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

Excitation powder dependent optical temperature behavior of Er3+ doped transparent Sr0.69La0.31F2.31 glass ceramics

Open Access Open Access

Abstract

The knowledge of the pump power for which the population of thermally coupled energy levels (TCL) changes with power increase is of valuable importance for optical temperature sensors. In this paper, novel Er3+ doped transparent Sr0.69La0.31F2.31 glass ceramics was fabricated successfully, and its structure is studied by XRD, TEM and HRTEM analyses. The 2H11/2/4S3/2, 4F9/2(1)/4F9/2(2), and 4I9/2(1)/4I9/2(2) levels of Er3+ are proved as TCL by analyzing the temperature dependent fluorescence intensity ratios. The spectrum split, thermal quenching ratio, population stability, and temperature sensitivity from three TCL are observed to be dependent on the pump power. A new fitting method has been developed to establish the relation between fluorescence intensity ratios and temperature. It is found that the combined use of 2H11/2/4S3/2 and 4F9/2(1)/4F9/2(2) as thermally coupled energy levels will get a more precise temperature reading from 62.7 K to 800 K with the help of low excitation power at 66.8 mW/mm2.

© 2016 Optical Society of America

1. Introduction

In recent years, Er3+ doped phosphors were widely explored for applications in lasers, three-dimensional displays, optical temperature sensors, biological imaging probes, and solar cells [1–6], due to their intense visible up-conversion emissions induced by infrared photon excitation. Visible up-conversion emissions of Er3+ ions originated from transitions from 2H11/2, 4S3/2, and 4F9/2 excited states to 4I15/2 ground state after sequential absorption of infrared photons or two-step energy transfers from Yb3+ to Er3+ ions [7–10]. Fischer et. al. reported that Er3+ doped phosphors could have intense emissions at high excitation power where they are saturated [11]. Moreover, some nonlinear properties were also observed in Er3+ up-conversion process at high pumping power of lasers. Chen et.al. reported the “s”-shape power dependence for the green and red emissions in Gd2O3:Er3+ nanocrystals under 970 nm excitation [12]. Instead of three-photon process, the linear dependence of intensities of green and red emissions on the excitation power density was observed by Chen et.al. in LiYF4:Er3+ under 1490 nm excitation [13]. Qu et. al. reported that the green and red emissions of LaF3:Yb3+, Er3+ nanocrystals showed a typical two-photon upconversion process under lower power density of 980 nm laser, while they exhibited as the hysteresis loops when the pump power density turns to higher region [14]. Tian et. al. observed that the green emission of Yb3+-Er3+ co-doped Gd6WO12 nanoparticles was a two-photon process when the excitation power density of 980 nm laser was low, while a three-photon process occurred when the excitation power density was high [15]. These nonlinear changes at high excitation power, owing to the competition between the linear decay and the depletion of the intermediate excited states [16].

It was found that the intensity of green emission from 2H11/2 level increased with temperature increase at fixed excitation powder, while the intensity of green emission from 4S3/2 level decreased [17,18], due to the thermal population and depopulation at high temperature [19]. The fluorescence intensity ratio (FIR) of green emissions from 2H11/2 and 4S3/2 levels of Er3+ ion have been used to study optical temperature sensing in glass ceramics [20–23] and oxide bulks [24–28]. In these optical thermometry processes, it was assumed that the 2H11/2 and 4S3/2 levels of Er3+ ion were populated by a same two-photon up-conversion, and the FIR was independent of excitation powder of infrared lasers. In fact, almost all these 2H11/2 and 4S3/2 levels have the different pump power dependencies reported so far. This deviated a fact that the photon numbers needed to populate the 2H11/2 and 4S3/2 levels might be different at low and high pump powders. The conventional optical thermometry may lead to an error of sensitivity without consideration on the dependence of FIR on the pump power. Thus, it is necessary to explore the optical temperature behavior controlled by the excitation powder. In this work, novel Er3+ doped transparent Sr0.69La0.31F2.31 glass ceramics with high physical and chemical stability have been fabricated. The dependence of FIR on the pump power is studied through monitoring the up-conversion emissions from three couples of adjacent thermally coupled levels, 2H11/2/4S3/2, 4F9/2(1)/4F9/2(2), and 4I9/2(1)/4I9/2(2) at low (66.8 Mw/mm2) and high (374.8 mW/mm2) excitation powers. The thermal quenching ratio, population mechanism of thermally coupled levels, and temperature sensitivity are also studied by using the low and high excitation powers.

2. Experimental

Er3+ doped transparent Sr0.69La0.31F2.31 glass ceramics were fabricated by an improved melt-quenching method and subsequent heating. The raw materials are SiO2 (AR), Na2CO3 (AR), Al2O3 (AR), LaF3 (99.99%) and SrF2 (99.99%) with the composition (in mol%): 45SiO2–20Na2CO3–15Al2O3–6.2LaF3–13.8SrF2. Then 0.75 wt% ErF3 is doped into the raw materials. The materials were thoroughly mixed and ground in an mortar and heated at 1500°C for about 1 hour and then cast into a brass mold to proceed rapid quenching. The as-made glass was heat-treated at 660°C for 2 h to form transparent ceramics through a crystallization process.

The structure of the sample was investigated by X-ray diffraction (XRD) using a X'TRA (Switzerland ARL) equipment provided with Cu tube with Kα radiation at 1.54056 Å. The size and shape of the sample was observed by a JEM-2100 transmission electron microscope (JEOL Ltd., Tokyo, Japan). Luminescence spectra were obtained by the Acton SpectraPro Sp-2300 Spectrophotometer with a photomultiplier tube equipped with 980 nm laser as the excitation sources. Different temperature spectra were obtained by using an INSTEC HCS302 Hot and Cold System.

3. Results and discussions

Figure 1 illustrates the transmission electron microscope (TEM) and high-resolution transmission electron microscope (HRTEM) images, XRD, and unit cell structure of transparent Sr0.69La0.31F2.31 glass ceramics. TEM image in Fig. 1(a) confirms homogeneously distributed Sr0.69La0.31F2.31 spherical crystallites precipitated among the oxyfluoride glassy matrix. The size of Sr0.69La0.31F2.31 crystallite in the transparent ceramic is about 10 nm. The HRTEM observation in Fig. 1(b) confirms the well-defined lattice image and good crystallinity of the resulting nano-glass ceramics. The interplanar distance d value measured through two separated crystal planes from the HRTEM image is about 0.14 nm, which can be ascribed to the (111) crystal plane of Sr0.69La0.31F2.31 crystals. Structure of the resulting transparent nano-glass ceramics is studied by the XRD patterns in Fig. 1(c). One can find that the position and relative intensity of all the diffraction peaks can be readily indexed to the cubic Sr0.6La0.31F2.31 according to the JCPDS file No. 78-1143. Compared to the standard XRD pattern, the intensity ratio of the (111) peak of our sample is obviously larger than that of the standard value, suggesting that the sample tends to be preferentially oriented. The spatiality of the Sr0.69La0.31F2.31 structure along the [001] direction is illustrated in Fig. 1(d). The crystal structure is cubic Sr0.69La0.31F2.31 (ICSD 062285) [29]. The space group is Fm-3m. Each unit cell contains 4 formula units. In the unit cell of Sr0.69La0.31F2.31, there are 8c and 32f two F cation sites, and Sr and La occupy the same 4a cation site, respectively. La is substituted by the Er in Er3+-doped Sr0.6La0.31F2.31 crystal. To our knowledge, the synthesis and optical temperature sensing of transparent Er3+-doped Sr0.69La0.31F2.31 nano-glass ceramics has not been reported.

 figure: Fig. 1

Fig. 1 (a) TEM and (b) HRTEM images of Er3+ doped Sr0.69La0.31F2.31 glass ceramics. (c) XRD pattern of Er3+ doped Sr0.69La0.31F2.31 glass ceramics. The below standard data for cubic Sr0.69La0.31F2.31 (JCPDS 78-1143). (d) The schematic views of unit cell of Sr0.69La0.31F2.31 structure along b-direction.

Download Full Size | PDF

The transmission spectrum of Er3+ doped Sr0.69La0.31F2.31 glass ceramics is shown in Fig. 2. It shows that the resulting glass ceramics maintains a high transparency in the region from visible to infrared, which is due to the smaller size of the precipitated crystals much than the wavelength of visible and infrared light. The transmission peaks at 378 nm, 488 nm, 520 nm, 651nm, 797 nm, 980 nm and 1532 nm are observed, which are assigned to the transitions of 4I15/24G11/2, 4I15/24F7/2, 4I15/22H11/2, 4I15/24F9/2, 4I15/24I9/2, 4I15/24I11/2 and 4I15/24I13/2.

 figure: Fig. 2

Fig. 2 Transmission spectra of Er3+ doped Sr0.69La0.31F2.31 glass ceramics.

Download Full Size | PDF

In order to explore the property of optical temperature sensing, the temperature dependent photoluminescence spectra are studied in the temperature range from 298 K to 573 K by using a 980 nm laser with two kinds of powers of 66.8 mW/mm2 and 374.8mW/mm2, as shown in Fig. 3. It can be seen that the photoluminescence spectra in Fig. 3(a) and Fig. 3(b) contain green, red, and infrared emission bands, which are assigned to the 4I15/22H11/2 (522 nm), 4I15/24S3/2 (540 nm), 4I15/24F9/2(1) (650 nm), 4I15/24F9/2(2) (665 nm) and 4I15/24I9/2 (800 nm) transitions of Er3+ ion, respectively. One can find that the intensities of green and red emission bands greatly decrease as temperature increases without changing the peak positions of the emissions, while the intensity of infrared emission bands greatly increase as temperature increases with a peak shifting from 840 nm through 820 nm and to 800 nm. The emission spectra of Er3+ doped Sr0.69La0.31F2.31 glass ceramics are converted to the Commission International de I’Eclairage (CIE) 1931 chromaticity diagram [30], as shown in Fig. 4. With increasing the temperature from 298 K to 523 K, the luminescent color changes a little in green range at low and high excitation power.

 figure: Fig. 3

Fig. 3 Temperature dependent photoluminescence spectra of Er3+ doped Sr0.69La0.31F2.31 glass ceramics (a) at low 66.8 mW/mm2 excitation power, and (b) at high 374.8 mW/mm2 excitation power.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 CIE(X, Y) chromaticity coordinates diagrams (a) at low 66.8 mW/mm2 excitation power, and (b) at high 374.8 mW/mm2 excitation power.

Download Full Size | PDF

From the insets of Fig. 3(a) and Fig. 3(b), we can see only an infrared peak at 840 nm when the temperature is lower than 373 K. This peak begins to split when the temperature increases more than 373 K. To study peak splitting induced by the temperature, the temperature and excitation power dependent photoluminescence spectra from 750 nm to 900 nm are shown in Fig. 5. The infrared peak at 840 nm in Fig. 5(a) is not split at 298K with low excitation power of 66.8 mW/mm2, while the asymmetric emission band at 840 nm has been resolved into two Gaussian components peaking at 800 nm and 840 nm at 298 K with high excitation power of 374.8 mW/mm2, as shown in Fig. 5(b). Fixed the excitation power at 66.8 mW/mm2, the infrared peak at 840 nm obtained at 298 K (Fig. 5(a)) is resolved into three Gaussian components peaking at 800 nm, 820 nm, and 840 nm at high temperature of 573 K, as shown in Fig. 5(c). Fixed the excitation power at 374.8 mW/mm2, the asymmetric infrared peak is still resolved into two Gaussian components peaking at 800 nm and 840 nm at low (298 K) and high temperature (573 K), as shown in Fig. 5(b) and Fig. 5(d). Compared with the peak at 800 nm at 298 K, an obvious intensity increase of 800 nm emission is observed at 573 K. These change trends mean that the infrared emission band can be modified by changing the excitation power. At Sr0.69La0.31F2.31 host, the 4I9/2 level of Er3+ ion is split into three crystal field levels [31], such as 4I9/2(1), 4I9/2(2), and 4I9/2(3). The 4I9/2(3) is populated predominantly at low temperature, giving 840 nm emission with a 4I9/2(3)→4I15/2 transition. According to the Boltzmann distributing law [32], the ions on 4I9/2(3) level are thermally pumped onto the 4I9/2(2) and 4I9/2(1) levels step by step at high temperature. This thermal population decreases the intensity of 840 nm emission, and increases the intensities of 800 nm and 820 nm emissions. Thus, the peak shifting from 840 nm to 800 nm is observed with temperature increase.

 figure: Fig. 5

Fig. 5 Temperature and excitation power dependent photoluminescence spectra from 750 nm to 900 nm of Er3+ doped Sr0.69La0.31F2.31 glass ceramics. The black solid line represents the experimental data, and the colored dotted lines represent the fitting data.

Download Full Size | PDF

To evaluate the affection of temperature on luminescence quenching, the thermal quenching ratio (RQ) of emission bands induced by the temperature change is defined as follow:

RQ=1ITI0
where IT is luminescence intensity at different temperature T, and I0 is luminescence intensity at room temperature. Figure 6 shows the thermal quenching ratios of five emission bands through exciting the resulting glass ceramics with low and high excitation power. For green emission bands at 522 nm and 540 nm, it shows a reverse trend with the temperature increase. The value of RQ of 540 nm emission band increases with temperature increase from 298 K to 573 K. Oppositely, the value of RQ of 522 nm emission band shows some negative values, and decreases firstly and then increases with temperature increase, which means that the 2H11/2 state (522 nm) is populated thermally at high temperature [33]. Moreover, when the temperature is below 423 K, the values of RQ of 522 nm and 540 nm emission bands induced by high excitation power are more than those induced by low excitation power. However, when the temperature is above 423 K, the values of RQ of 522 nm and 540 nm emission bands induced by high excitation power are less than those induced by low excitation power. It means that the thermal quenching ratios of green emission bands can be inhibited by high excitation power. Differently, it shows that the thermal quenching ratios of red emissions at 650 nm and 665 nm can be increased by high excitation power. The intensity of 800 nm emission can be enhanced a lot by the increase of temperature and excitation power.

 figure: Fig. 6

Fig. 6 Thermal quenching ratios (RQ) of Er3+ doped Sr0.69La0.31F2.31 glass ceramics (a) at low 66.8 mW/cm2 excitation power, and (b) at high 374.8 mW/cm2 excitation power.

Download Full Size | PDF

To explore the origin of emissions of Er3+ ions at high temperatures, the relation between up-conversion emission intensity I and laser light intensity P is expressed as:

Ipn
where n is the number of absorbed photons in up-conversion process [16]. Figure 7 shows log–log plots of up-conversion intensity and pumping power for green, red, and infrared emissions at the different temperatures. The slopes of six fitted lines for 522 nm and 540 nm emissions change a little at three temperature points at 298 K, 373 K, and 573 K, and indicate that 522 nm and 540 nm emissions come from two-photon up-conversion process. The slopes of three six lines for 650 nm and 665 nm emissions are dependent on temperature, and indicate that two red emissions come from two-photon up-conversion process. The slopes of four lines for 800 nm and 820 nm emissions are dependent on temperature, and indicate that two infrared emissions come from one-photon up-conversion process. Thus, under 980 nm excitation, the excited state of 4I11/2 is populated by the ground state absorption with a 4I15/24I11/2 transition, shown in Fig. 8. The slopes for the 800 nm and 820 nm emissions have significant difference at 373K and 573K. It may be attributed to the competition between the nonradiative relaxation and thermal population of the 4I9/2(1) and 4I9/2(2) excited states [34,35]. Associated with phonons, three adjacent energy levels split from 4I9/2 are populated thermally. After an excited state absorption, the 4F7/2 state is populated by a 4I11/24F7/2 transition. The ions in the 4F7/2 state relax to the next lower energy levels 2H11/2/4S3/2 and 4F9/2(1)/4F9/2(2) through non-radiative relaxation, giving green and red emissions. According to Boltzmann distributing law, two adjacent energy levels, upper 2H11/2 (4F9/2(1)) level and lower 4S3/2 (4F9/2(2)) level, can be thermally populated and depopulated with temperature increase [36]. The temperature-dependent fluorescence intensity ratios, IU1/IL1 (522 nm/540 nm), IU2/IL2 (650 nm/665 nm), and IU3/IL3 (800 nm/820 nm), may be used to be as thermally coupled levels, such as TCL1 (2H11/2/4S3/2), TCL2 (4F9/2(1)/4F9/2(2)), and TCL3 (4I9/2(1)/4I9/2(2)), for optical temperature sensing.

 figure: Fig. 7

Fig. 7 Log–log plots of intensity and pumping power for (a) 522 nm, (b) 540 nm, (c) 650 nm, (d) 665 nm, (e) 800 nm, and (f) 820 nm emissions at different temperatures.

Download Full Size | PDF

 figure: Fig. 8

Fig. 8 Population mechanism of Er3+ at high temperature under 980 nm excitation.

Download Full Size | PDF

It was reported that the emission intensity ratio (R) from the thermally coupled energy levels of active ions is expressed as

R=AeΔEkT+B
where ΔE is the energy difference between thermally coupled levels, T the absolute temperature, and k is the Boltzmann constant. A and B are constants [3]. From the spectra in Fig. 3, the experimental values of ΔE are about 638.6 cm−1, 347.02 cm−1, and 304.9 cm−1 for thermally coupled levels of TCL1, TCL2 and TCL3. If fitted with the Eq. (3), the values of ΔE are about 735.1 cm−1 (66.8 mW/cm2 excitation power) and 653.7 cm−1 (374.8 mW/cm2 excitation power) for thermally coupled level of TCL1, 88.4 cm−1 (66.8 mW/cm2 excitation power) and 93.5 cm−1 (374.8 mW/cm2 excitation power) for thermally coupled level of TCL2, 1097.3 cm−1 (66.8 mW/cm2 excitation power), 154.4 cm−1 (374.8 mW/cm2 excitation power) for thermally coupled level of TCL3. Compared with the experimental values of ΔE, the errors are too big to ignore. The experimental ΔE is calculated from the two emission peaks originated from 2H11/2 and 4S3/2 in the experimental spectrum. The theoretical ΔE will be obtained through fitting some data points of luminescence intensity ratios at different temperature by using the Eq. (3). The Eq. (3) was a theoretical model and established by only considering the Boltzmann distribution between two thermally coupled levels and radiative transitions from thermally coupled levels to ground state [37], without considering nonradiative relaxation and energy transfer between the host and the rare-earth ions. However, nonradiative relaxation and energy transfer are not neglected in real photoluminescence process at high temperature. Thus, the error between experimental and theoretical values of ΔE appears. A large error value means that the thermally coupled levels are populated and dispopulated at high temperature through a complex dynamic process, not only Boltzmann distribution and radiative transitions, but also nonradiative relaxation and energy transfer.

Notably, the slopes for red and infrared emissions in Fig. 7 at low and high temperatures are not identical, which suggests that the fluorescence intensity ratios of thermally coupled levels of the TCL2 and TCL3, are expected to be susceptible to the pumping powers. This unsteady phenomenon is attributed to the competition between the thermal population and the nonradiative relaxation of the intermediate excited states [34]. The nonradiative relaxation possibility is expressed as

wij=w0(1ew/kT)P
where w0 is its value at zero temperature, ħw is the phonon energy, and the exponent p is the phonon numbers. Fluorescence intensity ratios of adjacent thermally coupled levels are not be fitted well with a single exponential model from Eq. (3), due to the fact that nonradiative relaxation possibility changes also with the temperature.

Considering the similar exponential model of R and wij, the relation between R and T is modified as

LnR=aT+b
where a is constant. The b is a correction term for the comprehensive population of thermally coupled energy levels induced by the thermal population, nonradiative relaxation and so on. Figure 9 shows temperature dependent emission intensity ratios of green, red, and infrared emissions at the different excitation powers. One can find that the experimental points can be fitted well with a linear mode. The slope values of green, red, and infrared emissions are dependent on excitation powers. It means that the fluorescence intensity ratios of thermally coupled levels of TCL1, TCL2, and TCL3 are susceptible to the pumping power. It is noting that the values of fitted slopes in Fig. 9 are not equal to the theoretical values of the ΔE/K, due to a complex dynamic up-conversion process induced by the high temperature and high excitation powers.

 figure: Fig. 9

Fig. 9 Arrhenius plots of temperature dependent emission intensity ratios of (a) 522 nm/540 nm, (b) 650 nm/665 nm, (c) 800 nm/820 nm at low 66.8 mW/cm2 excitation power and at high 374.8 mW/cm2 excitation power.

Download Full Size | PDF

The sensitivity is one of the key parameter to determine the property of optical thermometry [3]. The sensitivity is defined as

S=dRdT=aT2ebTaT
where a and b are the constants from Eq. (5). The excitation power dependent sensitivity for three thermally coupled energy levels, such as TCL1, TCL2, and TCL3, are calculated in Fig. 10. All the sensitivity lines increase and then decrease with the increase of temperature. Three thermally coupled energy levels show the high sensitivity at low excitation power. The maximum values at (452.6K, 0.0014/K), (494.8K, 0.0015/K), (62.7K, 0.0063/K), (65.8K, 0.0061/K), (584K, 0.0039/K), and (934.4K, 0.0045/K) are observed for the TCL1, TCL2, and TCL3. From Fig. 7, the slopes of fitted lines for TCL1 are very closer at three temperature points at 298 K, 373K, and 573 K, revealing that TCL1 is especially stable for monitoring the temperature in wide temperature range. It is very sensitive to measure temperature below 250 K by using the fluorescence intensity ratio of the TCL2, shown in Fig. 10. Most of sensors based on up-conversion luminescence of Er3+ ion showed excellent sensitivity property for temperatures ranging from room to high temperatures, with the maximum sensitivity values at high temperature more than 300 K [17–25]. Reports on optical thermometry below room temperature were very scarce. The slope change induced by the temperature in Fig. 7 shows the TCL2 is not stable when the temperature excesses 373 K. Thus, TCL2 is suitable for monitoring the temperature below 373 K. Even though the TCL3 has high sensitivity at high temperature range more than 500 K, Fig. 7 shows the TCL3 is not stable with the temperature increase. Thus, the combined use of TCL1 and TCL2 as thermally coupled energy levels will get a more precise temperature reading at low excitation power.

 figure: Fig. 10

Fig. 10 The excitation power dependence of sensitivity.

Download Full Size | PDF

4. Conclusion

In summary, Er3+ doped transparent Sr0.69La0.31F2.31 glass ceramics was fabricated by a improved melt-quenching method. XRD, TEM and HRTEM analysis show that cubic Sr0.69La0.31F2.31 crystals are homogeneously precipitated among the silicate aluminum glass matrix. It was observed that the spectrum structure, thermal quenching ratio, fluorescence intensity ratio, and sensitivity from three thermally coupled levels, such as 2H11/2/4S3/2, 4F9/2(1)/4F9/2(2), and 4I9/2(1)/4I9/2(2), are strongly dependent on the change of pump powers. A new fitting method is proposed to establish the relation between fluorescence intensity ratios and temperature. The 2H11/2/4S3/2 levels has high population stability at high temperature, while 4F9/2(1)/4F9/2(2) levels has high population stability at low temperature. A more precise temperature measuring can be obtained through the combining use of 2H11/2/4S3/2 and 4F9/2(1)/4F9/2(2) as thermally coupled energy levels at low excitation power at 66.8 mW/mm2. This work also overcomes difficulty in low temperature measurement for conventional optical temperature sensing technology. The resulting glass ceramic is a very promising candidate for accurate optical temperature sensors with much higher sensitivity and wide temperature range.

Funding

This work was supported by National Natural Science Foundation of China (NSFC) (11404171), Natural Science Youth Foundation of Jiangsu Province (BK20130865), the Six Categories of Summit Talents of Jiangsu Province of China (2014-XCL-021), and the Natural Science Foundation of the Jiangsu Higher Education Institutions of China (14KJB430020), and Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT & Future Planning (NRF-2013RA1A2009154).

References and links

1. H. Xia, J. Feng, Y. Wang, J. Li, Z. Jia, and C. Tu, “Evaluation of spectroscopic properties of Er(3+)/Yb(3+)/Pr(3+): SrGdGa3O7 crystal for use in mid-infrared lasers,” Sci. Rep. 5, 13988 (2015). [CrossRef]   [PubMed]  

2. C. Yan, X. Liu, H. Li, X. Xia, H. Lu, and W. Zheng, “Color three-dimensional display with omnidirectional view based on a light-emitting diode projector,” Appl. Opt. 48(22), 4490–4495 (2009). [CrossRef]   [PubMed]  

3. X. Wang, Q. Liu, Y. Bu, C.-S. Liu, T. Liu, and X. Yan, “Optical temperature sensing of rare-earth ion doped phosphors,” RSC Advances 5(105), 86219–86236 (2015). [CrossRef]  

4. J. C. Boyer, F. Vetrone, L. A. Cuccia, and J. A. Capobianco, “Synthesis of Colloidal Upconverting NaYF4 Nanocrystals Doped with Er3+, Yb3+ and Tm3+, Yb3+ via Thermal Decomposition Of Lanthanide Trifluoroacetate Precursors,” J. Am. Chem. Soc. 128(23), 7444–7445 (2006). [CrossRef]   [PubMed]  

5. 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]  

6. J. Feenstra, I. F. Six, M. A. H. Asselbergs, R. H. van Leest, J. de Wild, A. Meijerink, R. E. I. Schropp, A. E. Rowan, and J. J. Schermer, “Er(3+)/Yb(3+) upconverters for InGaP solar cells under concentrated broadband illumination,” Phys. Chem. Chem. Phys. 17(17), 11234–11243 (2015). [CrossRef]   [PubMed]  

7. X. Zhai, S. Liu, X. Liu, F. Wang, D. Zhang, G. Qin, and W. Qin, “Sub-10 nm BaYF5:Yb3+,Er3+ core–shell nanoparticles with intense 1.53 μm fluorescence for polymer-based waveguide amplifiers,” J. Mater. Chem. C Mater. Opt. Electron. Devices 1(7), 1525–1530 (2013). [CrossRef]  

8. D. Chen, Y. Zhou, Z. Wan, H. Yu, H. Lu, Z. Ji, and P. Huang, “Tunable upconversion luminescence in self-crystallized Er(3+):K(Y(1-x)Yb(x))3F10 nano-glass-ceramics,” Phys. Chem. Chem. Phys. 17(11), 7100–7103 (2015). [CrossRef]   [PubMed]  

9. W. Zhang, F. Ding, and S. Y. Chou, “Large Enhancement of Upconversion Luminescence of NaYF4:Yb3+/Er3+ Nanocrystal by 3D Plasmonic Nano-Antennas,” Adv. Mater. 24(35), 236–241 (2012). [CrossRef]  

10. Y. Zhu, W. Xu, S. Cui, M. Liu, C. Lu, H. Song, and D. Kim, “Controlled size and morphology, and phase transition of YF3:Yb3+,Er3+ and YOF:Yb3+,Er3+ nanocrystals for fine color tuning,” J. Mater. Chem. C Mater. Opt. Electron. Devices 4(2), 331–339 (2016). [CrossRef]  

11. S. Fischer, B. Fröhlich, K. W. Krämer, and J. C. Goldschmidt, “Relation between excitation power density and Er3+ doping yielding the highest absolute upconversion quantum yield,” J. Phys. Chem. C 118(51), 30106–30114 (2014). [CrossRef]  

12. G. Y. Chen, H. J. Liang, H. C. Liu, G. Somesfalean, and Z. G. Zhang, “Anomalous power dependence of upconversion emissions in Gd2O3:Er3+ nanocrystals under diode laser excitation of 970 nm,” J. Appl. Phys. 105(11), 114315 (2009). [CrossRef]  

13. G. Chen, T. Y. Ohulchanskyy, A. Kachynski, H. Agren, and P. N. Prasad, “Intense visible and near-infrared upconversion photoluminescence in colloidal LiYF4:Er3+ nanocrystals under excitation at 1490 nm,” ACS Nano 5(6), 4981–4986 (2011). [CrossRef]   [PubMed]  

14. Y. Qu, X. Kong, Y. Sun, Q. Zeng, and H. Zhang, “Effect of excitation power density on the upconversion luminescence of LaF3:Yb3+, Er3+ nanocrystals,” J. Alloys Compd. 485(1–2), 493–496 (2009). [CrossRef]  

15. Y. Tian, R. Hua, J. Yu, J. Sun, and B. Chen, “The effect of excitation power density on frequency upconversion in Yb3+/Er3+ codoped Gd6WO12 nanoparticles,” Mater. Chem. Phys. 133(2–3), 617–620 (2012). [CrossRef]  

16. F. Auzel, “Upconversion and anti-Stokes processes with f and d ions in solids,” Chem. Rev. 104(1), 139–174 (2004). [CrossRef]   [PubMed]  

17. K. Zheng, W. Song, G. He, Z. Yuan, and W. Qin, “Five-photon UV upconversion emissions of Er3+ for temperature sensing,” Opt. Express 23(6), 7653–7658 (2015). [CrossRef]   [PubMed]  

18. S. F. LeÓn-Luis, U. R. Rodríguez-Mendoza, E. Lalla, and V. Lavín, “Temperature sensor based on the Er3+ green upconverted emission in a fluorotellurite glass,” Sens. Actuators B Chem. 158(1), 208–213 (2011). [CrossRef]  

19. D. Jaque and F. Vetrone, “Luminescence nanothermometry,” Nanoscale 4(15), 4301–4326 (2012). [CrossRef]   [PubMed]  

20. S. F. Leon-Luis, U. R. Rodríguez-Mendoza, P. Haro-González, I. R. Martín, and V. Lavín, “Role of the host matrix on the thermal sensitivity of Er3+ luminescence in optical temperature sensors,” Sens. Actuators B Chem. 174, 176–186 (2012). [CrossRef]  

21. S. F. Leon-Luis, U. R. Rodríguez-Mendoza, I. R. Martín, E. Lalla, and V. Lavín, “Effects of Er3+ concentration on thermal sensitivity in optical temperature fluorotellurite glass sensors,” Sens. Actuators B Chem. 176, 1167–1175 (2013). [CrossRef]  

22. F. Vetrone, R. Naccache, A. Zamarrón, A. Juarranz de la Fuente, F. Sanz-Rodríguez, L. Martinez Maestro, E. Martín Rodriguez, D. Jaque, J. García Solé, and J. A. Capobianco, “Temperature sensing using fluorescent nanothermometers,” ACS Nano 4(6), 3254–3258 (2010). [CrossRef]   [PubMed]  

23. P. Du, L. H. Luo, W. P. Li, Q. Y. Yue, and H. B. Chen, “Optical temperature sensor based on upconversion emission in Er-doped ferroelectric 0.5Ba(Zr0.2Ti0.8)O3-0.5(Ba0.7Ca0.3)TiO3 ceramic,” Appl. Phys. Lett. 104(15), 152902 (2014). [CrossRef]  

24. V. K. Rai, A. Pandey, and R. Dey, “Photoluminescence study of Y2O3:Er3+-Eu3+-Yb3+ phosphor for lighting and sensing applications,” J. Appl. Phys. 113(8), 083104 (2013). [CrossRef]  

25. R. Dey, A. Pandey, and V. K. Rai, “Er3+-Yb3+ and Eu3+-Er3+-Yb3+ codoped Y2O3 phosphors as optical heater,” Sens. Actuators B Chem. 190, 512–515 (2014). [CrossRef]  

26. V. Lojpur, G. Nikoli, and M. D. Dramianin, “Luminescence thermometry below room temperature via up-conversion emission of Y2O3:Yb3+,Er3+ nanophosphors,” J. Appl. Phys. 115(20), 203106 (2014). [CrossRef]  

27. A. K. Soni, A. Kumari, and V. K. Rai, “Optical investigation in shuttle like BaMoO4:Er3+–Yb3+ phosphor in display and temperature sensing,” Sens. Actuators B Chem. 216, 64–71 (2015). [CrossRef]  

28. B. Dong, D. P. Liu, X. J. Wang, T. Yang, S. M. Miao, and C. R. Li, “Optical thermometry through infrared excited green upconversion emissions in Er3+–Yb3+Er3+–Yb3+ codoped Al2O3,” Appl. Phys. Lett. 90(18), 181117 (2007). [CrossRef]  

29. L. A. Muradyan, B. A. Maksimov, B. F. Mamin, N. N. Bydanov, V. A. Sarin, B. P. Sobolev, and V. I. Simonov, “Atomic structure of nonstoichiometric phase Srsub (0, 69) Lasub (0, 31) Fsub (2, 31),” Kristallografiya 31(2), 248–251 (1986).

30. X. F. Wang, X. H. Yan, Y. Y. Bu, J. Zhen, and Y. Xuan, “Fabrication, photoluminescence, and potential application in white light emitting diode of Dy3+–Tm3+ doped transparent glass ceramics containing GdSr2F7 nanocrystals,” Appl. Phys., A Mater. Sci. Process. 112(2), 317–322 (2013). [CrossRef]  

31. K. P. O’Donnell and V. Dierolf, “Rare-Earth Doped III-Nitrides for Optoelectronic and Spintronic Applications,” Springer, ISBN 978–90–481–2876–1, 124,1–355 (2010)

32. M. D. Shinn, W. A. Sibley, M. G. Drexhage, and R. N. Brown, “Optical transitions of Er3+ ions in fluorozirconate glass,” Phys. Rev. B 27(11), 6635–6648 (1983). [CrossRef]  

33. W. Xu, Z. Zhang, and W. Cao, “Excellent optical thermometry based on short-wavelength upconversion emissions in Er3+/Yb3+ codoped CaWO4.,” Opt. Lett. 37(23), 4865–4867 (2012). [CrossRef]   [PubMed]  

34. L. Xing, Y. Xu, R. Wang, and W. Xu, “Influence of temperature on upconversion multicolor luminescence in Ho3+/Yb3+/Tm(3+)-doped LiNbO3 single crystal,” Opt. Lett. 38(14), 2535–2537 (2013). [CrossRef]   [PubMed]  

35. J. F. Suyver, A. Aebischer, S. García-Revilla, P. Gerner, and H. U. Güdel, “Anomalous power dependence of sensitized upconversion luminescence,” Phys. Rev. B 71(12), 125123 (2005). [CrossRef]  

36. S. A. Wade, S. F. Collins, and G. W. Baxter, “Fluorescence intensity ratio technique for optical fiber point temperature sensing,” J. Appl. Phys. 94(8), 4743–4756 (2003). [CrossRef]  

37. G. K. Liu and X. Y. Chen, “Chapter 233: Spectroscopic properties of lanthanides in nanomaterials,” Handbook on the Physics and Chemistry of Rare Earths 37(07), 99–169, (2007).

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 (10)

Fig. 1
Fig. 1 (a) TEM and (b) HRTEM images of Er3+ doped Sr0.69La0.31F2.31 glass ceramics. (c) XRD pattern of Er3+ doped Sr0.69La0.31F2.31 glass ceramics. The below standard data for cubic Sr0.69La0.31F2.31 (JCPDS 78-1143). (d) The schematic views of unit cell of Sr0.69La0.31F2.31 structure along b-direction.
Fig. 2
Fig. 2 Transmission spectra of Er3+ doped Sr0.69La0.31F2.31 glass ceramics.
Fig. 3
Fig. 3 Temperature dependent photoluminescence spectra of Er3+ doped Sr0.69La0.31F2.31 glass ceramics (a) at low 66.8 mW/mm2 excitation power, and (b) at high 374.8 mW/mm2 excitation power.
Fig. 4
Fig. 4 CIE(X, Y) chromaticity coordinates diagrams (a) at low 66.8 mW/mm2 excitation power, and (b) at high 374.8 mW/mm2 excitation power.
Fig. 5
Fig. 5 Temperature and excitation power dependent photoluminescence spectra from 750 nm to 900 nm of Er3+ doped Sr0.69La0.31F2.31 glass ceramics. The black solid line represents the experimental data, and the colored dotted lines represent the fitting data.
Fig. 6
Fig. 6 Thermal quenching ratios (RQ) of Er3+ doped Sr0.69La0.31F2.31 glass ceramics (a) at low 66.8 mW/cm2 excitation power, and (b) at high 374.8 mW/cm2 excitation power.
Fig. 7
Fig. 7 Log–log plots of intensity and pumping power for (a) 522 nm, (b) 540 nm, (c) 650 nm, (d) 665 nm, (e) 800 nm, and (f) 820 nm emissions at different temperatures.
Fig. 8
Fig. 8 Population mechanism of Er3+ at high temperature under 980 nm excitation.
Fig. 9
Fig. 9 Arrhenius plots of temperature dependent emission intensity ratios of (a) 522 nm/540 nm, (b) 650 nm/665 nm, (c) 800 nm/820 nm at low 66.8 mW/cm2 excitation power and at high 374.8 mW/cm2 excitation power.
Fig. 10
Fig. 10 The excitation power dependence of sensitivity.

Equations (6)

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

R Q = 1 I T I 0
I p n
R = A e Δ E k T + B
w i j = w 0 ( 1 e w / k T ) P
L n R = a T + b
S = d R d T = a T 2 e b T a T
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.