Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Phosphorescence performance of La3GaGe5O16:Tb3+ by enhancing the bridging oxygen vacancies

Open Access Open Access

Abstract

La3GaGe5O16:Tb3+ phosphors deliver a long color-tuning persistent luminescence arising from the defect related emission and the 4f - 4f transitions of Tb3+. The persistent luminescence color was tuned by changing the doping Tb3+ concentration. The effect of the different atmospheres on the persistent luminescence of Tb3+-doped La3GaGe5O16 was investigated. Furthermore, we modified the composition around Tb3+ by adjusting the Ge/O content and improved the performance of persistent luminescence of the La3GaGe5O16:Tb3+. The samples synthesized in the poor-oxygen atmosphere most likely exhibited more exceptional persistent phosphorescence by enhancing the oxygen vacancies. To gain insight into the trapping and releasing processes involved in the persistent luminescence, we conducted a series of TL measurements by combining with the first-principles theory calculation. In addition, more investigations were carried out to unravel the nature of traps and also to verify the rationality of the material design.

© 2016 Optical Society of America

1. Introduction

Long persistent luminescence (LPL) is a “self-sustained” luminescence phenomenon whereby luminescence can last for minutes to hours at room temperature after the stoppage of the excitation. For long persistent phosphors (LPPs), the energy stored in suitable traps could be released which was induced by thermal energy available at room temperature. To date, LPPs have received much attention and interests due to their wide application, e.g., such as emergency escape routes and exit signs, optical energy media, thermal sensors [1–6]. Besides the traditional applications, new fields of interest emerged in medical field such as cancer photodynamic therapy, cancer diagnoses and in vivo imaging [7–10]. ZnS:Cu persistent phosphor was first applied for making watch dials in the middle of 1990s [3,11]. In 1996, a new persistent aluminate material (SrAl2O4:Eu2+, Dy3+) was reported by Matsuzawa. et al, which can emit in visible for 10h after ceasing the irradiation and gave us a new benchmark for the performance of persistent luminescence materials [12]. Since then, this new finding boosted the research on the persistent luminescence phosphors and so many efficient persistent luminescence materials were discovered based on stable aluminate and silicate materials, i.e. CaAl2O4:Eu2+, Nd3+ [13], Sr4Al14O25:Eu2+, Dy3+ [14], and Sr2MgSi2O7:Eu2+, Dy3+ [15]. Meanwhile, a large amount of the trivalent rare-earth (R3+) -doped LPPs were also found such as CaTiO3:Pr3+ [16], Y2O2S:Eu3+, Mg2+, Ti4+ [17], and Ca2SnO4:Sm3+ [18]. However, neither the persistent luminescence nor the afterglow times are less than that of Eu2+-doped LPPs. More importantly, the details of the nature and distribution of the traps in R3+-doped LPPs are still a subject of challenge, which are critical to explore and design the desired long persistent phosphors. Based on our team’s works about R3+-doped long persistent phosphors such as Sr3Al2O5Cl2:Pr3+ [19], CdGeO3:Tb3+ [20], La3GaGe5O16:Pr3+ [21], BaZrSi3O9:R3+ (R = Eu, Sm, Dy, Tb and Pr) [22], SrZrO3:Pr3+ [23] etc, a doubt comes to my mind that how to improve the phosphorescence performance of R3+-doped LPPs. A persistent luminescence requires the presence of abundant traps able to intercept free carriers, and to immobilize them for an appropriate time. Thus, how to enhance the lattice defects (oxygen vacancies) or the impurities is key to optimize the phosphorescence performance of R3+-doped LPPs. Obviously, the lattice defects largely depend on the proportion of the starting prepared materials and the synthesis condition, for example, the fabrication atmosphere or the sintering temperature. The chief reason why La3GaGe5O16 compound is selected to be the host is that this host is rich in the bridging oxygen and could be the potentially suitable for R3+-doped long persistent phosphors.

As shown in Fig. 1(a), there are five types of the bridging oxygen (O) based on the crystal structure of La3GaGe5O16, as denoted O1~O5. O1 is located between the tetracoordinated and hexacoordinated germanium (GeO4 and GeO6); O2 lies between the tetracoordinated gallium (GaO4) and the hesacoordinated germanium (GeO6); O3 comes from the joint between the tetracoordinated gallium (GaO4) and pentacoordinated germanium (GeO5); O4 exists between two pentacoordinated germanium (GeO5); O5 originates between the tetracoordinated germanium and pentacoordinated germanium (GeO5). On the other hand, Ge-O and Ga-O units are the basic framework of the structure, so the formation of Ge vacancies (VGe) and Ga vacancies (VGa) are both energetically costly. Meanwhile, the La vacancies (VLa) in this host can hardly occur as well, because La layers are located in a loose and comfortable space. As far as we know, oxygen vacancies, concerning linking (bridging) oxygen, cost the lower energy than the formation of other vacancy defects and thus stand out as a most likely cause of the observed persistent luminescence in Sr2MgSi2O7 [24,25]. If it is true in the case of La3GaGe5O16 phosphor, we expect that the reducing atmosphere will lead to a good performance of La3GaGe5O16 phosphor, since there are various types of bridging oxygen in La3GaGe5O16 host and the oxygen vacancies may be presented in high concentration under high-temperature reducing atmosphere annealing conditions. Therefore, oxygen vacancies are considered as candidates for the lattice defects in La3GaGe5O16 host. In order to gain a theory about the trapping of charge carriers at oxygen vacancies, Freysoldt et al gave us a concrete procedure that how density functional theory (DFT) calculations with periodic boundary conditions are performed on supercells containing an oxygen vacancy [26].

 figure: Fig. 1

Fig. 1 View of the structure of La3GaGe5O16 compound and the different cation are shown, respectively.

Download Full Size | PDF

In our previous works, we presented the photoluminescence and phosphorescence properties of La3GaGe5O16 host and also observed a pink persistent luminescence from La3GaGe5O16:Pr3+ [21]. In this paper, we observed a green long persistent luminescence from La3GaGe5O16:Tb3+ phosphors due to the defect related emission, and the 4f - 4f transitions of Tb3+. We tried to tune the multicolor persistent luminescence by doping Tb3+ concentration in La3GaGe5O16. In order to improve the phosphorescence performance, we prepared the La3GaGe5O16:Tb3+ phosphors under different flowing gas atmospheres and modified the composition around Tb3+ by adjusting the Ge/O content. For better understand the nature of traps involved in trapping and releasing processes, La3GaGe5O16:Tb3+ phosphors were studied by a combination of both experiments and first-principle calculation.

2. Experimental

2.1 Synthesis

The samples were prepared by a two-step traditional solid-state reaction. La2O3(99.99%), Ga2O3 (99.99%), GeO2 (99.99%) and Tb2O3 (99.99%) were used as starting materials. After the raw materials were weighed according to the composition of La3(1-x)GaGe5O16:xTb3+ (x = 0, 0.001, 0.005, 0.01, 0.03, 0.05, 0.07, and 0.09, etc) and La2.99GaGe5(1-x)O16:0.01Tb3+ (−0.05≤x≤0.05), the powders were mixed and milled thoroughly for 1h in an agate mortar and prefired at 900 °C for 8 h in air. After ground again, the mixtures were sintered at 1300 °C for 12 h under different atmospheres such as vacuum, air, 100% N2 and 5%H2/95% N2. After being cooled down to room temperature naturally, the as-prepared powder samples were obtained.

2.2 Measurement

The phase purity of the prepared phosphors was measured by an X-ray diffractometer with Cu Ka radiation (wavelength = 0.15406 nm) at 36 kV tube voltage and 20 mA tube current. Room-temperature photoluminescence excitation, emission spectra and afterglow spectra of all the samples were measured by a Hitachi F-7000 Fluorescence Spectrophotometer equipped with a 150 W xenon lamp as excitation source. The persistent luminescence spectrum was also measured using the spectrophotometer by shutting off the Xe lamp. The decay curves were measured by a GFZF-2A single-photo-counter system. The thermoluminescence (TL) spectrum was measured with a FJ-427A1 thermoluminescence reader and analyser system. For the TL glow curve measurements, the heating rate was 1°C/s and the range of the measurement is from room temperature to 300°C. A delay for 3 min was used between the irradiation and measurement. All measurements were carried out at room temperature except for the TL spectrum. Electron paramagnetic resonance (ESR) spectra were recorded with an X-band spectrometer (Bruker A300) at 100K before and after irradiation by a xenon lamp. The afterglow images were recorded with a classic Reflex digital camera (Nikon camera) with the same exposure time.

The representative XRD pattern of La3GaGe5O16: Tb3+ is shown in Fig. 10 in the Appendix. Our calculations were based on DFT within the generalized gradient approximation (GGA) [26] combined with the projector augment wave (PAW) method as implemented in VASP code [27,28]. The valence configurations of O, Ga, Ge, and La were 2s22p4, 4s24p1, 4s24p2, and 5d16s2, respectively. A plane-wave basis set with an energy cutoff of 400eV was adopted. Considering the size of the cell used (Fig. 11 in the Appendix), a 3 × 2 × 1 monkhorst-Pack k-point mesh was used for integrations over the Brillouin zone. The ionic positions were fully relaxed until the residual force acting on each ion was less than 0.01eV/ Ǻ. The calculated crystal structure was optimized based on the experimental data, as shown in (see Fig. 12 in the Appendix). The optimized lattice parameters are a = 4.827 Å, b = 8.090 Å, c = 15.707 Å, α = 90.76°, β = 94.01°, and γ = 90.01°, respectively, in good agreement with the experimental results. Test calculations show that the energy difference to be within 0.2 eV by using 150- and 450-atom supercells for one sample. Therefore, oxygen vacancies were simulated by removing one O atom from a 150-atom supercell with periodic boundary conditions, corresponding to the defect concentration of 1% for oxygen vacancies.

3. Results and discussion

3.1 Color tuning of persistent luminescence

The photoluminescence properties of La3GaGe5O16:Tb3+ were not discussed in detail, but the excitation and emission spectra were presented (see Fig. 13 in the Appendix). As shown in Fig. 2(a), La3GaGe5O16:Tb3+ gives a combined parts including the host emission and the Tb3+ characteristic emission. Thus, we can tune the persistent luminescence by changing the Tb3+ doping concentration. In our previous work, La3GaGe5O16 is proved to be a new self-activated luminescent host [21]. From Fig. 2, we can see that the host emission decreases with the increasing doping content of Tb3+, indicating that energy transfer from the host to Tb3+ ions occurs but it is inefficient. As the result of the different doping content, the persistent luminescence is observed to vary from cyan to green as shown in Fig. 2(b). We also give the persistent emission spectra of La3(1-x)GaGe5O16:xTb3+ as a function of the afterglow time (see Fig. 14 in the Appendix).

 figure: Fig. 2

Fig. 2 (a) the persistent luminescence of La3(1-x)GaGe5O16:xTb3+ with different doping concentration (x = 0.1%, 0.2%, 0.3%, 0.5%, 1.0% and 1.5%) measured immediately after stopping the light source (excitation slit Δλ = 10 nm, emission slit Δλ = 10 nm); (b) CIE chromaticity coordinates of the Tb3+-doped samples.

Download Full Size | PDF

3.2 Phosphorescence properties

Usually, the thermoluminescence (TL) method is considered as an efficient tool to probe the traps, such as the evaluating for the depth and density of traps in host material [29,30]. Clearly, knowing the depth and density of traps is crucial when trying to understand the mechanism of persistent luminescence, and when developing new afterglow materials. To gain insight into the trapping and releasing processes involved in the persistent phosphor La3GaGe5O16:Tb3+, we conducted a series of TL measurements by varying the delay time after the excitation and the thermal cleaning temperature. The TL curves of a typical sample La3GaGe5O16:0.015Tb3+ at different delay times (t = 0.5, 1, 3, 5, 7 and 10 min) are shown in Fig. 3(a). Obviously, the TL bands have a large shift from lower temperature to higher temperature with the elapse of time. It is caused by the lacking of low-temperature tail because the charge carriers escape from shallower traps much faster than that in deep ones. In addition, it reveals that TL band is composed of a series of traps with continuous distribution and close depths. With the longer time, the depth of predominated TL band becomes deeper and deeper.

 figure: Fig. 3

Fig. 3 (a) TL glow curves of La2.99GaGe5O16:1.0%Tb3+ measured by varying the delay time ; (b) TL glow curves by preheating the La2.99GaGe5O16:1.0%Tb3+ sample at the different thermal cleaning temperature; (c) Initial rise analysis of the corresponding TL glow curves; (d) Trap depth distribution of La2.99GaGe5O16:1.0%Tb3+ sample.

Download Full Size | PDF

To further probe the energy distribution of traps, it is imperative to vary the preheating temperature at which the sample is excited. If a phosphor is excited at a higher temperature with a continuous trap depth distribution, only deeper fractions of the traps are filled. The shallower traps are immediately bleached due to the increased thermal energy available. TL glow curves of La3GaGe5O16:0.015Tb3+ sample were collected after preheated at different preheating temperature in Fig. 3(b). This sample was first irradiated for 1 min at 300K and then heated to a certain temperature (Tstop) to partially clean the occupied traps. From Fig. 3(b), the gradual deepening of the trap depth for increasing Tstop proves the presence of a continuous trap distribution in La3GaGe5O16:Tb3+ phosphor.

An initial rising method, proposed by Van den Eeckhout et al [31,32], is further applied to the measured TL curves to uncover how the trap depth of La3GaGe5O16:Tb3+ phosphor distributes [32]. Briefly, this method for estimating trap depths starts from a assumption that the concentration of trapped electron on the low-temperature side of the TL curves remains relatively constant; only a tiny fraction of the charge carriers can escaped under the small amount of thermal energy available. Thus, the TL intensity (I(t)) can be approximately expressed as:

I(t)=Cexp(EkT)
where C is the constant including the frequency factor s, and k is the Boltzmann constant [33,34]. By plotting the TL curves as ln(I) vs 1000/T, i.e. in an Arrhenius plot, the shallowest trap depth can be readily estimated by the slope of a fitted straight section at the low-temperature side. The transformed Arrhenius plots for each curves in Fig. 3(b) were presented in Fig. 3(c). Obviously, all the plots give a straight section in the low-temperature side, indicating that our assumption is valid for the initial rising analysis. From the fitted slope, the depth of the shallowest trap can be obtained after different thermal cleaning temperature. As shown in Fig. 3(c), the estimated trap depths gradually deepens from 0.690 eV to 0.750eV, demonstrating the continuous distribution of the trap depth, thus. In Fig. 3(d), the trap density between two adjacent depths can be roughly calculated from the difference between the integrated intensities of the corresponding two TL curves. In general, the first and second order kinetics are only valid in ideal situations, i.e. when the retrapping probability is negligible or when it is equal to the recombination probability. Thus, the general order kinetics might play a primary role in the trapping and releasing processes involved in the persistent phosphor La3GaGe5O16:Tb3+. The depth of TL band is situated close to the ideal trap depth (0.6-0.8 eV) for the release at room temperature of the energy storage [35,36]. Thus, La3GaGe5O16:Tb3+ phosphor can present a long green persistent luminescence.

Figure 4 gives the TL glow curves of La2.99GaGe5O16:0.01Tb3+ synthesized under different synthesis atmosphere (5%H2 + 95%N2, 100% N2, vacuum and the air). For comparison, TL glow curve of La3GaGe5O16 host was also given. As a rule, each TL peak stands for a kind of trapping center and the relative trap density or trapping capacity is roughly proportional to the integral TL intensity [33–35].Compared the TL curves between La2.99GaGe5O16:0.01Tb3+ and La3GaGe5O16 host (which were prepared in air), we infer that intrinsic defects might play main role in the persistent luminescence because TL curve of Tb3+-doped La3GaGe5O16 is similar to that of La3GaGe5O16 host. For further prove our speculation, we made a comparison about the TL curves of La2.99GaGe5O16:0.01Tb3+ synthesized in a vacuum and air, and found the persistent performance prepared in the vacuum is much superior to that in air. Thus, oxygen vacancies might be considered as candidates for the lattice defects in La3GaGe5O16 host under high-temperature annealing conditions. If it is true in the case of La3GaGe5O16 phosphor, a better persistent performance of La3GaGe5O16:Tb3+ phosphor can be expected in a reducing atmosphere. Therefore, we tried to synthesize the La3GaGe5O16:Tb3+ phosphors under the reducing flowing gas atmosphere (5%H2/95%N2) to improve the phosphorescence performance. From Fig. 4, we can see the integral TL intensity of La3GaGe5O16:Tb3+ synthesized in the reduction atmosphere (5%H2 + 95%N2), is the largest compared with other two samples, suggesting this sample presents the best performance in the persistent luminescence. The reason for this is that the traps might be closely related with the oxygen vacancies (VO). In other word, VO can be developed helpfully when La3GaGe5O16:Tb3+ was prepared in the reduction atmosphere (5%H2 + 95%N2), however VO is suppressed partially when in an oxygenated atmosphere (the air). This is why La3GaGe5O16:Tb3+ samples synthesized at different atmosphere give the different behaviors in persistent luminescence. Thus, a conclusion might be made that the trapping center probably originates from the intrinsic defects not the extrinsic defects. Both the position and the shape of TL curves are the same besides the intensity of the TL curves. The role of the doping Tb3+ ions might be dual, as they can modify the intrinsic defects or enrich the instinct defects in the host and sever as luminescent centers as well. So, the integral TL intensity of La3GaGe5O16:Tb3+ samples are higher than that of La3GaGe5O16 host.

 figure: Fig. 4

Fig. 4 TL glow curves of La3GaGe5O16:Tb3+ synthesized at different atmosphere (a:5%H2 + 95%N2, b: 100% N2, c: the air and d: vacuum).

Download Full Size | PDF

By adjusting the Ge/O content ratio slightly, we aim to prolong the afterglow time and intensity of the La3GaGe5O16:Tb3+ by modifying composition around Tb3+, since the intrinsic traps of host probably play an important role in the persistent luminescence. Thus, we prepared a series of Tb3+-doped La3GaGe5+xO16+4x (−0.05≤x≤0.05) samples in the reduction atmosphere (5%H2 + 95%N2). The XRD patterns of La3GaGe5+xO16+4x (−0.05≤x≤0.05) samples were presented (see Fig. 15 in the Appendix), proving that small Ge/O content deficiency can still maintain the crystal structure intact. The persistent luminescence and afterglow times of La3GaGe5+xO16+4x:0.01Tb3+ (−0.05≤x≤0.05) are presented in Fig. 5(a) and 5(b). From Fig. 5(a) and 5(b), one can see that Ge/O content deficiency benefits to improve the persistent luminescence and afterglow time, whereas these samples, a composition of Ge/O content excess, exhibit poorer afterglow properties. This one aspect demonstrates the persistent luminescence of La3GaGe5O16:Tb3+ is closely related to the intrinsic defects in host. Thus, we can make a assumption that composition with slight Ge/O content deficiency is of importance for La3GaGe5O16:Tb3+ to enrich the traps concentration. In order to further prove our proposed assumption, the TL glow curves of La2.99GaGe5+xO16+4x:0.01Tb3+ (−0.05≤x≤0.05) were measured. As shown in Fig. 5(c), the integral intensity of TL glow curves gives a trend of increasing as the value of x decreases, indicating that composition of La3GaGe5+xO16+4x:0.01Tb3+ samples with slight Ge/O content deficiency contributes to enriching the traps concentration. Furthermore, new TL bands peaked in the higher-temperature region were detected when the Ge/O content is deficient. However, this is only the first insight. A detailed analysis of the TL curves should adopt the deconvolution method based on the Gaussian equation. Thus, a good agreement between the experimental and calculated glow curves was obtained by using the above method. As shown in Fig. 5(d), the TL glow curve of La2.99GaGe4.95O15.2:0.01Tb3+ consists of three bands located at 345K, 393K, and 420K. By comparing the TL curves of La2.99GaGe5+xO16+4x:0.01Tb3+ (−0.05≤x≤0.05), the TL band located at 345K should be attributed to the trapping center of oxygen vacancies. The other TL bands located at 393K, and 420K are so weak that they can make a little contribution for the persistent luminescence; and these two new TL bands might be caused by the new oxygen vacancies because there are various types of bridging oxygen in La3GaGe5O16 host, as illustrated in Fig. 1.

 figure: Fig. 5

Fig. 5 (a) The emission images of La3GaGe5+xO16+4x:0.01Tb3+ (−0.05≤x≤0.05) recorded by a classic Reflex digital camera with the same exposure time varying with the different afterglow time (t = 0.5, 3, 5, 10 and 15 min); (b) the afterglow curves of La3GaGe5+xO16+4x:0.01Tb3+ (−0.05≤x≤0.05) samples; (c) TL glow curves of La3GaGe5+xO16+4x:0.01Tb3+ (−0.05≤x≤0.05) sample (Irradiation time: 2 min; delay time: 3 min and heating rate:1°C/s); (d) deconvolution of the TL glow curve of the La3GaGe4.95O15.20:0.01Tb3+ sample.

Download Full Size | PDF

3.3 Defects and traps

As far as we know, effects, either intrinsic or extrinsic defects or both of them, play important roles in energy storage of persistent luminescence. Therefore, understanding of defect properties is indispensable for further development and applications of certain potential persistent phosphors. In order to distinguish between the extrinsic and intrinsic defects, further investigation of trap types should be the key to unlock the puzzle. Figure 6 gives the ESR spectra of La3GaGe5O16 host and La3GaGe5O16:0.01Tb3+ sample before and after irradiation measured at 100K. A sharpening signal at g = 3.9176, g = 3.6985 (Trap І) and a weak one at g = 1.9588 (Trap П) were detected after irradiation, in contrast to a lack of signal before irradiation in the inset of Fig. 6.The width and constriction of the signals remain practically unchanged between La3GaGe5O16 host and La3GaGe5O16:0.01Tb3+ sample. This implies that nothing is changed in the trap types and also means the doping Tb3+ ions don’t induce the extrinsic defects or traps. These two signals are attributed to oxygen vacancies [37–39]. Except for the signals of oxygen vacancies, others signals were not observed in Fig. 6. Thus, it is proved that oxygen vacancies are considered as candidates for the lattice defects in La3GaGe5O16 host. The intensity of the EPR signal varies obviously between them, confirming the assumption we proposed in above section that the doping Tb3+ ions can modify the intrinsic defects or enrich the instinct defects in the host. Thus, the role of the extrinsic defects can be overlooked in persistent luminescence, and this afterglow phenomenon probably originates from the oxygen vacancies. However, which kind of of the bridging oxygen vacancies (O) play the key role in this persistent phosphor? The answer comes upon total-energy calculations (see the following section).

 figure: Fig. 6

Fig. 6 The trap types: EPR spectra of La3GaGe5O16 host and La3GaGe5O16:0.015Tb3+ sample recorded at 100K after irradiation; inset show the EPR spectra of La3GaGe5O16 host before irradiation.

Download Full Size | PDF

3.4 Electronic structures

Understanding the electronic structures of the host material and its lattice defects is essential to control the carrier trapping and detrapping process in the persistent phosphors [33–35]. Before investigating the lattice defects in La3GaGe5O16, we first studied the electronic structure of perfect La3GaGe5O16 crystal. Calculated band structure and density of states of La3GaGe5O16 are shown in Fig. 7(a) and 7(b). We see that La3GaGe5O16 is an insulator with an indirect gap: the valence band maximum (VBM) composed of O 2p orbitals is at the V point and the conduction band minimum (CBM) characteristic of Ge 4s orbitals is at the Γ point; the band-gap energy was calculated to be 3.2002 eV. In order to compare this with the experimental band gap values, we calculated the measured value of 4.352 eV for La3GaGe5O16 host based on the diffuse reflection spectrum (see Fig. 16 in the Appendix). The obtained value for La3GaGe5O16 is 4.352 eV. This value is actually the optical gap rather than the electronic band gap. Considering the uncertainty of the experiment, the actual electronic band gap is expected to be less than 4.352 eV. Compared with the experimental value, our calculated GGA Kohn-Sham band gap (3.2002 eV) is around 2/3 of the experimental values, which falls into the error range of GGA calculations. The band structure is composed of six parts (P1-P6), P1, P4 and P5 is considered as O 2s and Ga 3d, La 5p and Ge 3d semicore states, respectively, showing localized (nonbonding) character, whereas P5 and P6 consist of the hybridized Ge 4s4p, La 5p5d, Ga 4s4p and O 2p states. The hybridization separates the bonding states below the VBM and the antibonding states above the CBM. The bonding states of Ge-O are located at the lowest energy range relative to that of Ga-O and La-O, indicating that the Ge-O constituent has a largest contribution to the cohesive energy of La3GaGe5O16.

 figure: Fig. 7

Fig. 7 (a) Computed band structure, density of states (DOS); (b) momentum projected local density of states (PDOS) for La3GaGe5O16 (Zero of energy is set at the Fermi level.

Download Full Size | PDF

3.5 Energy band structure of VO

As well defined by Freysoldt et al [26], the defect formation energy of a neutral defect Ef[defect] depends on the atomic chemical potential μi: Ef[defect] = Etot[defect] − Etot[host] + niμi. Here, Etot[defect] is the total energy of a supercell containing the defect, Etot[host] is the total energy of the defect-free supercell, ni indicates the number of atoms of type i that have been added to (ni<0) or removed from (ni>0) the supercell when the defect is created. To compare the formation of the same kind neutral defects, it suffices to know the difference between Etot[defect] and Etot[host]. For this reason, we can discuss the stability of O1~O5 in terms of their total energies. However, since the values of the total energy lacks physical meaning, we have replaced the absolute total energies of O1~O5 in Table 1 with the relative values with respect to the perfect supercell (i.e. Etot[VO] − Etot[La3GaGe5O16]) (these supercells with different relaxed vacancy defect are presented (see Fig. 17 in the Appendix).

Tables Icon

Table 1. The calculated total energy difference between the supercell containing one relaxed oxygen vacancy in La3GaGe5O16 and the perfect supercell.

One can see that the O4 has the lowest ∆E, suggesting that as the intrinsic traps, O4 is the most possible formed and one of the potential sources that explain the persistent luminescence in La3GaGe5O16:Tb3+. But by contrasting the O3 and O4, the energy differences between O3 and O4 is only 0.01eV. Except for O4, therefore, O3 should contribute to the persistent luminescence of La3GaGe5O16:Tb3+ phosphor as well. In fact, even for O1, O5, the differences with O4 are only about 0.15 eV and 0.17 eV, respectively. At the same time, it can explain why the trap distribution in La3GaGe5O16:Tb3+ phosphor is continuous ranging from 0.69eV to 0.75eV in theory.

Our focus in this section is to study the defect states associated with oxygen vacancies. According to the explanations by Freysoldt et al [26], the Kohn-Sham levels that result from a band structure calculation for a defect cannot directly be identified with the defect levels that are of experimental relevance, because the experimental levels involve transition between different charge states of the defect. In view that the calculations of defect in different charge states are of a demanding work, we just discuss the relevance of the Kohn-Sham levels of the defect states with the optical levels qualitatively.

Band structure calculation in Fig. 8 shows that the neutral O vacancy introduces an s-like deep donor level (a1) close to the VBM. Its two unpaired electrons can be easily ionized. So the neutral O vacancy (VO0) is not stable, and it can be ionized in 2 + charge state (VO2+). Besides, the neutral O vacancy can also introduce another dispersive empty band that is continuous with the conduction band within 1eV below the CBM. Such empty states distributed a wide energy range are considered to contribute to prolonging the afterglow decay time [40,41]. We then study the relevance of the energy distribution of these empty states with the trap depth measured from the thermoluminescence glow curves. The measured trap depth is in the range of 0.69~0.75 eV for La2.99GaGe5O16:0.01Tb3+, which falls into the ideal trap depth (0.6~0.8 eV) for the electron detrapping at room temperature. Nevertheless, the trap depth obtained from the Kohn-Sham levels is ~1eV, which appears somewhat deeper than the experimental value. This difference reflects that the Kohn-Sham levels should be corrected to compare with experiments. We illustrate the following simplified picture of the correction [26,42], as shown in Fig. 9. Under UV light excitation, the two electrons in the a1 state of VO0 are trapped at the VO2+ center. The equilibrium configuration of the VO2+ state is Egεa1 higher than the equilibrium configuration of VO0 − 2e, where εa1 is the Kohn-Sham level of the a1 state. Electrons at the VO2+ center can then recombine with the holes on the VO0 center, leading to emission of a photon with energy EPL. However, during this emission process, the atomic configuration remains fixed in the 2 + charge state. The difference between the energy of this configuration and that of the equilibrium configuration of VO0 is the relaxation energy Erel (Frank-Condon relaxation). From Fig. 9, it is clear that EPL = Egεa1Erel. Since the optical level of the a1 state εa1opt is defined as the energy difference between the band gap and the PL line, we find that εa1opt = EgEPL = εa1 + Erel, i.e. εa1 = εa1optErel. This means that the Kohn-Sham level of the a1 state should be smaller than the optical measurements by an amount of Erel. When such correction in terms of Erel is applied to the a1 level, the other dispersive empty defect band levels will be shifted upward accordingly. Therefore, the resulting trap depth obtained from the Kohn-Sham levels should be shallower than the former ~1eV (below the CBM), and thus more close to the experimental value.

 figure: Fig. 8

Fig. 8 Computed band structure of La3GaGe5O16 host and La3GaGe5O16 + VO4 (a: the Fermi level is set to 0 eV; b: the hole energy level above the VB is set to 0 eV).

Download Full Size | PDF

 figure: Fig. 9

Fig. 9 A simplified picture of Frank-Condon shift in the case of VO.

Download Full Size | PDF

4. Conclusion

In a word, we tuned the persistent luminescence color by changing the Tb3+ concentration, and the persistent color can be tuned from cyan to green. The afterglow time and intensity of La3GaGe5O16:Tb3+ can be prolonged and enhanced in a poor-oxygen atmosphere (5%H2 + 95%N2). Furthermore, the performance of persistent luminescence La3GaGe5O16:Tb3+ can be improved by adjusting the Ge/O content to modify composition around Tb3+ ions. In order to reveal the nature of persistent luminescence, the electronic structures of La3GaGe5O16 host and VO4 were calculated based on the first-principles, indicating that the suitable impurity levels from VO4 cannot only trap carriers but release the trapped carriers through activation. This study gives us some hints that the materials, like La3GaGe5O16, are rich in the bridging oxygen and could be the potentially suitable for RE3+-doped long persistent phosphors; the samples synthesized in the poor-oxygen atmosphere probably exhibit more excellent long persistent phosphorescence. At last, we hope insights obtained from this paper can offer us a guiding principle for the selection of the host and the luminescent center for the design of RE3+-doped persistent phosphors.

Appendix

 figure: Fig. 10

Fig. 10 The XRD patterns of La3GaGe5O16: xTb3+ (x=0, 0.1%, 0.3%, 0.5% and 1.0%).

Download Full Size | PDF

 figure: Fig. 11

Fig. 11 Unit cell of La3GaGe5O16 ( centrosymmetric point: Г(0, 0, 0) Z(0, 0, 0.5) Y or F (0, 0.5, 0) X or B (0.5, 0, 0) V(0.5, 0.5, 0) U(0.5, 0, 0.5) T or Q (0, 0.5, 0.5) R(0.5, 0.5, 0.5).

Download Full Size | PDF

 figure: Fig. 12

Fig. 12 Experimental (a) and calculated (b) crystal structure of La3GaGe5O16.

Download Full Size | PDF

 figure: Fig. 13

Fig. 13 Excitation and emission spectra of La2.99GaGe5O16: 0.01Tb3+ sample.

Download Full Size | PDF

 figure: Fig. 14

Fig. 14 The persistent emission spectra of La3(1-x)GaGe5O16:xTb3+ as a function of the afterglow time.

Download Full Size | PDF

 figure: Fig. 15

Fig. 15 XRD patterns of La3GaGe5+xO16+4x:0.01Tb3+ (-0.05≤x≤0.05) samples.

Download Full Size | PDF

 figure: Fig. 16

Fig. 16 UV-visible diffuse reflection spectrum of undoped La3GaGe5O16 ; the inset gives the relationship between (F(R)hν)2and hv.

Download Full Size | PDF

 figure: Fig. 17

Fig. 17 (a) the perfect compound La3GaGe5O16;(b-f) presents the different crystal structures after relaxation with different oxygen vacancies such as VO1, VO2, VO3, VO4 and VO5.

Download Full Size | PDF

Acknowledgments

This work are supported by the National Natural Science Foundation of China (No. 21471038).

References and links

1. Z. Pan, Y. Y. Lu, and F. Liu, “Sunlight-activated long-persistent luminescence in the near-infrared from Cr(3+)-doped zinc gallogermanates,” Nat. Mater. 11(1), 58–63 (2011). [CrossRef]   [PubMed]  

2. J. Liu, R. Wu, N. Li, X. Zhang, Q. Zhan, and S. He, “Deep, high contrast microscopic cell imaging using three-photon luminescence of β-(NaYF4:Er(3+)/NaYF4) nanoprobe excited by 1480-nm CW laser of only 1.5-mW,” Biomed. Opt. Express 6(5), 1857–1866 (2015). [CrossRef]   [PubMed]  

3. P. Wang, X. Xu, D. Zhou, X. Yu, and J. Qiu, “Sunlight activated long-lasting luminescence from Ba5Si8O21: Eu(2+),Dy(3+) phosphor,” Inorg. Chem. 54(4), 1690–1697 (2015). [CrossRef]   [PubMed]  

4. J. Ueda, P. Dorenbos, A. J. Bos, K. Kuroishi, and S. Tanabe, “Control of electron transfer between Ce3+ and Cr3+ in the Y3Al5-xGaxO12 host via conduction band engineering,” J. Mater. Chem. C Mater. Opt. Electron. Devices 3(22), 5642–5651 (2015). [CrossRef]  

5. X. Shi, R. Dou, T. Ma, W. Liu, X. Lu, K. J. Shea, Y. Song, and L. Jiang, “Bioinspired lotus-like self-illuminous coating,” ACS Appl. Mater. Interfaces 7(33), 18424–18428 (2015). [CrossRef]   [PubMed]  

6. R. Weissleder and M. J. Pittet, “Imaging in the era of molecular oncology,” Nature 452(7187), 580–589 (2008). [CrossRef]   [PubMed]  

7. A. Hellebust and R. Richards-Kortum, “Advances in molecular imaging: targeted optical contrast agents for cancer diagnostics,” Nanomedicine (Lond.) 7(3), 429–445 (2012). [CrossRef]   [PubMed]  

8. M. Nyk, R. Kumar, T. Y. Ohulchanskyy, E. J. Bergey, and P. N. Prasad, “High contrast in vitro and in vivo photoluminescence bioimaging using near infrared to near infrared up-conversion in Tm3+ and Yb3+ doped fluoride nanophosphors,” Nano Lett. 8(11), 3834–3838 (2008). [CrossRef]   [PubMed]  

9. J. Zhou, Z. Liu, and F. Li, “Upconversion nanophosphors for small-animal imaging,” Chem. Soc. Rev. 41(3), 1323–1349 (2012). [CrossRef]   [PubMed]  

10. S. Gai, C. Li, P. Yang, and J. Lin, “Recent progress in rare earth micro/nanocrystals: soft chemical synthesis, luminescent properties, and biomedical applications,” Chem. Rev. 114(4), 2343–2389 (2014). [CrossRef]   [PubMed]  

11. R. Bowers and N. T. Melamed, “Luminescent centers in ZnS: Cu: Cl phosphors,” Phys. Rev. 99(6), 1781–1787 (1955). [CrossRef]  

12. T. Matsuzawa, Y. Aoki, N. Takeuchi, and Y. Murayama, “A new long phosphorescent phosphor with high brightness, SrAl2O4:Eu2+, Dy3+,” J. Electrochem. Soc. 143(8), 2670–2673 (1996). [CrossRef]  

13. B. Y. Qu, B. Zhang, L. Wang, R. L. Zhou, and X. C. Zeng, “Mechanistic study of the persistent luminescence of CaAl2O4: Eu, Nd,” Chem. Mater. 27(6), 2195–2202 (2015). [CrossRef]  

14. Y. Lin, Z. Tang, and Z. Zhang, “Preparation of long-afterglow Sr4Al14O25-based luminescent material and its optical properties,” Mater. Lett. 51(1), 14–18 (2011). [CrossRef]  

15. K. Van den Eeckhout, P. F. Smet, and D. Poelman, “Persistent luminescence in Eu2+-doped compounds: a review,” Materials (Basel) 3(4), 2536–2566 (2010). [CrossRef]  

16. X. Zhang, J. Zhang, Z. Nie, M. Wang, X. Ren, and X. J. Wang, “Enhanced red phosphorescence in nanosized CaTiO3:Pr3+ phosphors,” Appl. Phys. Lett. 90(15), 151911 (2007). [CrossRef]  

17. Y. Kawahara, V. Petrykin, T. Ichihara, N. Kijima, and M. Kakihana, “Synthesis of high-brightness sub-micrometer Y2O2S red phosphor powders by complex homogeneous precipitation method,” Chem. Mater. 18(26), 6303–6307 (2006). [CrossRef]  

18. Z. Ju, R. Wei, J. Zheng, X. Gao, S. Zhang, and W. Liu, “Synthesis and phosphorescence mechanism of a reddish orange emissive long afterglow phosphor Sm3+-doped Ca2SnO4,” Appl. Phys. Lett. 98(12), 121906 (2011). [CrossRef]  

19. R. Chen, Y. H. Hu, H. Y. Wu, H. Y. Jin, Z. F. Mou, and S. A. Zhang, “Luminescent properties of blue green Sr3Al2O5Cl2:Pr3+ and orange red Sr3Al2O5Cl2:Eu2+, Pr3+ afterglow phosphors,” Radiat. Meas. 80, 38–45 (2015). [CrossRef]  

20. S. A. Zhang, Y. H. Hu, L. Chen, G. F. Ju, Z. H. Wang, and J. Lin, “Luminescent properties of a green emitting persistent phosphor CdGeO3:Tb3+,” Opt. Mater. 47, 203–210 (2015). [CrossRef]  

21. S. A. Zhang, Y. H. Hu, L. Chen, G. F. Ju, T. Wang, and Z. Wang, “Luminescence properties of the pink emitting persistent phosphor Pr3+-doped La3GaGe5O16,” RSC Advances 5(47), 37172–37179 (2015). [CrossRef]  

22. Y. H. Jin, Y. H. Hu, L. Chen, and X. J. Wang, “Luminescence properties of long persistent phosphors BaZrSi3O9: R3+ (R = Eu, Sm, Dy, Tb and Pr) based on host sensitization,” Opt. Mater. 36(11), 1814–1818 (2014). [CrossRef]  

23. Y. H. Jin, Y. H. Hu, L. Chen, X. J. Wang, G. F. Ju, and Z. F. Mou, “Luminescence Properties of Dual-Emission (UV/Visible) Long Afterglow Phosphor SrZrO3:Pr3+,” J. Am. Ceram. Soc. 96(12), 3821–3827 (2013). [CrossRef]  

24. P. Dorenbos, “Mechanism of persistent luminescence in Sr2MgSi2O7: Eu2+, Dy3+,” Phys. Status Solidi, B Basic Res. 242(1), R7–R9 (2005). [CrossRef]  

25. A. A. Setlur, A. M. Srivastava, H. L. Pham, M. E. Hannah, and U. Happek, “Charge creation, trapping, and long phosphorescence in Sr2MgSi2O7:Eu2+, RE3+,” J. Appl. Phys. 103(5), 053513 (2008). [CrossRef]  

26. C. Freysoldt, B. Grabowski, T. Hickel, J. Neugebauer, G. Kresse, A. Janotti, and C. G. Van de Walle, “First-principles calculations for point defects in solids,” Rev. Mod. Phys. 86(1), 253–304 (2014). [CrossRef]  

27. G. Kresse and J. Furthmüller, “Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set,” Comput. Mater. Sci. 6(1), 15–50 (1996). [CrossRef]  

28. G. Kresse and J. Hafner, “Norm-conserving and ultrasoft pseudopotentials for first-row and transition elements,” J. Phys. Condens. Matter 6(40), 8245–8257 (1994). [CrossRef]  

29. J. Hölsä, M. Lastusaari, M. Maryško, and M. Tukia, “A few remarks on the simulation and use of crystal field energy level schemes of the rare earth ions,” J. Solid State Chem. 178(2), 435–440 (2005). [CrossRef]  

30. P. Dorenbos, ““Systematic behaviour in trivalent lanthanide charge transfer energies,” J. Phys-Condense Mat. 15(49), 8417 (2003).

31. H. Lin, J. Xu, Q. Huang, B. Wang, H. Chen, Z. Lin, and Y. Wang, “Bandgap Tailoring via Si Doping in Inverse-Garnet Mg3Y2Ge3O12:Ce3+ Persistent Phosphor Potentially Applicable in AC-LED,” ACS Appl. Mater. Inter. 7(39), 21835–21843 (1969). [CrossRef]  

32. K. Van den Eeckhout, A. J. Bos, D. Poelman, and P. F. Smet, “Revealing trap depth distributions in persistent phosphors,” Phys. Rev. B 87(4), 045126 (2013). [CrossRef]  

33. S. W. S. McKeever, Thermoluminescence of Solids, Chapter 3 (Cambridge University Press 1985), pp 64–115.

34. A. J. J. Bos, “Theory of Thermoluminescence,” Radiat. Meas. 41, S45–S56 (2006). [CrossRef]  

35. L. C. Rodrigues, J. Hölsä, M. Lastusaari, M. C. Felinto, and H. F. Brito, “Defect to R3+ energy transfer: colour tuning of persistent luminescence in CdSiO3,” J. Mater. Chem. C Mater. Opt. Electron. Devices 2(9), 1612–1618 (2014). [CrossRef]  

36. P. Dorenbos, “Lanthanide charge transfer energies and related luminescence, charge carrier trapping, and redox phenomena,” J. Alloys Compd. 488(2), 568–573 (2009). [CrossRef]  

37. Y. H. Lee and J. W. Corbett, “EPR studies of defects in electron-irradiated silicon: A triplet state of vacancy-oxygen complexes,” Phys. Rev. B 13(6), 2653–2666 (1976). [CrossRef]  

38. C. Drouilly, J. M. Krafft, F. Averseng, S. Casale, D. Bazer-Bachi, C. Chizallet, and G. Costentin, “ZnO oxygen vacancies formation and filling followed by in situ photoluminescence and in situ EPR,” J. Phys. Chem. C 116(40), 21297–21307 (2012). [CrossRef]  

39. L. A. Kappers, O. R. Gilliam, S. M. Evans, L. E. Halliburton, and N. C. Giles, “EPR and optical study of oxygen and zinc vacancies in electron-irradiated ZnO,” Nucl. Instrum. Meth. B 266(12-13), 2953–2957 (2008). [CrossRef]  

40. H. Duan, Y. Z. Dong, Y. Huang, Y. H. Hu, and X. S. Chen, “First-principles study of intrinsic vacancy defects in Sr2MgSi2O7 phosphorescent host material,” J. Phys. D Appl. Phys. 49(2), 025304 (2016). [CrossRef]  

41. H. Duan, Y. Z. Dong, Y. Huang, Y. H. Hu, and X. S. Chen, “The important role of oxygen vacancies in Sr2MgSi2O7 phosphor,” Phys. Lett. A 380(9–10), 1056–1062 (2016). [CrossRef]  

42. C. G. Van de Walle and J. Neugebauer, “First-principles calculations for defects and impurities: Applications to III-nitrides,” J. Appl. Phys. 95(8), 3851–3879 (2004). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (17)

Fig. 1
Fig. 1 View of the structure of La3GaGe5O16 compound and the different cation are shown, respectively.
Fig. 2
Fig. 2 (a) the persistent luminescence of La3(1-x)GaGe5O16:xTb3+ with different doping concentration (x = 0.1%, 0.2%, 0.3%, 0.5%, 1.0% and 1.5%) measured immediately after stopping the light source (excitation slit Δλ = 10 nm, emission slit Δλ = 10 nm); (b) CIE chromaticity coordinates of the Tb3+-doped samples.
Fig. 3
Fig. 3 (a) TL glow curves of La2.99GaGe5O16:1.0%Tb3+ measured by varying the delay time ; (b) TL glow curves by preheating the La2.99GaGe5O16:1.0%Tb3+ sample at the different thermal cleaning temperature; (c) Initial rise analysis of the corresponding TL glow curves; (d) Trap depth distribution of La2.99GaGe5O16:1.0%Tb3+ sample.
Fig. 4
Fig. 4 TL glow curves of La3GaGe5O16:Tb3+ synthesized at different atmosphere (a:5%H2 + 95%N2, b: 100% N2, c: the air and d: vacuum).
Fig. 5
Fig. 5 (a) The emission images of La3GaGe5+xO16+4x:0.01Tb3+ (−0.05≤x≤0.05) recorded by a classic Reflex digital camera with the same exposure time varying with the different afterglow time (t = 0.5, 3, 5, 10 and 15 min); (b) the afterglow curves of La3GaGe5+xO16+4x:0.01Tb3+ (−0.05≤x≤0.05) samples; (c) TL glow curves of La3GaGe5+xO16+4x:0.01Tb3+ (−0.05≤x≤0.05) sample (Irradiation time: 2 min; delay time: 3 min and heating rate:1°C/s); (d) deconvolution of the TL glow curve of the La3GaGe4.95O15.20:0.01Tb3+ sample.
Fig. 6
Fig. 6 The trap types: EPR spectra of La3GaGe5O16 host and La3GaGe5O16:0.015Tb3+ sample recorded at 100K after irradiation; inset show the EPR spectra of La3GaGe5O16 host before irradiation.
Fig. 7
Fig. 7 (a) Computed band structure, density of states (DOS); (b) momentum projected local density of states (PDOS) for La3GaGe5O16 (Zero of energy is set at the Fermi level.
Fig. 8
Fig. 8 Computed band structure of La3GaGe5O16 host and La3GaGe5O16 + VO4 (a: the Fermi level is set to 0 eV; b: the hole energy level above the VB is set to 0 eV).
Fig. 9
Fig. 9 A simplified picture of Frank-Condon shift in the case of VO.
Fig. 10
Fig. 10 The XRD patterns of La3GaGe5O16: xTb3+ (x=0, 0.1%, 0.3%, 0.5% and 1.0%).
Fig. 11
Fig. 11 Unit cell of La3GaGe5O16 ( centrosymmetric point: Г(0, 0, 0) Z(0, 0, 0.5) Y or F (0, 0.5, 0) X or B (0.5, 0, 0) V(0.5, 0.5, 0) U(0.5, 0, 0.5) T or Q (0, 0.5, 0.5) R(0.5, 0.5, 0.5).
Fig. 12
Fig. 12 Experimental (a) and calculated (b) crystal structure of La3GaGe5O16.
Fig. 13
Fig. 13 Excitation and emission spectra of La2.99GaGe5O16: 0.01Tb3+ sample.
Fig. 14
Fig. 14 The persistent emission spectra of La3(1-x)GaGe5O16:xTb3+ as a function of the afterglow time.
Fig. 15
Fig. 15 XRD patterns of La3GaGe5+xO16+4x:0.01Tb3+ (-0.05≤x≤0.05) samples.
Fig. 16
Fig. 16 UV-visible diffuse reflection spectrum of undoped La3GaGe5O16 ; the inset gives the relationship between (F( R )hν) 2 and hv.
Fig. 17
Fig. 17 (a) the perfect compound La3GaGe5O16;(b-f) presents the different crystal structures after relaxation with different oxygen vacancies such as VO1, VO2, VO3, VO4 and VO5.

Tables (1)

Tables Icon

Table 1 The calculated total energy difference between the supercell containing one relaxed oxygen vacancy in La3GaGe5O16 and the perfect supercell.

Equations (1)

Equations on this page are rendered with MathJax. Learn more.

I(t)=Cexp( E kT )
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.