity of NaCaPO4:0.01Eu2+ with tNaCl increased with the increase in Eu2+ ions, reached a maximum at t = 0.03, and then decreased (inset of Fig. 3). Compared to NaCaPO4:0.01Eu2+, the intensity of NaCaPO4:0.01Eu2+ with 0.03NaCl was 10 times enhanced. Therefore, to further investigate the influence of doping concentration on luminescence and improve the luminescent properties, a series of Na(Ba/Sr/Ca)PO4:Eu2+ with 0.03NaCl was synthesized.


image file: c7ra03557h-f3.tif
Fig. 3 Emission spectra of NaCaPO4:0.01Eu2+ and NaCaPO4:0.01Eu2+ with tNaCl (t = 0.01, 0.02, 0.03, 0.04, and 0.05) excited at 365 nm. The inset shows the emission intensity at 515 nm depending on t concentration.

3.2 Na(Ba/Sr/Ca)PO4:Eu2+ with 0.03NaCl

3.2.1 Phase formation and structure. The XRD patterns and the magnified patterns between 22° and 24° of NaCaPO4:xEu2+ with 0.03NaCl with x = 0.001, 0.01, 0.03, and 0.05 are shown in Fig. 4a and b, respectively. It can be seen that all the peaks could be indexed to standard JCPDS no. 29-1193, suggesting that the phosphors are high-purity. According to the effective ionic radii of cations, Eu2+ ions were proposed to occupy the Ca2+ sites in NaCaPO4. The ionic radii for the seven- and eight-coordinated Ca2+ ions were 0.106 and 0.112 nm, respectively; thus, the Ca2+ sites in NaCaPO4 may be occupied by Eu2+ ions with similar ionic radii of 0.121 and 0.125 nm for the same coordination, respectively. The diffraction peaks of NaCaPO4:xEu2+ with 0.03NaCl slightly shifted to a lower angle, and it can be seen from Table 1 that the unit cell expanded with the increasing Eu2+ concentration. Because the radius of Ca2+ is slightly smaller than that of Eu2+, Eu2+ can be easily doped into the host lattice and substituted for the site of Ca2+ ions.
image file: c7ra03557h-f4.tif
Fig. 4 XRD patterns (a) and magnified patterns between 22° and 24° (b) of NaCaPO4:xEu2+ with 0.03NaCl (x = 0.001, 0.01, 0.03, and 0.05). The standard data for NaCaPO4 (JCPDS no. 29-1193) are shown as a reference.
Table 1 Cell parameters and volumes of NaCaPO4:xEu2+ with 0.03NaCl
Formula Cell parameters Volumes
a (nm) b (nm) c (nm) V (nm3)
NaCaPO4:0.001Eu2+ 0.6798 0.9158 0.5412 0.3369
NaCaPO4:0.01Eu2+ 0.6800 0.9160 0.5420 0.3376
NaCaPO4:0.03Eu2+ 0.6810 0.9168 0.5430 0.3390
NaCaPO4:0.05Eu2+ 0.6815 0.9169 0.5464 0.3414


XRD patterns of Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl (a[thin space (1/6-em)]:[thin space (1/6-em)]b = 0[thin space (1/6-em)]:[thin space (1/6-em)]1, 0.1[thin space (1/6-em)]:[thin space (1/6-em)]0.9, 0.3[thin space (1/6-em)]:[thin space (1/6-em)]0.7, 0.5[thin space (1/6-em)]:[thin space (1/6-em)]0.5, 0.7[thin space (1/6-em)]:[thin space (1/6-em)]0.3, and 1[thin space (1/6-em)]:[thin space (1/6-em)]0) phosphors are given in Fig. 5. The standard data for NaCaPO4 (JCPDS no. 29-1193) and NaBaPO4 (JCPDS no. 33-1210) are also shown as a reference in Fig. 5. We can see that Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl can be basically indexed to the NaCaPO4 phase (JCPDS no. 29-1193) when a ≤ 0.5. The phases of Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl agreed well with the standard data for the NaBaPO4 phase (JCPDS no. 33-1210) with the increase of a.


image file: c7ra03557h-f5.tif
Fig. 5 XRD patterns of Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl and standard data (JCPDS no. 29-1193 and JCPDS no. 33-1210).

To further investigate the phase formation depending on the Sr/Ca substitution of Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl phosphors, XRD patterns for the selected samples were obtained and are shown in Fig. 6. All the diffraction peaks of the selected samples agreed well with the standard data for the NaCaPO4 phase (JCPDS no. 29-1193) and NaSrPO4 phase (JCPDS no. 33-1282), indicating that Eu2+ ions were successfully incorporated in the host without noticeably changing the crystal structure.


image file: c7ra03557h-f6.tif
Fig. 6 XRD patterns of Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl and standard data (JCPDS no. 29-1193 and JCPDS no. 33-1282).

Fig. 7a–c shows the representative SEM images and EDX spectra of NaCaPO4:0.01Eu2+ with 0.03NaCl, Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl (a[thin space (1/6-em)]:[thin space (1/6-em)]b = 0.3[thin space (1/6-em)]:[thin space (1/6-em)]0.7) and Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl (c[thin space (1/6-em)]:[thin space (1/6-em)]d = 0.3[thin space (1/6-em)]:[thin space (1/6-em)]0.7). It can be clearly observed that the grains of NaCaPO4:0.01Eu2+ with 0.03NaCl have a spherical shape, and the EDX results indicates that the phosphor has a chemical composition of Na, Ca, P, O, Cl, and Eu. Phosphors showed an irregular shape of blocky particles when Ba2+ and Sr2+ were doped in NaCaPO4:0.01Eu2+ with 0.03NaCl. The EDX spectra of Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl (a[thin space (1/6-em)]:[thin space (1/6-em)]b = 0.3[thin space (1/6-em)]:[thin space (1/6-em)]0.7) and Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl (c[thin space (1/6-em)]:[thin space (1/6-em)]d = 0.3[thin space (1/6-em)]:[thin space (1/6-em)]0.7) confirmed the presence of all elements in the samples. These results suggest that well-crystallized powders were obtained.


image file: c7ra03557h-f7.tif
Fig. 7 SEM images and EDX spectra of NaCaPO4:0.01Eu2+ with 0.03NaCl, Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl (a[thin space (1/6-em)]:[thin space (1/6-em)]b = 0.3[thin space (1/6-em)]:[thin space (1/6-em)]0.7) and Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl (c[thin space (1/6-em)]:[thin space (1/6-em)]d = 0.3[thin space (1/6-em)]:[thin space (1/6-em)]0.7).
3.2.2 Luminescence properties.
3.2.2.1 NaCaPO4:xEu2+ with 0.03NaCl. Fig. 8a–f show the emission spectra of NaCaPO4:xEu2+ with 0.03NaCl (x = 0.001, 0.01, 0.03, 0.05, 0.07, and 0.10) excited at 365 nm, respectively. We fitted the emission spectra with a Gaussian function for NaCaPO4:xEu2+ with 0.03NaCl. The orange dashed lines denote two Gaussian functions, which successfully fitted with the maxima at 510 (Eu2) and 542 (Eu1) nm. Moreover, two different emission bands were obtained due to the different crystal surroundings of Eu2+ sites. From the XRD patterns and structural analysis, it can be inferred that Eu2+ ions occupied Ca2+ sites in NaCaPO4. Hence, three distinct crystallographic sites were formed (i.e., Ca(1), Ca(2), and Ca(3), in which both the crystallographic sites of Ca(2) and Ca(3) were eight-coordinated).
image file: c7ra03557h-f8.tif
Fig. 8 Emission spectra of NaCaPO4:xEu2+ with 0.03NaCl (x = 0.001, 0.01, 0.03, 0.05, 0.07, and 0.10) and the corresponding fitted curves.

Therefore, the emission spectra were fitted by two Gaussian curves, and due to this, the luminescence of the luminescence centers with similar coordinate environments was indistinguishable, according to crystal field theory.29 In addition, to investigate the relationship between the coordinate environment and emission peaks, the emission position of Eu2+ can be simply estimated using the Van Uitert eqn (1):30,31

 
image file: c7ra03557h-t1.tif(1)
where E is the energy location of the lower d-band edge for Eu2+ (cm−1), Q is the energy location for the lower d-band edge of the free ion (Q = 34[thin space (1/6-em)]000 cm−1 for Eu2+), V is the valence of the active cation (V = 2 for Eu2+), n is the coordination number, and r is the radius of the host cation (Ca2+) replaced by the activator Eu2+ ion (Å). The value of Ea was difficult to obtain due to the complexity of the host, but it was constant in the same host. E is proportional to the n and r of the sample; hence, Eu2+ ions with a higher coordination number generally emit at a higher energy, corresponding to a lower wavelength. As a result, 510 (Eu2) and 542 nm (Eu1) emission bands correspond to Ca(2) (Ca(3)) and Ca(1), respectively.

Fig. 9a shows the emission intensities of 510 (Eu2) and 542 nm (Eu1) as a function of Eu2+ contents; it can be observed that their emission intensities gradually increased before reaching the maximum at x = 0.01 for 510 nm emission and x = 0.05 for 542 nm emission, respectively. The emission intensities began to decrease with the increasing concentrations of Eu2+ due to the concentration quenching effect, resulting from the non-radiative energy migration among the activator Eu2+ ions.


image file: c7ra03557h-f9.tif
Fig. 9 (a) Variation of emission intensity for NaCaPO4:xEu2+ with 0.03NaCl as a function of doped Eu2+ concentration. (b) and (c) The relationship of log(I/x) versus log(x). (d) DR, excitation spectra, and fitted curves of NaCaPO4:0.01Eu2+ with 0.03NaCl, monitored at 510 and 542 nm.

Therefore, to further investigate the process of concentration quenching, the type of interaction among the activator Eu2+ ions was calculated by the following equation:32–35

 
I/x = K[1 + β(x)θ/3]−1 (2)
where I and x represent the emission intensity and the concentration of the activator ion, respectively; β and K are the specific constants for a given host crystal and excitation condition, and θ = 3, 6, 8 or 10 denotes the non-radiative energy transfer mechanism of exchange coupling, dipole–dipole, dipole–quadrupole, and quadrupole–quadrupole interactions, respectively. The fitted lines of log(I/x) versus log(x) for two emission bands centered at 510 and 542 nm are shown in Fig. 9b and c, respectively. The slopes of two fitted lines for 510 and 542 nm emission bands were −1.14 and −0.99, respectively. Therefore, the values of θ were 3.42 and 2.97, close to 3, implying that the main concentration quenching mechanism of Eu2+ ions for both 510 and 542 nm emission bands in NaCaPO4 host were the exchange coupling interactions. Fig. 9d shows the DR and excitation spectra of NaCaPO4:0.01Eu2+ with 0.03 NaCl, monitored at 510 and 542 nm. It can be seen that the DR spectrum matched well with the excitation spectra of NaCaPO4:0.01Eu2+ with 0.03NaCl. NaCaPO4:0.01Eu2+ with 0.03NaCl exhibited a broad absorption band from 200 to 450 nm, which was due to the transition of Eu2+ that originated from 4f7 ground state to 4f65d excitation state. For the excitation spectra monitored at 510 and 542 nm, the spectral width at 542 nm was larger than that at 510 nm due to emission bands that originate from different Eu2+ luminescence centers. To further study the reason for different excitation spectra monitored at 510 and 542 nm, their decay curves were measured, as shown in Fig. 10a and b, respectively. A single exponential was used to fit the decay curves, and the effective lifetimes could be defined as:36,37
 
image file: c7ra03557h-t2.tif(3)
where I(t) is the emission intensity at time t and τ is the decay lifetime. The corresponding lifetimes at 510 and 542 nm for NaCaPO4:xEu2+ with 0.03NaCl are shown in Fig. 10a and b, respectively. For example, the lifetimes of NaCaPO4:0.01Eu2+ with 0.03NaCl for 510 and 542 nm emission bands were calculated to be 339.01 ns and 320.22 ns, respectively, which implies that two emission bands were attributed to two different Eu2+ ion luminescence centers.


image file: c7ra03557h-f10.tif
Fig. 10 Decay curves of NaCaPO4:xEu2+ with 0.03NaCl monitored at 510 nm (a) and 542 nm (b).

3.2.2.2 Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl and Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl. Na(BaaCab)PO4:0.01Eu2+ and Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl samples were synthesized using a fixed Eu2+ concentration of 0.01. The normalized emission spectra of Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl and Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl are shown in Fig. 11a and b, respectively, excited at 365 nm. It can be seen that the emission spectra of Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl show an obvious blue shift with an increase in Ba2+ ions. However, when the ratio of Ba and Ca exceeded 0.5[thin space (1/6-em)]:[thin space (1/6-em)]0.5, the full width at half maximum decreased. For Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl, the emission spectra show a broadened full width at half maximum with the increasing Ba2+ ions. Based on the abovementioned discussion, the emission spectra could be deconvoluted into two Gaussian components with the maxima at 510 nm (Eu(2)) and 542 nm (Eu(1)). For Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl and Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl, the fitted spectra of 510 nm and 542 nm are shown in Fig. 12a–d. The intensities of the 510 nm and 542 nm spectra for Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl increased with the increasing Ba2+ concentration up to a[thin space (1/6-em)]:[thin space (1/6-em)]b = 0.2[thin space (1/6-em)]:[thin space (1/6-em)]0.8. Due to the concentration quenching effect, the intensities decreased with further increase in the concentration of Ba2+ ions. For Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl, the intensities of 510 nm and 542 nm increased and reached maxima at c[thin space (1/6-em)]:[thin space (1/6-em)]d = 0.3[thin space (1/6-em)]:[thin space (1/6-em)]0.7 and then decreased with the increment of Sr2+ ions concentration. Fig. 13a–d show the normalized spectra of 510 nm and 542 nm for Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl and Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl excited at 365 nm. For Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl, it can be seen that the spectra of 510 nm and 542 nm shifted to a lower wavelength, as shown in Fig. 13a and b, respectively. Moreover, the full width at half maximum of the 510 nm spectra for Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl decreased when the ratio of Ba to Ca exceeded 0.5[thin space (1/6-em)]:[thin space (1/6-em)]0.5. However, for Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl, the full width at half maximum of 510 nm spectra decreased when the radio of Sr and Ca was greater than 0.7[thin space (1/6-em)]:[thin space (1/6-em)]0.3, and no shift could be observed. The 542 nm spectra showed a blue shift with the increase of Sr2+ ions. Generally, the crystal field splitting (Dq) trends with bond length can be determined by the following equation:38–40
 
Dq = Ze2r4/6R5 (4)
where Dq is the magnitude of the 5d energy level separation, Z represents the anion charge or valence, e is the electron charge, r is the radius of the d wave function, and R is the bond length. The bond length of Ce–O will increase due to the lattice expansion because the radii of Ba2+ and Sr2+ are larger than that of Ca2+. Hence, the crystal field splitting will become weaker, which causes a blue shift and decreases full width at half maximum. The schematic mechanism of the decreased full width at half maximum and blue shift of Eu2+ emission is shown in Fig. 14 when Ba2+ and Sr2+ were substituted for Ca2+ in NaCaPO4:0.01Eu2+ with 0.03NaCl. The blue shift of Na(BaaCab)PO4:0.01Eu2+ with 0.03 NaCl was greater than that of Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl, which is due to that fact that the radius of Ba2+ (rBa = 1.34 nm) is larger than that of Sr2+ (rSr = 1.12 nm). The crystal field splitting of Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl was weaker, and a more obvious blue shift could be observed. For Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl, comparing the decreased full width at half maximum of the 510 nm band and the blue shift of the 542 nm band, the difference may be that Eu2+ had different coordinations (i.e., eight and seven coordination). Therefore, the emission properties could be tuned by controlling the concentration of Ba2+ or Sr2+. Fig. 15a and b show CIE chromaticity coordinates and images of Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl and Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl under a 365 nm UV lamp. It can be seen that the emission color of Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl and Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl could change from green (0.1996, 0.4380) to blue (0.1578, 0.0978) or to cyan (0.2285, 0.2853), respectively. The corresponding images of the samples are shown in Fig. 15.

image file: c7ra03557h-f11.tif
Fig. 11 Normalized emission spectra of Na(BaaCab)PO4:0.01Eu2+ and Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl excited at 365 nm.

image file: c7ra03557h-f12.tif
Fig. 12 (a)–(d) Fitted curves at 510 nm and 542 nm for Na(BaaCab)PO4:0.01Eu2+ and Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl excited at 365 nm.

image file: c7ra03557h-f13.tif
Fig. 13 (a)–(d) Normalized spectra at 510 nm and 542 nm for Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl and Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl excited at 365 nm.

image file: c7ra03557h-f14.tif
Fig. 14 Schematic of the mechanism accounting for the decreased full width at half maximum and blue shift of Eu2+ emission.

image file: c7ra03557h-f15.tif
Fig. 15 CIE chromaticity coordinates and images of Na(BaaCab)PO4:0.01Eu2+ with 0.03NaCl (a) and Na(SrcCad)PO4:0.01Eu2+ with 0.03NaCl (b) under a 365 nm UV lamp.

For LED applications, the thermal stability of the phosphor is one of the important parameters. The temperature-dependent emission spectra of Na(Ba0.7Ca0.3)PO4:0.01Eu2+ with 0.03NaCl and Na(Sr0.8Ca0.2)PO4:0.01Eu2+ with 0.03NaCl under the excitation of 365 nm were investigated, as shown in Fig. 16a and b, respectively. It was observed that the emission intensity continuously decreased with the increasing temperature from room temperature to 200 °C. The emission intensity of Na(Ba0.7Ca0.3)PO4:0.01Eu2+ with 0.03NaCl and Na(Sr0.8Ca0.2)PO4:0.01Eu2+ with 0.03NaCl decreased to 90% and 76% of the initial emission intensity, respectively, corresponding to 100 °C. It can be seen that the thermal quenching of Na(Sr0.8Ca0.2)PO4:0.01Eu2+ with 0.03NaCl was inferior to that of Na(Ba0.7Ca0.3)PO4:0.01Eu2+ with 0.03NaCl. The reason is that in Na(Ba/Sr/Ca)PO4:Eu2+ with NaCl, with the replacement of Eu2+–Eu2+ neighbors by Eu2+–Ba2+/Sr2+ pairs, the emission intensity will decrease with an increment in temperature due to the lower nonradiative decay rate from the lowest excited state, according to Peng et al.41 The bond length will become shorter when Ba2+ is co-doped in NaCaPO4:Eu2+ with NaCl with respect to Sr2+ co-doping because the radius of Ba2+ is larger than that of Sr2+. The thermal excitation from Eu2+–Ba2+/Sr2+ may occur differently with an increase in temperature. Hence, co-doping Sr will lead to worse thermal quenching compared to co-doping Ba due to covalent effects. The slightly decreased intensity indicated that the Na(Ba/Sr/Ca)PO4:0.01Eu2+ with 0.03NaCl phosphor could be applied to a white LED. To investigate the relationship of luminescence with temperature and to calculate the activation energy from thermal quenching, the activation energy (Ea) can be expressed by the following formula:42

 
I = I0/[1 + c[thin space (1/6-em)]exp(−Ea/kT)] (5)
where I and I0 are the luminescence intensities of the phosphor at the testing temperature and room temperature, respectively, Ea represents the thermal quenching activation energy of the phosphor, c is the rate constant for thermally activated escape, and k is the Boltzmann constant (8.629 × 10−5 eV K−1). The inset of Fig. 16a and b show the plots of ln[(I0/I) − 1] versus 1/T for Na(Ba0.7Ca0.3)PO4:0.01Eu2+ with 0.03NaCl and Na(Sr0.8Ca0.2)PO4:0.01Eu2+ with 0.03NaCl, respectively. The calculated Ea were 0.1506 and 0.2399 eV for Na(Ba0.7Ca0.3)PO4:0.01Eu2+ with 0.03NaCl and Na(Sr0.8Ca0.2)PO4:0.01Eu2+ with 0.03NaCl, which indicate that the phosphors show relatively good thermal stability and can be used in an LED.


image file: c7ra03557h-f16.tif
Fig. 16 Temperature-dependent emission spectra of Na(Ba0.7Ca0.3)PO4:0.01Eu2+ with 0.03NaCl (a) and Na(Sr0.8Ca0.2)PO4:0.01Eu2+ with 0.03NaCl (b) excited at 365 nm. The inset shows the activation energy of Na(Ba0.7Ca0.3)PO4:0.01Eu2+ with 0.03NaCl (a) and Na(Sr0.8Ca0.2)PO4:0.01Eu2+ with 0.03NaCl (b).

To verify the actual application of the phosphor, we chose the sample with the broadest full width at half maximum to fabricate a white LED. The devices were combined with a 380 nm UV chip (CaSr)AlSiN3:Eu2+ and Na(Ba0.7Ca0.3)PO4:0.01Eu2+ with 0.03NaCl phosphors, Fig. 17a. CIE coordinates were (0.3750, 0.3634), and the image shows an excellent white light, as shown in Fig. 17b. Fig. 17c shows the electroluminescence spectrum of the white LED. Tt can be seen that the spectrum shows a stronger emission in the range of 480–510 nm. The quantum efficiency of Na(Ba0.7Ca0.3)PO4:0.01Eu2+ with 0.03NaCl was 45.06%. Therefore, Na(Ba/Sr/Ca)PO4:0.01Eu2+ with 0.03NaCl has potential applications in white LEDs.


image file: c7ra03557h-f17.tif
Fig. 17 (a) Diagram of white LED package. (b) CIE chromaticity coordinates and photos of white LED, which is fabricated by Na(Ba0.7Ca0.3)PO4:0.01Eu2+ with 0.03 NaCl and (CaSr)AlSiN3:Eu2+ phosphors. (c) Electroluminescence spectra of white LED composed of a 380 nm UV chip.

4 Conclusions

In summary, a series of Na(Ba/Sr/Ca)PO4:0.01Eu2+ with 0.03NaCl phosphors were synthesized via a high-temperature solid-state reaction method. On the one hand, the luminescence intensity of NaCaPO4:Eu2+ was enhanced using NaCl as the aid-sintering additives, and the phases were indexed to pure NaCaPO4. NaCaPO4:Eu2+ exhibited the strongest emission when the concentration of NaCl was 0.03. NaCaPO4:Eu2+ with 0.03NaCl exhibited two green emission bands at 510 and 542 nm, which were ascribed to two different Ca sites with different coordination (i.e., eight and seven coordination). On the other hand, by gradually introducing Ba2+ or Sr2+ into NaCaPO4:0.01Eu2+ with 0.03NaCl, 510 and 542 nm emission bands showed various full width at half maximum and a blue shift caused by the crystal field splitting. The intensity of cyan region from 480 to 510 nm increased by controlling the Ba or Sr concentration. Compared to a Sr2+-doped phosphor, Na(BaCa)PO4:0.01Eu2+ with 0.03NaCl exhibited a more obvious blue shift because the radius of Ba2+ is larger than that of Sr2+. Therefore, luminescence of Na(Ba/Sr/Ca)PO4:0.01Eu2+ with 0.03NaCl could be tuned from green (0.1996, 0.4380) to blue (0.1578, 0.0978) under the same excitation by introducing Sr2+ or Ba2+ ions. Moreover, these phosphors can be used in UV/n-UV-pumped white LEDs.

Acknowledgements

The work was supported by the National Natural Science Foundation of China (No. 51672066), the Funds for Distinguished Young Scientists of Hebei Province, China (No. A2015201129), the personnel training project of Hebei Province, China (No. A2016002013), and the Post-graduate's Innovation Fund Project of Hebei University (No. X2016063, X2016064).

References

  1. H. Ji, L. Wang, M. S. Molokeev, N. Hirosaki, R. Xie, Z. Huang, Z. Xia, O. M. Kate, L. Liu and V. V. Atuchin, J. Mater. Chem. C, 2016, 4, 6855–6863 RSC.
  2. F. Kang, H. Zhang, L. Wondraczek, X. Yang, Y. Zhang, D. Y. Lei and M. Peng, Chem. Mater., 2016, 28, 2692–2703 CrossRef CAS.
  3. S.-P. Lee, S.-D. Liu, T.-S. Chan and T.-M. Chen, ACS Appl. Mater. Interfaces, 2016, 8, 9218–9223 CAS.
  4. X. Y. Liu, H. Guo, Y. Liu, S. Ye, M. Y. Peng and Q. Y. Zhang, J. Mater. Chem. C, 2016, 4, 2506–2512 RSC.
  5. H. C. Yoon, H. Kang, S. Lee, J. H. Oh, H. Yang and Y. R. Do, ACS Appl. Mater. Interfaces, 2016, 8, 18189–18200 CAS.
  6. L. Wang, B. K. Moon, S. H. Park, J. H. Kim, J. Shi, K. H. Kim and J. H. Jeong, RSC Adv., 2016, 6, 79317–79324 RSC.
  7. F. Pan, M. Zhou, J. Zhang, X. Zhang, J. Wang, L. Huang, X. Kuang and M. Wu, J. Mater. Chem. C, 2016, 4, 5671–5678 RSC.
  8. A. A. Setlur, W. J. Heward, Y. Gao, A. M. Srivastava, R. G. Chandran and M. V. Shankar, Chem. Mater., 2006, 18, 3314–3322 CrossRef CAS.
  9. X. Piao, T. Horikawa, H. Hanzawa and K. I. Machida, Appl. Phys. Lett., 2006, 88, 161908 CrossRef.
  10. X. Gong, J. Huang, Y. Chen, Y. Lin, Z. Luo and Y. Huang, Inorg. Chem., 2014, 53, 6607–6614 CrossRef CAS PubMed.
  11. K. Li, Y. Zhang, X. Li, M. Shang, H. Lian and J. Lin, Phys. Chem. Chem. Phys., 2015, 17, 4283–4292 RSC.
  12. H. S. Jang, W. B. Im, D. C. Lee, D. Y. Jeon and S. S. Kim, J. Lumin., 2007, 126, 371–377 CrossRef CAS.
  13. Y. Zhang, X. Li, K. Li, H. Lian, M. Shang and J. Lin, ACS Appl. Mater. Interfaces, 2015, 7, 2715–2725 CAS.
  14. D. Wen, J. Feng, J. Li, J. Shi, M. Wu and Q. Su, J. Mater. Chem. C, 2015, 3, 2107–2114 RSC.
  15. L. Jiang, R. Pang, D. Li, W. Sun, Y. Jia, H. Li, J. Fu, C. Li and S. Zhang, Dalton Trans., 2015, 44, 17241–17250 RSC.
  16. C. C. Lin, Z. R. Xiao, G.-Y. Guo, T.-S. Chan and R.-S. Liu, J. Am. Chem. Soc., 2010, 132, 3020–3028 CrossRef CAS PubMed.
  17. M.-H. Fang, C. Ni, X. Zhang, Y.-T. Tsai, S. Mahlik, A. Lazarowska, M. Grinberg, H.-S. Sheu, J.-F. Lee, B.-M. Cheng and R.-S. Liu, ACS Appl. Mater. Interfaces, 2016, 8, 30677–30682 CAS.
  18. F. Kang, M. Peng, D. Y. Lei and Q. Zhang, Chem. Mater., 2016, 28, 7807–7815 CrossRef CAS.
  19. W. L. Wanmaker and H. L. Spier, J. Electrochem. Soc., 1962, 109, 109–114 CrossRef CAS.
  20. M. S. Waite, J. Electrochem. Soc., 1974, 121, 1122–1123 CrossRef CAS.
  21. W. L. Wanmaker, A. Bril and J. W. ter Vrugt, Appl. Phys. Lett., 1966, 8, 260–261 CrossRef CAS.
  22. C. S. Liang, H. Eckert, T. E. Gier and G. D. Stucky, Chem. Mater., 1993, 5, 597–603 CrossRef CAS.
  23. Z. C. Wu, J. X. Shi, M. L. Gong, J. Wang and Q. Su, Mater. Chem. Phys., 2007, 103, 415–418 CrossRef CAS.
  24. Z. C. Wu, J. X. Shi, J. Wang, M. L. Gong and Q. Su, J. Solid State Chem., 2006, 179, 2356–2360 CrossRef CAS.
  25. Y. S. Tang, S. F. Hu, C. C. Lin, N. C. BagKar and R. S. Liu, Appl. Phys. Lett., 2007, 90, 151108 CrossRef.
  26. Z. P. Yang, G. W. Yang, S. L. Wang, J. Tian, X. N. Li, Q. L. Guo and G. S. Fu, Mater. Lett., 2008, 62, 1884–1886 CrossRef CAS.
  27. Y. Wang, M. G. Brik, P. Dorenbos, Y. Huang, Y. Tao and H. Liang, J. Phys. Chem. C, 2014, 118, 7002–7009 CAS.
  28. C. Larson and R. B. Von Dreele, Los Alamos National Laboratory Report LAUR, Los Alamos National Laboratory, Los Alamos, 2000, vol. 86, pp. 748–789 Search PubMed.
  29. W. Tang and Z. Zhang, J. Mater. Chem. C, 2015, 3, 5339–5346 RSC.
  30. X. J. Wang, L. Wang, T. Takeda, S. Funahashi, T. Suehiro, N. Hirosaki and R. Xie, Chem. Mater., 2015, 27, 7689–7697 CrossRef CAS.
  31. L. G. Van Uitert, J. Lumin., 1984, 29, 1–9 CrossRef CAS.
  32. L. Ozawa and P. M. Jaffe, J. Electrochem. Soc., 1971, 118, 1678–1679 CrossRef CAS.
  33. L. G. Van Uitert, J. Electrochem. Soc., 1967, 114, 1048–1053 CrossRef CAS.
  34. C. Huang, Y. Chiu, Y. Yeh, T. Chan and T. Chen, ACS Appl. Mater. Interfaces, 2012, 4, 6661–6668 CAS.
  35. B. Wang, H. Lin, J. Xu, H. Chen and Y. Wang, ACS Appl. Mater. Interfaces, 2014, 6, 22905–22913 CAS.
  36. Y. Jia, H. Qiao, Y. Zheng, N. Guo and H. You, Phys. Chem. Chem. Phys., 2012, 14, 3537–3542 RSC.
  37. Y. Shimomura, T. Honma, M. Shigeiwa, T. Akai, K. Okamoto and N. Kijima, J. Electrochem. Soc., 2007, 154, J35–J38 CrossRef CAS.
  38. C. K. Jorgensen, Modern Aspects of Ligand-Field Theory, Amsterdam, North-Holland, 1971 Search PubMed.
  39. M. Shang, J. Fan, H. Lian, Y. Zhang, D. Geng and J. Lin, Inorg. Chem., 2014, 53, 7748–7755 CrossRef CAS PubMed.
  40. Y. Xia, J. Chen, Y. Liu, M. S. Molokeev, M. Guan, Z. Huang and M. Fang, Dalton Trans., 2016, 45, 1007–1015 RSC.
  41. M. Peng, X. Yin, P. A. Tanner, M. G. Brik and P. Li, Chem. Mater., 2015, 27, 2938–2945 CrossRef CAS.
  42. S. Bhushan and M. V. Chukichev, J. Mater. Sci. Lett., 1988, 74, 319–321 CrossRef.

This journal is © The Royal Society of Chemistry 2017
Click here to see how this site uses Cookies. View our privacy policy here.