Next Article in Journal
Transcriptomic Analysis of Vibrio parahaemolyticus Reveals Different Virulence Gene Expression in Response to Benzyl Isothiocyanate
Previous Article in Journal
Heparinized Polyurethane Surface Via a One-Step Photografting Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Controlled Synthesis of Tb3+/Eu3+ Co-Doped Gd2O3 Phosphors with Enhanced Red Emission

1
School of Materials Science and Engineering, University of Jinan, Jinan, Shandong 250022, China
2
Shandong Provincial Academy of Building Research, Jinan, Shandong 250031, China
3
Department of Physics and Astronomy, KU Leuven, 3001 Leuven, Belgium
*
Authors to whom correspondence should be addressed.
The two authors contributed equally to this paper.
Molecules 2019, 24(4), 759; https://doi.org/10.3390/molecules24040759
Submission received: 29 December 2018 / Revised: 14 February 2019 / Accepted: 15 February 2019 / Published: 20 February 2019

Abstract

:
(Gd0.93−xTb0.07Eux)2O3 (x = 0–0.10) phosphors shows great potential for applications in the lighting and display areas. (Gd0.93−xTb0.07Eux)2O3 phosphors with controlled morphology were prepared by a hydrothermal method, followed by calcination at 1100 °C. XRD, FE-SEM, PL/PLE, luminescent decay analysis and thermal stability have been performed to investigate the Eu3+ content and the effects of hydrothermal conditions on the phase variation, microstructure, luminescent properties and energy transfer. Optimum excitation wavelength at ~308 nm nanometer ascribed to the 4f8-4f75d1 transition of Tb3+, the (Gd0.93−xTb0.07Eux)2O3 phosphors display both Tb3+and Eu3+ emission with the strongest emission band at ~611 nm. For increasing Eu3+ content, the Eu3+ emission intensity increased as well while the Tb3+ emission intensity decreased owing to Tb3+→Eu3+ energy transfer. The energy transfer efficiencies were calculated and the energy transfer mechanism was discussed in detail. The lifetime for both the Eu3+ and Tb3+ emission decreases with the Eu3+ addition, the former is due to the formation of resonant energy transfer net, and the latter is because of contribution by Tb3+→Eu3+ energy transfer. The phosphor morphology can be controlled by adjusting the hydrothermal condition (reaction pH), and the morphological influence to the luminescent properties (PL/PLE, decay lifetime, etc.) has been studied in detail.

1. Introduction

The stable physical and chemical properties of Gd2O3 with cubic structure make it an important inorganic compound in luminescence applications. The Gd3+ in Gd2O3 could be easily substituted by an alternative rare earth activator ion (Eu3+, Tb3+, etc.) due to their similar ion radius (Gd3+, Eu3+, and Tb3+ have ion radii of 1.053 Å, 1.066 Å and 1.040 Å for coordination number 8) [1]. The Eu3+, Tb3+and Dy3+ doped Gd2O3 matrix can emit vivid red, green and yellow colors, which in turn supports their use in the field of lighting and display [2,3,4].
The (Gd0.93−xTb0.07Eux)2O3 system was chosen in light of: (1) the luminescent properties of phosphor are greatly affected by the particle morphology and size, which relied on the synthesis route used. [5,6,7]. The hydrothermal method is usually selected to control the particle morphology and size [8,9,10], which is also applied in the preparation of Gd0.93−xTb0.07EuxO3 systems in this work. Based on this, luminescent properties due to particle morphology and size were studied in detail; (2) due to higher 6IJ excited state of Gd3+ compared to 5D3,4 and 5D0,1 emission states of Tb3+ and Eu3+, the Gd3+ can sensitize the luminescence of Tb3+ and Eu3+ through Gd3+→Tb3+, Gd3+→Eu3+ energy transfer [11,12]. Meanwhile, Tb3+→Eu3+ energy transfer reported in numerous works can also boost Eu3+ red emission [13], and the energy transfer of Gd3+→Tb3+→Eu3+ may also occur; (3) the lower electronegativity (1.20) of Gd3+ compared to Y3+ (1.22) and Lu3+ (1.27) may result in easier inter- configurational transition, which can induce new properties and further improve the red emission intensity. Better luminescence features of Eu3+ and Tb3+ in Gd2O3 than Y2O3 and Lu2O3 lattices may then be obtained, which is further validated by experiments in this work.
In this paper, a series of (Gd0.93−xTb0.07Eux)2O3 (x = 0–0.10) phosphors were prepared through hydrothermal method, and the particle size and morphology were tuned by varying the reaction pH values. The phase structure, microstructure, luminescent properties, energy transfer efficiency and mechanism were analyzed by the combination of XRD, FE-SEM, PLE/PL and luminescent decay analysis. Moreover, morphology and size effect of the particle on the luminescent properties were investigated. In the sections that follow, we report in detail the synthesis, morphology/size controlled, luminescent traits, energy transfer and thermal stability of the phosphors.

2. Results and Discussion

The XRD patterns of precursors with different Eu3+ content are shown in Figure 1a. The diffraction peaks can be indexed as pure Gd(OH)3 (JCPDS NO.38-1042). All the samples show the same diffraction behavior, indicating that the Eu3+ addition does not significantly affect the crystal structure of the precursor. Figure 1b displays the XRD patterns of (Gd0.93−xTb0.07Eux)2O3 (x = 0–0.10) sintered at 1100 °C as a function of Eu3+ content (reaction pH = 9.0, hydrothermal temperature: 140 °C). The diffraction peaks of the calcined products can be indexed as pure Gd2O3 phase (JCPDS NO. 43-1014) and no other phases are observed. All the samples show the same diffraction behavior indicating that the Eu3+ addition does not affect the crystal structure.
Figure 2 illustrates the FE-SEM images of the (Gd0.93−xTb0.07Eux)2O3 precursor sintered at 1100 °C with x = 0.04 (a) and x = 0.1 (b), respectively (reaction pH = 9.0, hydrothermal temperature: 140 °C). All the precursors display rod-resemble structures with diameters of ~100 nm and lengths of ~500 nm. Comparison of the FE-SEM images in Figure 2a (x = 0.04) and Figure 2b (x = 0.1) shows that the Eu3+ incorporation does not alter the particle morphology. The particles (Gd0.93−xTb0.07Eux)2O3 calcined at 1100 °C possess good dispersion and uniform morphology (Figure 2c,d), and the rod-like morphology of the precursor persists. The main variation was that the particles grew and the overall outline was clearer and more easily distinguished.
Figure 3 shows FE-SEM micrographs of (Gd0.89Tb0.07Eu0.04)2O3 precursor synthesized at various pH values (pH 8–12, hydrothermal temperature: 140 °C). As we can see the particle morphology and size can be controlled by varying the pH value during synthesis. For the pH value of 8.0, the particles exhibit a tubular morphology (Figure 3a) with diameter of ~200 nm and length of ~800 nm. In contrast, a pH value of 9.0 results in a rod-like particle morphology (Figure 3b). The formation of tubular and rod-shaped phosphors strongly depends on the mass transfer rate. At a low pH value of 8.0, the mass transfer speed of inner part is lower than the outer region, which leads to tube formation. As the pH increased to 9.0, the mass transfer speed between inner and outer region is comparable which leads to the formation of the rod morphology. While the pH value is further adjusted from 9.0 to 12.0, the precursor size with rod-like shape gradually decreased from diameter of ~120 nm and length of ~500 nm to ~80 nm and ~100 nm, respectively. The reduction of the size is principally attributed to large nucleation density resulting from large pH value [14].
Figure 4 shows the excitation spectrum of the (Gd0.93−xTb0.07Eux)2O3 (x = 0.02–0.1) samples (reaction pH = 9.0, hydrothermal temperature: 140 °C, calcined temperature: 1100 °C) as a function of Eu3+ content at an emission wavelength of 542 nm (Tb3+ emission, Figure 4a) and 611 nm (Eu3+ emission, Figure 4b), respectively. With monitoring at 542 nm, the PLE spectra of the (Gd0.93−xTb0.07Eux)2O3 (x = 0.02–0.1) system displays one strong and broad peak centered at ~308 nm which is ascribed to the 4f8-4f75d1 transition of Tb3+ [15], whereas by monitoring at 611 nm (Figure 4b), the PLE spectra of (Gd0.93−xTb0.07Eux)2O3 phosphors contain two excitation bands at ~248 nm and ~308 nm which is ascribed to the charge transfer band (CTB) of Eu3+ [16] and the 4f8-4f75d1 transition of Tb3+, respectively. In addition, as we can see from the inline graph of b, the CTB excitation peak of Eu3+ at ~258 nm overlapped the characteristic transition 8S7/2-6IJ of Gd3+ implying the Gd3+→Eu3+ energy transfer. The occurrence of Gd3+ and Tb3+ on the PLE spectra monitoring the Eu3+ emission provide clear information for energy transfer of the Gd3+→Eu3+ and Tb3+→Eu3+ [17,18]. Therefore, not only the Tb3+ but also Eu3+ ions can be energized at ~308 nm. The PL spectra with 308 nm excitation are analyzed and presented in Figure 4c.
The PL spectra show the strongest emission band at ~611 nm (5D0-7F2 transition of Eu3+) accompanied by other relatively weak emission bands at ~542 nm, ~580 nm, ~593 nm, ~654 nm and ~687 nm contributed to the 5D4-7F5 transition of Tb3+, 5D0-7F0 transition of Eu3+, 5D0-7F1 transition of Eu3+, 5D0-7F3 transition of Eu3+, and 5D0-7F4 transition of Eu3+, respectively [19,20,21,22]. Both the appearance of the 5D0-7F0 transition of Eu3+ and the higher emission intensity of 5D0-7F2 transition of Eu3+ (~611 nm) compared with 5D0-7F1 transition of Eu3+ (~593 nm) imply that more Eu3+ occupies the relatively low symmetric lattice (C2) [23,24]. The intensity of the emission at 611 nm increases with an increasing Eu3+ content (up to x = 0.04), and then decreases because of the concentration quenching. Furthermore, the emission intensity of Tb3+ at ~542 nm (the inset in Figure 4c) decreases resulting from the energy transfer of Tb3+→Eu3+. Comparing the PL spectra of (Gd0.89Tb0.07Eu0.04)2O3, (Gd0.96Eu0.04)2O3 and (Y0.96Eu0.04)2O3 (Figure 4c), the emission intensity is found in the order (Gd0.89Tb0.07Eu0.04)2O3 > (Gd0.96Eu0.04)2O3 > (Y0.96Eu0.04)2O3 due to the efficient Gd3+→Eu3+ and Tb3+→Eu3+ energy transfer.
The luminescence quenching type of Eu3+ in solid phosphors can be obtained through evaluating the parameter s as indicated in Equation (1) [25,26,27,28]:
log ( I c ) = ( s d ) log ( c ) + log f
where I represents the Eu3+ emission intensity, c is the Eu3+ concentration, d = 3 for a regular sample, f is a constant, and s is the electric multipole index. When values of 3, 6, 8 and 10 are assigned to s, different exchange interaction, dipole-dipole, dipole-quadrupole, and quadrupole-quadrupole electric interactions are obtained, respectively. The log(I/c)-log(c) plot that corresponds to emission at 611 nm is shown in Figure 4d. The fitted slope (−s/3) was calculated to be −1.13, thus s = 3.42 (~3) for the (Gd0.93−xTb0.07Eux)2O3 systems, indicating that concentration quenching is mostly caused by the energy transfer between Eu3+ ions [26,29].
The energy level diagram and energy transfer between Gd3+, Tb3+ and Eu3+ are shown in Figure 5. At 275 nm excitation, the electrons of Gd3+ are excited from the 8S7/2 to the 6IJ state, then relaxed to 6P7/2 state. On the other hand, UV excitation makes the electrons of Tb3+ and Eu3+ shift from the 7FJ (J = 3, 4, 5, 6 for Tb3+) and 7FJ (J = 0, 1, 2, 3, 4 for Eu3+) to the 5D3 (Tb3+) and 5D1 (Eu3+) states followed by relaxation to 5D4 (Tb3+) and 5D0 (Eu3+), respectively. Because the energy level of the 6P7/2 state lies higher than the 5D4 levels of Tb3+ and the 5D0 level of Eu3+, the part energy of Gd3+ can be transferred to Tb3+ and Eu3+ [30], respectively. Meanwhile, energy transfer from Tb3+ to Eu3+ due to the higher energy level of 5D4 (Tb3+) compared to 5D0 (Eu3+) can happen. The electrons of 5D4 (Tb3+) and 5D0 (Eu3+) states jump back to the ground state 7FJ, thereby producing green (Tb3+) and red (Eu3+) emissions [31].
In order to calculate the energy transfer efficiency between Tb3+ and Eu3+, the luminescence decay behavior of Tb3+ at 542 nm was investigated using the x = 0.04 and the results are shown in Figure 6a. As we can see that the kinetics of decay follow a single exponential decay behavior:
I = A exp ( t τ R ) + B
where I refers to luminescence intensity, t represents the decay time τR denotes the lifetime and A and B are the constants [32]. The fitted result yields A = 8424.79 ± 821.77 (au), B = 100.24 ± 24.60 (au) and τR = 0.16 ± 0.01 ms. The lifetime values for Tb3+ shown in the inset of Figure 6a decrease gradually with increasing Eu3+ content because of energy transfer Tb3+→Eu3+. with transfer efficiency (ηET) being obtained by evaluating the lifetime of Tb3+ with (τS) and without (τS0) Eu3+ doping through Equation (3) [33]:
η E T = 1 τ S τ S 0
The results of energy transfer efficiency calculation are shown in Figure 6b. As can be seen, ηET has a positive correlation with Eu3+ concentration where increased Eu3+ content, from x = 0.02 to x = 0.10, leads to gradually enhanced efficiency of energy transfer, from 89.7% to 98.7%. By consequence, the sensitizer of Tb3+ plays a critical part in the luminescence emission of Eu3+ with large ηET value predominantly generating from substantial overlapping of spectra between the 5D47FJ emissions of Tb3+ and the 7F0,15D0,1 absorption of Eu3+ [34]. Figure 6c shows the lifetime value of Eu3+ for 611 nm emission relative to Eu3+ content, through where we can see that the lifetime of Eu3+ decreases from 2.24 to 1.19 ms with Eu3+ addition from x = 0.02 to x = 0.10, resulting from the formation of a resonant energy transfer net among the activators. Figure 6d depicts the CIE chromaticity coordinates for (Gd0.89Tb0.07Eu0.04)2O3 phosphors with 308 nm excitation. The CIE chromaticity coordinate and color temperature are determined to be (~0.64, ~0.35) and ~2439 K, respectively, as a result the phosphors gives a vivid red color.
The energy transfer mechanism between Tb3+→Eu3+ can be analyzed according to Dexter’ and Reisfeld’s theory [35,36], and the explanation is given as in the equations below:
ln I S 0 I S C
I S 0 I S C n 3
where C is the summed concentration of doped ions Tb3+ and Eu3+; IS0 and IS are the emission intensities of Tb3+ for 542 nm emission with and without Eu3+; lnIs0/Is-C corresponds to exchange interactions, and lnIs0/Is-Cn/3 for n = 6, 8, 10 represent the dipole-dipole, dipole-quadrupole and quadrupole-quadrupole electric interactions, respectively. The plots of lnIs0/Is-C and lnIs0/Is-Cn/3 are illustrated in Figure 7. By comparing the fitted factor values (R), the best linear relationship was found for n = 10, which clearly shows energy transfer from Tb3+→Eu3+ in the (Gd1−xTb0.07Eux)2O3 phosphor is dominated by quadrupole-quadrupole electric interactions [27].
Considering that the change of hydrothermal pH values can alter the particle morphology (Figure 3), and the shape/size has a significant effect on the luminescent properties, we investigated the PL spectra of the (Gd0.89Tb0.07Eu0.04)2O3 sample as a function of pH value (pH = 8–12, Figure 8a, hydrothermal temperature: 140 °C, calcined temperature: 1100 °C). From Figure 8, we can conclude that the pH value variation has no influence to the shape of the emission peak, however it affects the emission intensity of Eu3+ dramatically. The emission intensity first decreases with the increasing pH till pH = 9.0. Thereafter it increases as the pH further increases up to 12.0. When the pH varies from 8.0 to 9.0, and the particle morphology changes from tubular to rods, with the latter presenting directional growth as described in Figure 3b–e. The phosphors with rod-like morphology could decrease the electric dipole transition probabilities of Eu3+, therefore decreasing the luminescence intensity [28]. For pH changing from 9.0 to 12.0, the particle dimension progressively decreases while the surface area gradually increases. As a result, the luminescent center number on the particle surface increases leading to an improved intensity of emission.
Figure 8b displays the lifetime values of the 611 nm emission with different synthesis pH values. The lifetime increases from 1.42 to 2.02 ms with the pH increasing from 8.0 to 12.0. The extended lifetime can be expressed via Equation (6) [25,37]:
τ R ~ 1 f ( E D ) λ 0 2 [ 1 3 ( n e f f 2 + 2 ) ] 2 n e f f
where f(ED) and λ0 are represent the dipole transition oscillator strength and the wavelength in vacuum, respectively. neff is the effective refractive index which is influenced by the particle size and decreases for smaller particles when applied to intermediately-sized particles as in this work. Thus, the neff decreased at a larger given pH value, and a longer lifetime was obtained. The influences of the defects of lattice on luminescent lifetime, nevertheless, can in no way be totally excluded. Deep traps are believed to be capable of arresting electrons temporarily, thus leading to a longer lifetime.
The thermal stability for phosphor materials is an important parameter for its potential application. The influences of temperature variation to the intensity of emission was investigated in the range of 298–523 K using (Gd0.89Tb0.07Eu0.04)2O3 as an example (reaction pH = 8.0, hydrothermal temperature: 140 °C, calcination temperature: 1100 °C), and the activation energy was also calculated in this work. Owing to the thermal quenching, the emission intensity of (Gd0.89Tb0.07Eu0.04)2O3 phosphor decreased with increasing temperature (Figure 9a). The temperature resulted thermal quenching can be explained using Arrhenius equation [27,38]:
ln ( I 0 I 1 ) = ln A E a k T
where Ea is the activation energy, T denotes temperature, A is a constant and k refers to the Boltzmann constant. I0 is the emission intensity at room temperature while I corresponds to the emission intensity at the related operating temperature. The variation of ln[(I0I)/I] in terms of 1/kT for the thermal quenching is shown in Figure 9b. The slope of the fitting curve is −0.211, which corresponds to the Ea value of 0.211 eV being almost the same as the 0.212 eV value for the Gd2O3:Dy3+/Eu3+ system [39,40]. The larger activation energy means that the synthesized phosphor has a more stable thermal stability compared to other reported phosphors and can be potentially used in lighting and display areas [41].

3. Summary

Pure-phase (Gd0.93−xTb0.07Eux)2O3 (x = 0.02–0.1) phosphors with controlled morphology were synthesized by hydrothermal method, followed by calcination. The combined technologies of XRD, FE-SEM, PLE/PL, decay behavior and thermal stability have been applied to analyze the products. The analysis results can be summarized as follows:
(1)
Increasing Eu3+ content does not change the particle morphology, but both the particle shape and size can be controlled by tuning the pH value used in the hydrothermal synthesis. The particle morphology varies from tubular to rod-like when the pH value increases from 8.0 to 9.0. The rod-like particle size decreases with the pH value when increased from 9.0 to 12.0;
(2)
(Gd0.93−xTb0.07Eux)2O3 phosphors exhibit a vivid red emission with a CIE chromaticity coordinate and color temperature of (~0.64, ~0.35) and ~2439 K, respectively. The quenching concentration was x = 0.04, and determined to be due to energy transfer between Eu3+. Comparing to the (Gd0.96Eu0.04)2O3 and (Y0.96Eu0.04)2O3 oxides, the (Gd0.89Tb0.07Eu0.04)2O3 possesses better luminescent properties due to Tb3+→Eu3+, Gd3+→Eu3+ energy transfer;
(3)
The influence of particle shape or size on the luminescence features, e.g. PLE/PL, lifetime, of resultant phosphors was investigated. The related energy transfer efficiency, mechanism, process and thermal stability were also analyzed in detail.

4. Experimental Procedures

The chemical reagents used in the synthesis include rare earth oxides (Gd2O3, Tb4O7, and Eu2O3, 99.99% pure, Jining Zhongkai New Type Material Science Co. Ltd, Jining, China), ammonia (NH3∙H2O, analytical grade 25 wt%) and nitric acid (HNO3, analytical grade 68 wt%). Both acids were purchased from Sinopharm Chemical Reagent Co. Ltd. (Shanghai, China). All reagents were utilized as starting material with no additional purification.
The whole synthesis process is shown in Figure 10. The rare earth nitrates RE(NO3)3 (RE = Gd, Tb, Eu) were provided via dissolving the corresponding oxides, Gd2O3, Tb4O7 and Eu2O3, in hot nitric acid. RE(NO3)3 was mixed as mother salt and stirred for 30 minutes according to the stoichiometric ratio (Gd0.93−xTb0.07Eux)2O3. Ammonia was used to adjust the pH of the mother salt, and the resulting turbid liquid was aged for 30 min. The turbid liquids were transferred to an autoclave and heated in an oven for 24 h. Upon completion of the reaction, the suspension was cooled to room temperature, followed by centrifugation and repeated washing using distilled water and alcohol to give a precipitate. The wet precipitate was dried at 180 °C for 24 h in air. The precursors were firstly decomposed at 600 °C for 4 h in the air, and then calcined at 1100 °C for 4 h in Ar/H2 (5 vol.% H2) gas mixture to obtain the resultant oxides. The Eu3+ content (x = 0–0.10) and reaction pH (pH = 8.0–12.0) were varied to study their effects on the particle morphology and size.
Phosphor phases were identified by X-ray diffractometry (XRD, Model PW3040/60, PANALYTICAL B.V, Almelo, The Netherlands) with nickel-filtered CuKα radiation and a 4° 2θ/min scanning speed. Particle morphological distribution was studied by field-emission scanning electron microscopy (FE-SEM, Model JSM-7001F, JEOL, Tokyo, Japan). Photoluminescence excitation (PLE) and photoluminescence (PL) spectra of the phosphors were collected by a FP-6500 fluorospectrophotometer (JASCO, Tokyo, Japan) at room temperature, which has an integrating sphere (Model ISF-513, JASCO) of diameter of 60 mm and an excitation source, Xe lamp, 150 W. The decay kinetic of Eu3+ and Tb3+ emission was acquired at room temperature. By exciting the phosphor powder at a chosen wavelength, the emission intensity was detected as to the elapsed time immediately after the excitation light was blocked by a shutter.

Author Contributions

For research articles with several authors, a short paragraph specifying their individual contributions must be provided. The following statements should be used “conceptualization, D.Z.; methodology, X.G.; software, Q.L.; validation, H.W.; formal analysis, X.G.; investigation, D.Z.; resources, Z.L.; data curation, H.W.; writing—original draft preparation, D.Z.; writing—review and editing, J.L.; visualization, L.M. and J.L.; supervision, J.L. and Z.L.; project administration, Q.L.; funding acquisition, Z.L.

Funding

This work was supported in part by the National Natural Science Foundation of China (No. 51402125), China Postdoctoral Science Foundation (No. 2017M612175), the Research Fund for the Doctoral Program of University of Jinan (No. XBS1447), the Natural Science Foundation of University of Jinan (No. XKY1515), the Science Foundation for Post Doctorate Research from the University of Jinan (No. XBH1607), the Special Fund of Postdoctoral innovation project in Shandong province (No. 201603061).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Cryst. 1976, A32, 751–767. [Google Scholar] [CrossRef]
  2. Li, J.H.; Liang, Q.Y.; Hong, J.Y.; Yan, J.; Dolgov, L.; Meng, Y.Y.; Xu, Y.Q.; Shi, J.X.; Wu, M.M. White light-emitting and enhanced color stability in a single component host. ACS Appl. Mater. Inter. 2018, 10, 18066–18072. [Google Scholar] [CrossRef] [PubMed]
  3. Li, Y.; Geceviciusa, M.; Qiu, J.R. Long persistent phosphors-from fundamentals to applications. Chem. Soc. Rev. 2016, 45, 2090–2136. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, Y.; Jie, W.J.; Chen, P.; Liu, W.W.; Hao, J.H. Ferroelectric and piezoelectric effects on the optical process in advanced materials and devices. Adv. Mater. 2018, 34, 1707007. [Google Scholar] [CrossRef] [PubMed]
  5. Chhabra, V.; Pillai, V.; Mishra, B.K.; Morrone, A.; Shah, D.O. Synthesis, characterization, and properties of microemulsion-mediated nanophase TiO2 particles. Langmuir 1995, 11, 3307–3311. [Google Scholar] [CrossRef]
  6. Xuan, T.T.; Yang, X.F.; Lou, S.Q.; Huang, J.J.; Liu, Y.; Yu, J.B.; Li, H.L.; Wong, K.L.; Wang, C.X.; Wang, J. Highly stable CsPbBr3 quantum dots coated with alkyl phosphate for white light-emitting diodes. Nanoscale 2017, 9, 15286–15290. [Google Scholar] [CrossRef]
  7. Zou, R.; Gong, S.M.; Shi, J.P.; Jiao, J.; Wong, K.L.; Zhang, H.W.; Wang, J.; Su, Q. Magnetic-NIR persistent luminescent dual-modal ZGOCS@MSNs@Gd2O3 core−shell nanoprobes for In vivo imaging. Chem. Mater. 2017, 29, 3938–3946. [Google Scholar] [CrossRef]
  8. Pang, X.C.; He, Y.J.; Jung, J.; Lin, Z.Q. 1D nanocrystals with precisely controlled dimensions, compositions, and architectures. Science 2016, 353, 1268–1272. [Google Scholar] [CrossRef]
  9. Li, J.H.; Zhang, Z.H.; Li, X.H.; Xu, Y.Q.; Ai, Y.Y.; Yan, J.; Shi, J.X.; Wu, M.M. Luminescence properties and energy transfer of YGa1.5Al1.5(BO3)4: Tb3+, Eu3+ as a multi-colour emitting phosphor for WLEDs. J. Mater. Chem. C 2017, 25, 6294–6299. [Google Scholar] [CrossRef]
  10. Li, J.H.; Yan, J.; Wen, D.W.; Khan, W.U.; Shi, J.X.; Wu, M.M.; Su, Q.; Tanner, P.A. Advanced Red Phosphors for White Light-emitting Diodes. J. Mater. Chem. C 2016, 37, 8611–8623. [Google Scholar] [CrossRef]
  11. Liu, L.; Fu, G.R.; Li, B.N.; Lu, X.Q.; Wong, W.K.; Jones, R.A. Single-component Eu3+-Tb3+-Gd3+-grafted polymer with ultra-high color rendering index white-light emission. Rsc. Adv. 2017, 11, 6762–6771. [Google Scholar] [CrossRef]
  12. Anh, T.K.; Ngoc, T.; Nga, P.T.; Bich, V.T.; Long, P. Energy transfer between Tb3+ and Eu3+ in Y2O3 crystals. J. Lumin. 1988, 39, 215–221. [Google Scholar]
  13. Reddy, G.V.L.; Moorthy, L.R.; Chengaiah, T.; Jamalaiah, B.C. Multi-color emission tunability and energy transfer studies of YAl3(BO3)4: Eu3+/Tb3+ phosphors. Ceram. Int. 2014, 40, 3399–3410. [Google Scholar] [CrossRef]
  14. Li, G.R.; Lu, X.H.; Zhao, W.X.; Su, C.Y.; Tong, Y.X. Controllable electrochemical synthesis of Ce4+-doped ZnO nanostructures from nanotubes to nanorods and nanocages. Cryst. Growth Des. 2008, 8, 1276–1281. [Google Scholar] [CrossRef]
  15. Wu, X.L.; Li, J.G.; Li, J.K.; Zhu, Q.; Li, X.D.; Sun, X.D.; Sakka, Y. Layered rare-earth hydroxide and oxide nanoplates of the Y/Tb/Eu system: Phase-controlled processing, structure characterization and color-tunable photoluminescence via selective excitation and efficient energy transfer. Sci. Technol. Adv. Mater. 2013, 14, 015006. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, Y.; Zhang, X.J.; Zhang, H.R.; Zheng, L.L.; Zeng, Y.; Lin, Y.; Liu, Y.L.; Lei, B.F. Tunable emission from green to red in the GdSr2AlO5:Tb3+, Eu3+ phosphor via efficient energy transfer, nanophosphors. Rsc Adv. 2018, 8, 3530–3535. [Google Scholar] [CrossRef]
  17. Liu, L.L.; Wang, Q.; Gao, C.J.; Chen, H.; Liu, W.S.; Tang, Y. Dramatically enhanced luminescence of layered terbium hydroxides as induced by the synergistic effect of Gd3+ and organic sensitizers. J. Phys. Chem. C 2014, 118, 14511–14520. [Google Scholar] [CrossRef]
  18. Jain, A.; Hirata, G.A. Photoluminescence, size and morphology of red-emitting Gd2O3:Eu3+, nanophosphor synthesized by various methods. Ceram. Int. 2016, 42, 6428–6435. [Google Scholar] [CrossRef]
  19. Xin, F.X.; Zhao, S.L.; Xu, S.Q.; Huang, L.H.; Jia, G.H.; Deng, D.G.; Wang, H.P. Structure and luminescence properties of Eu/Tb codoped oxyfluoride glass ceramics containing Sr2GdF7 nanocrystals. Opt. Mater. 2011, 34, 85–88. [Google Scholar] [CrossRef]
  20. Li, S.; Guo, N.; Liang, Q.M.; Ding, Y.; Zhou, H.T.; Ouyang, R.Z.; Lü, W. Energy transfer and color tunable emission in Tb3+, Eu3+ co-doped Sr3LaNa(PO4)3F phosphors. Spectrochim Acta A 2018, 190, 246–252. [Google Scholar] [CrossRef]
  21. Li, J.; Chen, L.; Zhang, J.H.; Hao, Z.D.; Luo, Y.S.; Zhang, L.G. Photoluminescence properties of a novel red-emitting phosphor Eu3+ activated scandium molybdate for white light emitting diodes. Mater. Res. Bull. 2016, 83, 290–293. [Google Scholar] [CrossRef]
  22. Li, F.H.; Liu, H.; Wei, S.L.; Sun, W.; Yu, L.X. Photoluminescent properties of Eu3+ and Tb3+ codoped Gd2O3 nanowires and bulk materials. J. Rare Earth. 2013, 111, 1063–1068. [Google Scholar] [CrossRef]
  23. Gai, S.L.; Yang, P.P.; Wang, D.; Li, C.X.; Niu, N.; He, F.; Li, X.B. Monodisperse Gd2O3: Ln (Ln=Eu3+, Tb3+, Dy3+, Sm3+, Yb3+/Er3+, Yb3+/Tm3+, and Yb3+/Ho3+) nanocrystals with tunable size and multicolor luminescent properties. Cryst. Eng. Comm. 2011, 13, 5480–5487. [Google Scholar] [CrossRef]
  24. Bi, H.F.; Li, X. Self-assembled columnar structure Gd2O3:Eu3+: Solvothermal synthesis and luminescence properties. Integr. Ferroelectr. 2012, 135, 119–124. [Google Scholar] [CrossRef]
  25. Blasse, G. Energy transfer in oxidic phosphors. Philips Res. Rep. 1969, 24, 131–144. [Google Scholar] [CrossRef]
  26. Chen, Q.; Li, J.K.; Wang, W.Z. Synthesis and luminescence properties of Tb3+/Eu3+ co-doped GdAlO3 phosphors with enhanced red emission. J. Rare Earths 2018, 36, 924–930. [Google Scholar] [CrossRef]
  27. Dexter, D.L.; Schulman, J.H. Theory of concentration quenching in inorganic phosphors. J. Chem. Phys. 1954, 22, 1063–1070. [Google Scholar] [CrossRef]
  28. Dexter, D.L. A theory of sensitized luminescence in solids. J. Chem. Phys. 1953, 21, 836–850. [Google Scholar] [CrossRef]
  29. Reisfeld, R.; Greenberg, E.; Velapoldi, R.; Barnett, B. Luminescence quantum efficiency of Gd and Tb in borate glasses and the mechanism of energy transfer between them. J. Chem. Phys. 1972, 56, 1698–1705. [Google Scholar] [CrossRef]
  30. Zhu, Q.; Li, J.G.; Li, X.D.; Sun, X.D. Selective processing, structural characterization, and photoluminescence behaviors of single crystalline (Gd1-xEux)2O3 nanorods and nanotubes. Curr. Nanosci. 2010, 6, 496–504. [Google Scholar] [CrossRef]
  31. Dai, Q.L.; Song, H.W.; Wang, M.Y.; Bai, X.; Dong, B.; Qin, R.F.; Qu, X.S.; Zhang, H. Size and concentration effects on the photoluminescence of La2O2S:Eu3+ nanocrystals. J. Phys. Chem. C 2008, 112, 19399–19404. [Google Scholar] [CrossRef]
  32. Li, Y.; Hong, G. Synthesis and luminescence properties of nanocrystalline Gd2O3:Eu3+ by combustion process. J. Lumin. 2007, 124, 297–301. [Google Scholar] [CrossRef]
  33. Mutelet, B.; Perriat, P.; Ledoux, G.; Amans, D.; Lux, F.; Tillement, O.; Billotey, C.; Janier, M.; Villiers, C.; Bazzi, R.; et al. Suppression of luminescence quenching at the nanometer scale in Gd2O3 doped with Eu3+ or Tb3+: Systematic comparison between nanometric and macroscopic samples of life-time, quantum yield, radiative and non-radiative decay rates. J. Appl. Phys. 2011, 110, 4317–4326. [Google Scholar] [CrossRef]
  34. Jia, G.; Liu, K.; Zheng, Y.H.; Song, Y.H.; Yang, M.; You, H.P. Highly uniform Gd(OH)3 and Gd2O3:Eu3+ nanotubes: Facile synthesis and luminescence properties. J. Phy. Chem. C 2009, 113, 6050–6055. [Google Scholar] [CrossRef]
  35. Devaraju, M.K.; Yin, S.; Sato, T. Solvothermal synthesis, controlled morphology and optical properties of Y2O3:Eu3+ nanocrystals. J. Cryst. Growth 2009, 311, 580–584. [Google Scholar] [CrossRef]
  36. Ramírez, A.D.J.M.; Murillo, A.G.; Romo, F.D.J.C.; Salgado, J.R.; Luyer, C.L.; Chadeyron, G.; Boyer, D.; Palmerin, J.M. Preparation and studies of Eu3+ and Tb3+ co-doped Gd2O3and Y2O3 sol-gel scintillating films. Thin Solid Films 2009, 517, 6753–6758. [Google Scholar] [CrossRef]
  37. Zhang, K.; Holloway, T.; Pradhan, A.K.; Cui, Y.; Bhattacharya, P.; Burger, A.; Kar, A.; Patra, A. Synthesis and optical properties of Eu3+ and Tb3+ doped Gd2O3 nanotubes. Sci. Adv. Mate. 2012, 4, 649–655. [Google Scholar] [CrossRef]
  38. Xu, Z.H.; Yang, J.; Hou, Z.Y.; Li, C.X.; Zhang, C.M.; Huang, S.S.; Lin, J. Hydrothermal synthesis and luminescent properties of Y2O3:Tb3+ and Gd2O3:Tb3+ microrods. Mater. Res. Bull. 2009, 44, 1850–1857. [Google Scholar] [CrossRef]
  39. Meltzer, R.S.; Feofilov, S.P.; Tissue, B.; Yuan, H.B. Dependence of fluorescence lifetimes of Y2O3:Eu3+, nanoparticles on the surrounding medium. Phys. Rev. B 1999, 60, 14012–14015. [Google Scholar] [CrossRef]
  40. Liu, B.; Li, J.K.; Duan, G.B.; Li, Q.G.; Liu, Z.M. The synthesis and luminescent properties of morphology-controlled Gd2O3:Dy3+/Eu3+ phosphors with enhanced red emission via energy transfer. J. Lumin. 2019, 206, 348–358. [Google Scholar] [CrossRef]
  41. Yang, J.; Quan, Z.W.; Kong, D.Y.; Liu, X.M.; Lin, J. Y2O3:Eu3+ microspheres: Solvothermal synthesis and luminescence properties. Cryst. Growth Des. 2007, 7, 730–735. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds are not available from authors.
Figure 1. (a) is XRD spectra of precursors doped with different Eu contents, (b) is XRD spectra of (Gd0.93−xTb0.07Eux)2O3 (x = 0–0.10) precursors calcined at 1100 °C.
Figure 1. (a) is XRD spectra of precursors doped with different Eu contents, (b) is XRD spectra of (Gd0.93−xTb0.07Eux)2O3 (x = 0–0.10) precursors calcined at 1100 °C.
Molecules 24 00759 g001
Figure 2. FE-SEM micrograph of the (Gd0.93−xTb0.07Eux)2O3 precursor with x = 0.04 (a) and x = 0.1 (b) and of the resultant product calcined at 1000 °C, x = 0.04 (c) and x = 0.1 (d).
Figure 2. FE-SEM micrograph of the (Gd0.93−xTb0.07Eux)2O3 precursor with x = 0.04 (a) and x = 0.1 (b) and of the resultant product calcined at 1000 °C, x = 0.04 (c) and x = 0.1 (d).
Molecules 24 00759 g002
Figure 3. The FE-SEM morphologies of (Gd0.89Tb0.07Eu0.04)2O3 precursor synthesized with pH = 8 (a), 9 (b), 10 (c), 11 (d), 12 (e), respectively (hydrothermal temperature: 140 °C).
Figure 3. The FE-SEM morphologies of (Gd0.89Tb0.07Eu0.04)2O3 precursor synthesized with pH = 8 (a), 9 (b), 10 (c), 11 (d), 12 (e), respectively (hydrothermal temperature: 140 °C).
Molecules 24 00759 g003
Figure 4. Figures (a) and (b) are the excitation spectra (a, λem = 542 nm; b, λem = 611 nm) of (Gd0.93−xTb0.07Eux)2O3 (x = 0.02–0.1) phosphor calcined at 1100 °C as a function of Eu3+ content. The inset in (b) is the excitation spectrum contrast map of Gd1.92Eu0.08O3 and Y1.92Eu0.08O3 under λem = 611 nm. Figure (c) shows the emission spectra (λex = 308 nm) of (Gd0.93−xTb0.07Eux)2O3 (x = 0.02–0.1), the (Gd0.96Eu0.04)2O3 (λex = 258 nm) and (Y0.96Eu0.04)2O3 (λex = 258 nm) were included for comparison. Inset is excitation spectra corresponding to Gd1.78Tb0.14Eu0.08O3, Gd1.92Eu0.08O3 and Y1.92Eu0.08O3 with emission peak of 611 nm. The inset in (c) is the enlarged graph of the Tb3+ emission peak. Figure (d) describes log(I/c) variation as related to log(c) for the (Gd0.93−xTb0.07Eux)2O3 phosphors calcined at 1100 °C (611 nm emission).
Figure 4. Figures (a) and (b) are the excitation spectra (a, λem = 542 nm; b, λem = 611 nm) of (Gd0.93−xTb0.07Eux)2O3 (x = 0.02–0.1) phosphor calcined at 1100 °C as a function of Eu3+ content. The inset in (b) is the excitation spectrum contrast map of Gd1.92Eu0.08O3 and Y1.92Eu0.08O3 under λem = 611 nm. Figure (c) shows the emission spectra (λex = 308 nm) of (Gd0.93−xTb0.07Eux)2O3 (x = 0.02–0.1), the (Gd0.96Eu0.04)2O3 (λex = 258 nm) and (Y0.96Eu0.04)2O3 (λex = 258 nm) were included for comparison. Inset is excitation spectra corresponding to Gd1.78Tb0.14Eu0.08O3, Gd1.92Eu0.08O3 and Y1.92Eu0.08O3 with emission peak of 611 nm. The inset in (c) is the enlarged graph of the Tb3+ emission peak. Figure (d) describes log(I/c) variation as related to log(c) for the (Gd0.93−xTb0.07Eux)2O3 phosphors calcined at 1100 °C (611 nm emission).
Molecules 24 00759 g004
Figure 5. Energy level diagram and energy transfer mechanism of Gd3+, Tb3+ and Eu3+ in (Gd0.93−xTb0.07Eux)2O3 (x = 0.02–0.1) phosphor.
Figure 5. Energy level diagram and energy transfer mechanism of Gd3+, Tb3+ and Eu3+ in (Gd0.93−xTb0.07Eux)2O3 (x = 0.02–0.1) phosphor.
Molecules 24 00759 g005
Figure 6. (a) The decay behavior (Gd0.89Tb0.07Eu0.04)2O3 (reaction pH = 9.0, hydrothermal temperature: 140 °C, calcination temperature: 110 °C) for the 542 nm emission of Tb3+ (λex = 308 nm). The inset is the lifetime variation against Eu3+ content; (b) the calculated energy transfer efficiency between Tb3+ and Eu3+as function of Eu3+ content; (c) the lifetime value of Eu3+ for 611 nm emission with the change of Eu3+ content (λex = 308 nm); (d) the CIE chromaticity diagram for the emission of (Gd0.89Tb0.07Eu0.04)2O3 phosphors under 308 nm excitation.
Figure 6. (a) The decay behavior (Gd0.89Tb0.07Eu0.04)2O3 (reaction pH = 9.0, hydrothermal temperature: 140 °C, calcination temperature: 110 °C) for the 542 nm emission of Tb3+ (λex = 308 nm). The inset is the lifetime variation against Eu3+ content; (b) the calculated energy transfer efficiency between Tb3+ and Eu3+as function of Eu3+ content; (c) the lifetime value of Eu3+ for 611 nm emission with the change of Eu3+ content (λex = 308 nm); (d) the CIE chromaticity diagram for the emission of (Gd0.89Tb0.07Eu0.04)2O3 phosphors under 308 nm excitation.
Molecules 24 00759 g006
Figure 7. The relationship lnIs0/Is-C (a) and IS0/IS-Cn/3 of (Gd0.89Tb0.07Eu0.04)2O3 (reaction pH = 9.0, hydrothermal temperature: 140 °C, calcination temperature: 1100 °C) with n = 6 (b), n = 8 (c), n = 10 (d), respectively.
Figure 7. The relationship lnIs0/Is-C (a) and IS0/IS-Cn/3 of (Gd0.89Tb0.07Eu0.04)2O3 (reaction pH = 9.0, hydrothermal temperature: 140 °C, calcination temperature: 1100 °C) with n = 6 (b), n = 8 (c), n = 10 (d), respectively.
Molecules 24 00759 g007
Figure 8. (a) emission spectrum of (Gd0.89Tb0.07Eu0.04)2O3 synthesized with different pH values marked in the figure (λex = 308 nm); (b) lifetime values of (Gd0.89Tb0.07Eu0.04)2O3 for the 611 nm emission of Eu3+ as a function of pH value.
Figure 8. (a) emission spectrum of (Gd0.89Tb0.07Eu0.04)2O3 synthesized with different pH values marked in the figure (λex = 308 nm); (b) lifetime values of (Gd0.89Tb0.07Eu0.04)2O3 for the 611 nm emission of Eu3+ as a function of pH value.
Molecules 24 00759 g008
Figure 9. Temperature-dependent PL intensity is shown in (a), and the relationship between ln(I0/I−1) and 1/kT is displayed in (b).
Figure 9. Temperature-dependent PL intensity is shown in (a), and the relationship between ln(I0/I−1) and 1/kT is displayed in (b).
Molecules 24 00759 g009
Figure 10. The synthesis scheme of (Gd0.93−xTb0.07Eux)2O3 phosphors.
Figure 10. The synthesis scheme of (Gd0.93−xTb0.07Eux)2O3 phosphors.
Molecules 24 00759 g010

Share and Cite

MDPI and ACS Style

Zhu, D.; Li, J.; Guo, X.; Li, Q.; Wu, H.; Meng, L.; Liu, Z. Controlled Synthesis of Tb3+/Eu3+ Co-Doped Gd2O3 Phosphors with Enhanced Red Emission. Molecules 2019, 24, 759. https://doi.org/10.3390/molecules24040759

AMA Style

Zhu D, Li J, Guo X, Li Q, Wu H, Meng L, Liu Z. Controlled Synthesis of Tb3+/Eu3+ Co-Doped Gd2O3 Phosphors with Enhanced Red Emission. Molecules. 2019; 24(4):759. https://doi.org/10.3390/molecules24040759

Chicago/Turabian Style

Zhu, Dong, Jinkai Li, Xiangyang Guo, Qinggang Li, Hao Wu, Lei Meng, and Zongming Liu. 2019. "Controlled Synthesis of Tb3+/Eu3+ Co-Doped Gd2O3 Phosphors with Enhanced Red Emission" Molecules 24, no. 4: 759. https://doi.org/10.3390/molecules24040759

Article Metrics

Back to TopTop