Substitution priority of Eu2+ in multi-cation compound Sr0.8Ca0.2Al2Si2O8 and energy transfer
Yang Zhi-Ping, Li Zhen-Ling, Wang Zhi-Jun, Li Pan-Lai, Tian Miao-Miao, Cheng Jin-Ge, Wang Chao
College of Physics Science & Technology, Hebei Key Laboratory of Optic-Electronic Information and Materials, Hebei University, Baoding 071002, China

 

† Corresponding author. E-mail: wangzj1998@126.com li_panlai@126.com

Abstract

A blue phosphor was obtained by doping Eu2+ into a multi-cation host Sr0.8Ca0.2Al2Si2O8 through high temperature solid state reaction. The emission spectra show a continuous red-shift behavior from 413 nm to 435 nm with Eu2+ concentration increasing. The substitution priority of Eu2+ in Sr0.8Ca0.2Al2Si2O8 was investigated via x-ray diffraction (XRD) and temperature properties in detail: the Ca2+ ions are preferentially substituted by Eu2+at lower doping, and with the Eu2+ concentration increasing, the probability of substitution for Sr2+ is greater than that of replacing Ca2+. Accordingly, we propose the underlying method of thermal property to determine the substitution of Eu2+ in the multi-cation hosts. Moreover, the abnormal increase of emission intensity with increasing temperature was studied by the thermoluminescence spectra. The energy transfer mechanism between the Eu2+ ions occupying different cation sites was studied by the lifetime decay curves. A series of warm white light emitting diodes were successfully fabricated using the blue phosphors Sr0.8Ca0.2Al2Si2O8: Eu2+ with commercial red phosphor (Ca Sr)SiAlN3: Eu2+ and green phosphor (Y Lu)3Al5O12: Ce3+, and the luminescent efficiency can reach 45 lm/W.

1. Introduction

Recently, rare earth ion Eu2+ has been commonly applied to some kinds of phosphors due to the emission of Eu2+ ion being sensitive to the crystal field and covalence since the 5d states of Eu2+ are the outer orbital.[13] Moreover, the energy of the excited state 4f65d1 of the Eu2+ ion is lower than that of the lowest excited state 6P in the 4f7 electron configuration. Therefore, the Eu2+ ions in most compounds present a broad band emission varying from ultraviolet to red light as a result of the 4f65d1–4f7 transition.[46] The local environment of Eu2+ ion directly dominates the luminescence properties of the materials. So we can obtain the desired emission characteristics by adjusting the Eu2+ environment. Recently, more and more researchers are devoted to adjusting the luminescent properties of Eu2+ by changing the species or ratio of cations or anion polyhedrons.[711] However, for some multi-cation host, there are two or more cation sites, and Eu2+ in different sites will produce different luminescence properties.[12] Hence, understanding which site is preferentially replaced by the Eu2+ ion is of great significance for adjusting the luminescent properties of the multi-cation host phosphors. In some cases, the substitution of Eu2+ can be qualitatively judged by the movement of 2θ angle in x-ray diffraction (XRD) and the shift of emission spectra of phosphors.[1319] However, for some complicated multi-cation hosts, other ways need to be introduced for further verification. In this work, a new way, thermal characteristic, is proposed to determine the substitution of Eu2+ in the multi-cation host of Sr0.8Ca0.2Al2Si2O8 (SCAS). Moreover, the structure morphology, luminescent properties, decay lifetimes, CIE chromaticity coordinates, and quantum yields of Sr0.8Ca0.2Al2Si2O8: Eu2+ are investigated in detail, and a series of warm white LEDs with high CRI (86.6–89.6) and low CCT (3000–4000 K) are fabricated by employing blue phosphor Sr0.8Ca0.2Al2Si2O8: Eu2+, commercial green phosphor (Y Lu)3Al5O12: Ce3+, and red phosphor (Ca Sr)SiAlN3: Eu2+ on an ultraviolet chip.

2. Experimental details
2.1. Materials and synthesis

A series of Sr0.8 −xCa0.2Al2Si2O8: xEu2+ (x = 0, 0.1%, 1%, 3%, 5%, 7%, 8%, and 10%) phosphors were synthesized in this work through the solid-state reaction with CaCO3 (A.R.), Eu2O3 (99.99%), Al2O3 (A.R.), SrCO3 (A.R.), and SiO2 (A.R.) as raw materials. Stoichiometric mixtures of the raw materials were homogeneously mixed and ground, then placed in the alumina crucibles with covers and sintered at 1400 °C for 4 h under reductive atmosphere (95% H2+5% N2) in a high temperature furnace. After firing, the samples were cooled to room temperature, and ground into powder for subsequent use.

2.2. Structure characterizations

Phase purity of the samples was checked using a D8 x-ray diffractometer (XRD, 40 kV, 40 mA) with Cu radiation (λ = 1.5418 Å). The XRD data were collected in a 2θ range from 10° to 80°, with the continuous scan mode at the speed of 0.1 s per step with a step size of 0.05°. Furthermore, the XRD patterns submitted for Rietveld refinement were acquired at a step size of 0.03° with a counting time of 0.5 s per step. Scanning electron micrograph (SEM) images were obtained using a Nova Nano SEM 650 instrument.

2.3. Optical measurements

The excitation and emission spectra were measured using a Hitachi F4600 fluorescence spectrophotometer and a steady state fluorescence spectrometer equipped with a continuous xenon lamp (450 W), and both excitation and emission spectra were set up to be 0.2 nm with the width of the monochromator slits adjusted as 2.5 nm. The diffuse reflection spectra were measured with a Hitachi U4100 UV-VIS-NIR spectroscope, with the scanning wavelength from 200 nm to 800 nm at 240 nm/min. The thermoluminescence (TL) spectra of the samples were measured using an FJ-427Al TL dosimeter with a fixed heating rate of 1 °C/s within the range 25–300 °C. The luminescence decay curves of the samples were measured with a Horiba FL-1057 fluorescence spectrophotometer using an external LED (370 nm) as the excitation source. The thermally stable property was tested using a heating apparatus (TAP-02) in combination with a PL equipment. The color coordinates (X, Y) of the samples were obtained by a PMS-80 color analyzer at room temperature. The internal quantum yields (IQYs) of the phosphors were measured by a Horiba FL-1057.

3. Results and discussion
3.1. Phase formation and structural characteristics

The XRD patterns of SCAS: xEu2+ (x = 0, 0.1%, 1%, 3%, 5%, 7%, 8%, and 10%) samples are displayed in Fig. 1(a). It can be seen that all of the diffraction peaks can be indexed to the standard data of SrAl2Si2O8 (JCPDS 38-1454), indicating that the doped ions are completely dissolved into the SrAl2Si2O8 host without inducing significant changes of the crystal structure. The main magnified XRD patterns in the range of 35°–36.4° are shown in Fig. 1(b). It can be found that the main diffraction peaks of SCAS: xEu2+ first shift slightly to smaller angles before x = 3%, but shift to higher angles after x = 3%. On the basis of the effective ionic radii of the cations with different coordination numbers, the Eu2+ ions are expected to occupy smaller Ca2+ sites (n = 6, 7) and Sr2+ sites (n= 6) when , but occupy bigger Sr2+ sites (n = 7) when , because the effective ionic radii for six-coordinated Ca2+, Sr2+, and Eu2+ are 1 Å, 1.13 Å, and 1.17 Å, and seven-coordinated Ca2+ Sr2+, and Eu2+ are 1.06 Å, 1.13 Å, and 1.20 Å, respectively.[2022]

Fig. 1. (color online) (a) XRD patterns and (b) the detailed XRD patterns ranging from 35° to 36.4° of as-prepared samples SCAS: xEu2+ (x = 0, 0.1%, 1%, 3%, 5%, 7%, 8%, and 10%). The standard data for SrAl2Si2O8 (JCPDS 38-1454) is shown as a reference.

The Rietveld structural refinements of the XRD data profiles for SCAS: 3%Eu2+ are presented in Fig. 2. The results show that Sr0.8Ca0.2Al2Si2O8: 3%Eu2 + has a triclinic unit cell with cell parameters a = 8.35 Å, b = 12.976 Å, c = 14.227 Å, and V = 1392.16 Å3. The refinements finally converge to the goodness of fit coefficient with , , and , which indicates that the crystal structures of these phosphors match well with the starting model (Sr0.8Ca0.2Al2Si2O8) after the refinement.

Fig. 2. (color online) Rietveld refinements of powder XRD profiles of the representative SCAS: 3%Eu2+ phosphor.

Figure 3 presents the crystal structure of Sr0.8Ca0.2Al2Si2O8. The structure consists of a layered framework made from Si–O as well as Al–O tetrahedron.[23] The Si–O and Al–O tetrahedrons link to each other by sharing corners, forming an unusual eight-member ring. The Ca2+ and Sr2+ ions surrounded with oxygen neighbors in sheets compensate for the charge difference between Al3+ and Si4+.[24,25] The Sr0.8Ca0.2Al2Si2O8 host contains two kinds of cation sites denoted as Ca1/Sr1 and Ca2/Sr2. Ca1/Sr1 are coordinated by 7 oxygen atoms, and Ca2/Sr2 are coordinated by 6 oxygen atoms. Figure 4 show the SEM images of microstructures and shapes for the representative SCAS host and SCAS: 3%Eu2+ samples. It is observed that the particles have irregular blocky shape with a size of about .

Fig. 3. The crystal structure of Sr0.8Ca0.2Al2Si2O8.
Fig. 4. SEM images of (a) SCAS host and (b) SCAS: 3%Eu2+.
3.2. Photoluminescence properties

The diffuse reflectance spectra (DRS) of SCAS and Eu2+ doped SCAS phosphors are displayed in Fig. 5. It is obvious that the profiles of Eu2+-doped SCAS are very different from that of the SCAS host, which illustrates that the Eu2+ ions can be an effective activator and exhibit their characteristic excitation and emission spectral properties in this host. As can be seen in the DRS of the SCAS host, the absorption in the ultraviolet can be attributed to the valenceto-conduction band. The band gap of the SCAS host is shown in the inset of Fig. 5. The band gap of the SCAS host can be estimated using the following formula:[26] where h is Planck’s constant, υ is the frequency, is the band gap, A is the proportional constant, and exponent n is related to the nature of electron transition: denotes the direct allowed transition and n = 0.5 for an indirect allowed transition. is a Kubelka–Munk function defined as[27] where K represents the absorption coefficient, and S and R are the absorption coefficient and the scattering coefficient, respectively.

Fig. 5. (color online) UV–visible DRS of SCAS host. Inset shows the plot of versus photoenergy for the SCAS host.

From the reflection spectrum, the band gap is estimated to be about 3.683 eV. As the Eu2+ ions are introduced into the SCAS host, the DRS shows a broad deep valley from 280 nm to 650 nm, which is attributed to the 4f–5d absorption of the Eu2+ ions.[2830] In addition, it can also be found that the absorption extent increases continuously with Eu2+ concentration increasing, which further demonstrates the absorption bands deriving from the Eu2+ ions.

The excitation spectra of SCAS: xEu2+ ( ) phosphors are shown in Fig. 6. It can be seen that the intensity increases firstly and then decreases with Eu2+ concentration increasing. The quenching concentration is 3%. Under monitoring at 417 nm, an intense and broad excitation band (200 nm to 400 nm) with a peak at 353 nm can be observed, which is ascribed to the transition from the 4f75d0 ground state to the 4f65d1 excited state of the Eu2+ ion. The broad excitation spectra with high absorption ability from 350 nm to 380 nm can match well with n-UV LED chips.

Fig. 6. (color online) Excitation (PLE) spectra of SCAS: xEu2+ ( ) phosphors ( ).

Figure 7 shows the emission spectra of SCAS: xEu2+ ( ) phosphors under the excitation of 353 nm. It is clear to see that the emission spectra present an asymmetric broad band (400–650 nm), which indicates that the Eu2+ ions have more than one luminescent center in the lattice. In addition, photos of SCAS: xEu2+ ( ) under 365 nm excitation are also shown in Fig. 7. As we can see, the emission changes from deep blue to light blue, which indicates that a small red-shift appears with the increase of Eu2+ concentration. To clearly show the red-shift behavior, the normalized emission spectra of SCAS: xEu2+ ( ) are shown in Fig. 8.

Fig. 7. (color online) Emission (PL) spectra of SCAS: xEu2+ ( ) phosphors ( ). Insets show the relative intensity of peak and digital images ( ) of SCAS: Eu2+ phosphor with various concentrations of Eu2+.
Fig. 8. (color online) Normalized emission spectra of SCAS: xEu2+ phosphors ( ), and the inset shows the corresponding peak wavelength.

It can be seen that the peak of the emission band shifts to longer wavelength from 413 nm to 435 nm together with increasing full width at half maximum (FWHM) as the Eu2+ concentration increases. In order to find out the origin of the red shift and increasing FWHM, we carry out the following study. Firstly, as is mentioned in the inset of Fig. 3, there are four cation sites in the SCAS host (Ca1/Sr1 coordinated by seven oxygen atoms and Ca2/Sr2 coordinated by six oxygen atoms). To study the occupation of Eu2+ in the four cations, we conduct Gauss fitting on the emission spectra, as shown in Fig. 9.

Fig. 9. (color online) Emission spectra of SCAS: xEu2+ phosphors excited at 353 nm with Gaussian peaks.

The emission spectra are deconvoluted into four Gaussian components peaking at , , , and , respectively. In order to clarify the origin of the four emission bands, the well-known experiential equation given by Van Uitert is used as follows:[31] where E represents the energy position of the d-band edge for the rare-earth ion (cm−1), Q is the energy position for the lower d-band edge for the free ion and is 34000 cm−1 for Eu2+, V is the valence of the activator (for Eu2+ ion, V = 2), n is the number of anions in the immediate shell around the Eu2+ ion, r is the radius of the host cation replaced by Eu2+ ion, and is the electron affinity of the atoms that form anions relying on the anion type. Although is very complex and difficult to determine, it is a fixed value in a specific host. Therefore the value of E depends on the parameters n and r directly, and the value of E increases with the increase of . In order to make clear which sites the four emission bands come from, the values of are presented in Table 1. It is clear to see that the values of for Sr1 and Ca1 are 8.47 and 7.42, respectively. The values of for Sr2 and Ca2 are 6.78 and 6, respectively. Since E is proportional to , it can be deduced that the four Gaussian components from right to left belong to Eu2+ centers at Sr1, Ca1, Sr2, and Ca2, respectively. In addition, it is well known that the shift of the spectrum is closely related to the splitting of the crystal field. So we study the crystal field splitting caused by the substitution of Eu2+ for different cations. According to the reports of Robertson et al.[32] and Jang et al,[33] crystal field splitting can be determined by the following equation:[34,35] where is a measure of the energy level separation, Z is the anion charge, e is the electron charge, r is the radius of the d wavefunction, and R is the bond length. When the Eu2+ ion substitutes a smaller ion, the distance between Eu2+ and O2− becomes short. Since the crystal field splitting is proportional to , this shorter Eu2+–O2− distance also leads to the enhancement of the crystal field strength surrounding the Eu2+ ion and further results in a larger crystal field splitting of the Eu2+ 5d energy levels, which makes the lowest 5d state of Eu2+ closer to its ground state and finally gives a red shift of the emission peak of Eu2+. Oppositely, when the Eu2+ ion replaces a bigger ion, the emission spectrum shows a blue shift.[36] However, in this work, although the host contains both Ca2+ ( and Sr2+ ( , the obtained emission spectra just show the red shift, which may attribute to the combination effect of the following two cases.

Table 1.

Values of at different sites.

.
3.3. Substitution priority of Eu2+ ions

In order to find out the cause of the spectrum red shift, figuring out Eu2+ substitution priority is a primary work. As is known to us, the thermal quenching behavior is associated with the crystal structure of the phosphor. So the substitution of Eu2+ ions for different cations will result in different thermal quenching behaviors. Based on this, the emission spectra at different temperatures (25–300 °C) excited by 353 nm for SCAS: xEu2+ (x = 0, 0.1%, 3%, 5%, 8%, and 10%) are measured and plotted in Fig. 10. At the same time, as a comparison, the temperature properties of Eu2+ doped single-cation host phosphors SrAl2Si2O8: 3%Eu2+ and CaAl2Si2O8: 3%Eu2+ are also measured under the same conditions and presented at the top of Fig. 10. It can be seen that the emission intensity of SrAl2Si2O8: 3%Eu2+ decreases much faster than that of CaAl2Si2O8: 3%Eu2+ with temperature increasing, which indicates that Eu2+ replacing Ca2+ is better for the thermal stability of the phosphor. The temperature quenching behaviors of SCAS: xEu2+ phosphors show an interesting phenomenon. For x = 0.1%–3%, the emission intensities decline weakly with increasing temperature, which are close to the behavior of the CaAl2Si2O8: 3%Eu2+. Meanwhile, the emission intensity for x = 10% declines relatively quickly with the increase of temperature, which is similar to the thermal quenching behavior of SrAl2Si2O8: 3%Eu2+. For the intermediate contents, x = 5% and 8%, the samples present the transition state of the two. The results indicate that the Ca2+ ions are preferentially substituted by Eu2+ at lower doping. With the Eu2+ concentration increasing, the probability of substitution for Sr2+ is greater than that of replacing Ca2+, which is basically consistent with the above analysis of the XRD. In addition, it can be observed that the emission intensities show an abnormal increase with the increasing temperature for x = 3%, 5%, and 8% in Fig. 10. To investigate the cause of the phenomenon, the thermoluminescence spectra of SCAS: xEu2+ (x = 0.1%, 3%, 5%, 8% and 10%) are measured and shown in Fig. 11.

Fig. 10. (color online) Emission spectra of MAl2Si2O8: 3% Eu2+ (M = Sr and Ca) and SCAS: xEu2+ (x = 0.1%, 3%, 5%, 8%, 10%) at different temperatures.
Fig. 11. (color online) TL glow curves of SCAS: x Eu2+ (x = 0.1%, 3%, 5%, 8%, and 10%).

It is obvious that there is defect luminescence in all the samples, and moreover the TL intensities decrease with Eu2+ concentration increasing. So, for x = 10%, the TL intensity is so weak that it makes no obvious effect on the photoluminescence intensity. Additionally, as shown in Fig. 10, an obvious blue shift of the emission peak is observed with increasing temperature for SCAS: xEu2+, which can be explained by the thermally active phonon-assisted tunneling from lower energy sublevel to high-energy sublevel of the excited state of Eu2+.[3739] Two emissions (Eu(I) (occupying Ca(1) and Sr(1)) and Eu (II) (occupying Ca(2) and Sr(2)) arise from different minima on the potential energy surface of the relaxed excited state as given in Fig. 12. Since Eu(I) and Eu(II) ions are in different crystal fields, the 4f65d1 excited states of the two are located at different energy levels. When SCAS: Eu2+ phosphor is excited by the UV light, the electrons release different energies from different excited states to ground state. At low temperature, the thermal transfer ( ) over barrier can be overcome and the low-energy emission Eu (2) is dominant. At high temperature, the thermal back-transfer ( ) over the barrier is possible, and consequently the high energy emission Eu (I) is dominant. Thus, the blue shift behavior is observed with increasing temperature.

Fig. 12. (color online) Schematic illustration of a configurational coordinate diagram of the ground state (Eu2+: 4f75d0), two split excited states (Eu(I) and Eu(II), and showing the energy barrier of the transition Eu(I) state to Eu(II), and showing the energy barrier of the transition Eu(II) state to Eu(I).
3.4. Energy transfer between Eu2+ ions

Through the discussion of temperature properties, the substitution process is that the Ca2+ ions are preferentially substituted by Eu2+at lower doping, and with the Eu2+ concentration increasing, the probability of substitution for Sr2+ is greater than that of replacing Ca2+. Considering the red-shift of the emission spectra, it is far more likely that there is energy transfer from Eu2+ ions at higher energy to Eu2+ ions at lower energy. In order to prove this speculation, a series of fluorescent decay curves are measured by monitoring two different luminescence centers at Eu(I) and Eu (II) for SCAS: xEu2+ (x = 0.1%, 1%, 3%, 5%, 7%, 8%, and 10%) phosphors. The decay curves are obtained by monitoring the emission peaks at around 405 nm, 419 nm, 439 nm, and 451 nm, corresponding to the different emission centers, and plotted in Fig. 13.

Fig. 13. (color online) Decay curves of different luminescence centers (Eu(I) of Ca1 (a), Sr1 (b) and Eu (II) of Ca(2) (c), Sr (2) (d)).

The decay behaviors can be fitted with a second-order exponential decay mode because of the two luminescence centers (Eu(I) and Eu(II) as follows:[40] where I refers to the luminous intensity, A1 and A2 are constants, t is time, and τ1 and τ2 are the rapid and slow lifetimes for the exponential components, respectively. According to Eq. (6), A1, A2, τ1, and τ2 can be calculated based on the fitting of the decay curves. Therefore the effective lifetime constant τ can be calculated by[41]

The variations of τ for Eu(I) and Eu (II) with increasing Eu2+ concentration are depicted in Fig. 14. For x = 0.1%, is close to and is close to . As we discussed above, Eu2+ preferentially replaces the Ca2+ ion at low Eu2+ concentration. Considering the similar lifetimes and smaller 2θ shift of XRD, we deduce that the calculated and should be and in SCAS: 0.1%Eu2+. With the increase of the Eu2+ concentration, decreases while increases firstly and then decreases. Furthermore, decreases while increases. The above results can strongly prove the existence of energy transfer from Eu (I) to Eu (II). In summary, the size mismatch of Eu2+, Sr2+, and Ca2+ and the energy transfer between the Eu2+ ions together lead to the red shift of the emission spectra. The energy transfer from Eu (I) to Eu (II) also contributes to the increasing FWHM of the emission spectra. A schematic diagram of the red shift is shown in Fig. 15. In order to clarify the energy transfer mechanism from Eu (I) to Eu (II), it is necessary to know the critical distance ( between the activators. With the increasing Eu2+ content, the distance between the Eu2+ ions becomes shorter, thus the probability of energy transfer increases. When the distance becomes small enough, the concentration quenching[42] occurs and the energy migration is hindered. The calculation of has been given by Blasse[43] where V corresponds to the volume of the unit cell, N is the number of host cations in the unit cell, and is the critical concentration of doping ions. For the SCAS host, N = 8, V = 1400 Å3, and is 3% for Eu2+. Accordingly, the critical distance ( is estimated to be about 22.33 Å. In general, there are three mechanisms for non-radiant energy transfer, including exchange interaction, radiation reabsorption, and electric multipolar interaction.

Fig. 14. (color online) Lifetimes of Eu2+ ions in different luminescence centers for different Eu2+ contents.
Fig. 15. (color online) Schematic diagram of red-shift for the emission spectra.

The result obtained above indicates little possibility of exchange interaction since the exchange interaction is predominant only for about 5 Å which is far less than 22.33 Å. Consequently, we can conclude that the energy transfer mechanism between the Eu2+ ions is electric multipolar interaction. According to the formula proposed by Dexter and Van Uitert, the emission intensity (I) per activator ion follows the equation[44,45] where I represents the emission intensity, x is the activator ion concentration, K and β are constants for the given matrix under the identical excitation conditions. The type of energy transfer mechanism of electric multipolar interaction can be estimated by analyzing the constant θ from this formula. The value of θ is 6, 8, 10, corresponding to electric dipole–dipole, dipole–quadrupole, and quadrupole–quadrupole interactions, respectively. The curve of log (I/x) versus log(x) for SCAS: Eu2+ phosphors beyond the quenching content of Eu2+ is plotted in Fig. 16. The θ value is obtained by the fitted straight line with a slope equal to . The θ is determined to be 4.41, which is close to 6. The result indicates that the concentration quenching mechanism of the Eu2+ emission in SCAS host is dominated by the dipole–dipole interaction.

Fig. 16. (color online) Fitting line of vs. in SACA: xEu2+ ( ) phosphors beyond the quenching concentration.
3.5. CIE chromaticity coordinates, quantum yields, and application of SCAS: Eu2+

As important parameters, CIE chromaticity coordinates and internal quantum yields (IQYs) of SCAS phosphors are also measured and presented in Table 2. The CIE chromaticity coordinates vary from (0.1575, 0.0792) to (0.1599, 0.1935) corresponding to x = 1%–10%, and the color shifts from blue to cyan with Eu2+ concentration increasing. The maximum IQY in as-prepared samples is 66.53%. To demonstrate the advantages of this blue-emitting phosphor, a series of samples are used to fabricate warm white LEDs. In order to obtain white light, commercial red phosphor (Ca Sr)SiAlN3: Eu2+ and green phosphor (Y Lu)3Al5O12: Ce3+ are employed to adjust the luminescence. The white LED devices of various Eu2+ concentrations under a driven current of 350 mA are given in Fig. 17. The relevant parameters are presented in

Fig. 17. (color online) Photographs of the fabricated warm white-LEDs driven by 350 mA current (x = 3%, 5%, 7%, 8%, and 10%).
Table 2.

CIE chromaticity coordinates and IQYs of SCAS: xEu2+ phosphors excited at 365 nm UV radiation.

.

Table 3. The CIE chromaticity coordinates of white light turn from (0.3582, 0.3591) to (0.4346, 0.4063). The color rendering indexes range from 86.6 to 89.6. The low correlated color temperatures vary from 3000 K to 4500 K, which meets the indoor lighting requirements. Moreover, the luminescent efficiency of x= 3% reaches 45 lm/W. The above results demonstrate that the white LED device of x= 3% may be an excellent candidate for warm white-LEDs.

Table 3.

Main parameters of the selected white LED devices.

.
4. Conclusions

A series of SCAS: Eu2+ blue-emitting phosphors were successfully prepared by a high-temperature solid-state reaction. The substitution priority of Eu2+ in Sr0.8Ca0.2Al2Si2O8 was investigated via XRD and temperature properties. It was found that the Ca2+ ions are preferentially substituted by Eu2+ at lower doping, and with the Eu2+ concentration increasing, the probability of substitution for Sr2+ is greater than that of replacing Ca2+. A potential method of using thermal property to determine the substitution of Eu2+ in the multi-cation hosts was proposed. Moreover, the luminescence properties, thermoluminescence spectra, and energy transfer mechanism between Eu2+ were studied in detail. SCAS: Eu2+ phosphors were used to fabricate warm white LEDs, and the maximal luminescent efficiency reached 45 lm/W. On the basis of these results, the blue phosphors will have a good prospect of application in the field of warm white LEDs.

Reference
[1] Zhou T L Song Z Song X P Bian L Liu Q L 2010 Chin. Phys. 19 127808
[2] Huang P Cui C E Wang S 2009 Chin. Phys. 18 4524
[3] Xie W Wang Y H Hu Y H Wu H Y Deng L Y Liao F 2010 Acta Phys. Sin. 59 1148
[4] Jiang Z Q Wang Y H Gong Y 2010 Chin. Phys. 19 027801
[5] Zhang J C Qin Q S Yu M H Zhou H L Zhou M J 2011 Chin. Phys. 20 094211
[6] Li P L Xu Z Zhao S L Wang Y S Zhang F J 2012 Chin. Phys. 21 047803
[7] Dai P P Cao J Zhang X T Liu Y C 2016 J. Phys. Chem. 120 18713
[8] Ji H P Wang L Molokeev M S Hirosaki N Xie R J Huang Z H Xia Z G ten Kate O M Liu L H Atuchin V V 2016 J. Mater. Chem. 4 6855
[9] He L Z Song Z Xiang Q C Xia Z G Liu Q L 2016 J. Lumin. 180 163
[10] Ji H P Huang Z H Xia Z G Molokeev M S Atuchin V V Fang M H Liu Y G 2015 J. Phys. Chem. 119 2038
[11] Huang C H Wu P J Lee J F Chen T M 2011 J. Mater. Chem. 21 10489
[12] Li K Shang M M Lian H Z Lin J 2015 Inorg. Chem. 54 7992
[13] Park W B Singh S P Sohn K S 2014 J. Am. Chem. Soc. 136 2363
[14] Denault K A Brgoch J Gaultois M W Mikhailovsky A Petry R Winkler H DenBaars S P Seshadri R 2014 Chem. Mater. 26 2275
[15] Li G G Lin C C Chen W T Molokeev M S Atuchin V V Chiang C Y Zhou W Z Wang C W Li W H 2014 Chem. Mater. 26 2991
[16] Wang S S Chen W T Li Y Wang J Sheu H S Liu R S 2013 J. Am. Chem. Soc. 135 12504
[17] Tsai Y T Chiang C Y Zhou W Lee J F Sheu H S Liu R S 2015 J. Am. Chem. Soc. 137 8936
[18] Im W B George N Kurzman J Brinkley S Mikhailovsky A Hu J Chmelka B F DenBaars S P Seshadri R 2011 Adv. Mater. 23 2300
[19] Li G G Tian Y Zhao Y Lin J 2015 Chem. Soc. Rev. 44 8688
[20] Xia Y F Chen J Liu Y G Molokeev M S Guan M Huang Z H Fang M H 2016 Dalton. Trans. 45 1007
[21] Chen D Q Zhou Y Xu W Zhong J S Ji Z Xiang W D 2016 J. Mater. Chem. 4 1704
[22] Li K Xu J Cai X C Fan J Zhang Y Shang M M Lian H Z Lin J 2015 J. Mater. Chem. 3 6341
[23] Bai Z H Gong B Y Tian Y G Zhang Q Song C 2008 J. Chin. Ceram. Soc. 36 1753
[24] Woo K G Lee J H Kim M H Lee Y J 2004 Inform. Software. Tech. 46 255
[25] Hakeem D K Park K 2015 New J. Chem. 39 7768
[26] Jiang Z Q Wang Y H Wang L S 2010 J. Electro. Chem. 157 J155
[27] Jia Y L Pang R Li H F Sun W Z Fu J P Jiang L H Liu R S 2015 Dalton. Trans. 44 11399
[28] Ma M X Zhu D C Zhao C Han T Cao S X Tu M J 2012 Opt. Commun. 285 665
[29] Wegh R T Donker H Oskam K D Meijerink A 1999 J. Lumin. 82 93
[30] Ding X Xu Y Guo C F 2010 Acta Phys. Sin. 59 6632
[31] Wang X C Zhao Z Y Wu Q S Li Y Y Wang C Mao A J Wang Y H 2015 Dalton. Trans. 44 11057
[32] Li K Geng D L Shang M M Zhang Y Lian H Z Lin J 2014 J. Phys. Chem. 118 11026
[33] Jang H S Im W B Lee D C Jeon D Y Kim S K 2007 J. Lumin. 126 371
[34] Shang M M Li C X Lin J 2014 Chem. Soc. Rev. 43 1372
[35] Yuan L L Zhang X S Xu J P Sun J Jin H Liu X J Li L L Li L 2015 Chin. Phys. 24 087802
[36] Lu F Zhang X H Lu Z M Xu X W Tang C C 2012 Acta Phys. Sin. 61 144209
[37] Kim J S Park Y H Choi J C Park H L Kim G C Yoo J K 2006 Solid State 137 187
[38] Qin C X Huang Y L Shi L Chn G Q Qiao X B Seo H J 2009 Appl. Phys. 42 185105
[39] Xu S Y Zhang X S Zhou Y L Xi Q Li L 2011 Chin. Phys. 20 037804
[40] Zhang J C Zhou M J Wang Y H 2012 Chin. Phys. 21 124102
[41] Chen J Y Zhang N M Guo C F Pan F J Zhou X J Suo H Zhao X Q Goldys E M 2016 ACS Appl. Mater. 8 20856
[42] Sharma G Lochab S P Singh N 2010 J. Alloys Compd. 501 190
[43] Wu L Wang B Zhang Y Li L Wang H R Yi H Kong Y F Xu J J 2014 Dalton. Trans. 43 13845
[44] Li P L Wang Y S Zhao S L Zhang F J Xu Z 2012 Chin. Phys. 21 127804
[45] Cao S X Han T Tu M J 2011 Acta Phys. Sin. 60 127802