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

Constructing ultra-sensitive dual-mode optical thermometers: Utilizing FIR of Mn4+/Eu3+ and lifetime of Mn4+ based on double perovskite tellurite phosphor

Open Access Open Access

Abstract

A strategy of optical temperature sensing was developed by using various thermal quenching of Mn4+ and Eu3+ for double perovskite tellurite phosphor in optical thermometers. Herein, SrGdLiTeO6 (SGLT): Mn4+,Eu3+ phosphors were synthesized by a high-temperature solid-state reaction method. The temperature-dependent emission spectra indicated that two distinguishable emission peaks originated from Eu3+ and Mn4+ exhibited significantly diverse temperature responses. Therefore, optical thermometers with a dual-mode mechanism were designed by employing a fluorescence intensity ratio (FIR) of Mn4+ (2Eg4A2g) and Eu3+ (5D07F1,2) and the decay lifetime of Mn4+ as the temperature readouts. The temperature sensing of the phosphors ranging from 300 to 550 K were studied. The maximum relative sensitivities (Sr) are obtained as 4.9% K−1 at 550 K. Meanwhile, the 695 nm emission of Mn4+ possessed a temperature-dependent decay lifetime with Sr of 0.229% K−1 at 573 K. Relevant results demonstrate the SrGdLiTeO6:Mn4+, Eu3+ phosphor as an optical thermometer candidate and also provide constructive suggestions and guidance for constructing high-sensitivity dual-mode optical thermometers.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Optical thermometry technology based on rare earth (RE) and transition metal (TM) ions doped luminescent materials has attracted much attention thanks to non-contact, high spatial resolution, rapid response, and anti-electromagnetic interference [1,2]. In recent years, temperature-sensitive optical parameters, such as the luminescent intensity, spectral position, decay lifetime, bandwidth shift, and fluorescence intensity ratio (FIR) were adopted to measure temperature [3,4]. Especially, FIR and decay lifetime-based sensing of temperature independent of spectral losses, external interferences were attracted wide attention [5,6]. FIR strategy was usually concentrated on two thermally coupled levels (TCL) of RE, such as 2H11/2 and 4S3/2 states of Er3+, and non-thermally coupled levels are often adopted [7,8]. However, there are inherent limitations in this strategy such as low relative sensitivity and low signal discriminability owing to the restricted energy gap (200 <ΔE < 2000cm−1) [9,10]. For TM ions, although the luminescence intensity is very susceptible to variation of temperature owing to strong electron-phonon coupling, there are limitations in practicality [11]. Therefore, it is vital to design a new FIR temperature-sensing strategy and improve the performance of the optical thermometers.

A number of studies that showed that FIR strategy based on dual-emitting centers (such as Tb3+-Eu3+, Bi3+-Eu3+, Eu2+-Eu3+ and Pr3+-Tb3+, etc.) obeying different temperature dependence has been proposed [12,13]. This FIR strategy gives high relative sensitivity owing to independent of energy gap spacing [14,15]. To obtain a high dramatic variation of FIR value with temperature, choosing suitable dopant ions and matrix is very crucial. It was known that RE ions with 4fn electronic configurations have a weak electron-lattice interaction, whereas TM ions as well as 3dn configurations possess a strong electron-phonon coupling. Consequently, different thermal quenching were achieved for RE and TM ions. TM/Ln dopants exhibit high feasibility for a FIR optical thermometer [1618]. In order to realize it, considerable research efforts have been devoted to develop appropriate hosts to achieve efficient dual emissions for both RE and TM activators. Moreover, decay lifetime-based thermometry have attracted enormous attention, which is independent of calibration measurement and insusceptible to external electromagnetic interference [19,20]. Generally, RE3+ (Eu3+, Sm3+) and TM3+ (Mn4+, Cr3+) doped phosphors, nanocrystals, and glass ceramics were beneficial for decay lifetime-based optical thermometers [21,22]. Thus, it is very necessary for exploring novel dual-mode optical temperature sensors based on the FIR of TM3+/ RE3+ and the lifetimes of TM3+ ions.

As we all know, perovskite materials are well-known for its versatile optoelectronic applications in light-emitting diodes, solar cells, photon detection, photocatalyst and temperature sensor [23]. Cation substitution based on the A-site or B-site of the basic ABX3 structure of the perovskites accommodate for novel materials with fascinating properties. Especially, double perovskite compounds such as A2BB′O6, AA′BB′O6, and AA′B2O6 were extensively investigated owing to allowing for various radii and valence state ions [24]. Much work so far has focused on Mn4+ and RE3+ (Eu3+, Dy3+) doped double perovskite phosphor. Yin et al. have reported on double perovskite A2LaNbO6:Mn4+,Eu3+ and BaLaMgNbO6:Mn4+,Dy3+ phosphors for the application of optical temperature sensing [25,26]. Tellurates are intrigued to consider as host materials for optoelectronic applications owing to excellent chemical stabilities and lower phonon energy. The structure of SrLnLiTeO6 (Ln = La, Gd, Eu) was pointed out that smaller radius of A site in the SrLnLiTeO6 would result in bigger structure distortion [27]. Subodh has reported the phonon energy of ALaLiTeO6 (A = Ba, Sr) and showed that smaller phonon energy (725 cm-1) can suppress non-radiative multi-phonon transition, which indicating that these materials can be regarded as good luminescent matrix materials [28]. The crystal structure, luminescent properties and thermal quenching of Eu3+ doped SrLnLiTeO6 (Ln = Gd, La) phosphors had been investigated [29,30]. However, Mn4+ doped SrLnLiTeO6 (Ln = Gd, La) phosphors have not yet been investigated. Furthermore, the temperature sensing based on decay lifetime for Sm3+ doped Lu2MoO6 and Tb3+,Eu3+ codoped Ca8ZnLa(PO4)7 phosphors was already investigated in our previous work [31,32]. However, to the best of knowledge, the temperature sensing based on fluorescent intensity ratio of Mn4+/RE3+ and decay lifetimes of Mn4+ for SrLnLiTeO6 phosphors have not yet been investigated so far.

Herein, in view of this, we chose the SGLT as the matrix. The novel Mn4+, Eu3+ singly and codoped SGLT phosphors were synthesized by a facile high-temperature solid-state reaction method. The lifetime of Mn4+ and the FIR of Mn4+ and Eu3+ are both susceptible to temperature and are independent on measurement conditions and smaller measurement errors such as excitation power fluctuation, geometry change of the sample, and atmospheric pressure change, etc., so the dual-mode non-contact optical thermometry strategy was constructed based on fluorescent intensity ratio of Mn4+/Eu3+ and luminescence decay lifetimes of Mn4+. The dual-mode temperature readout can provide more accurate measurement than readout using only one type. The experimental results demonstrate that the resultant SGLT: Mn4+, Eu3+ phosphors have promising application in self-referencing optical thermometers. It was believed that this work provides a guidance for the exploration of well-performing temperature sensors.

2. Experimental

Mn4+, Eu3+ singly and codoped SGLT phosphors were prepared by high temperature solid state reaction method. The SrCO3, Li2CO3 (Analytical Reagent (A.R.)), TeO2 (A.R.), Gd2O3 (99.99%), Eu2O3 and MnCO3 (99.99%) were the raw materials. All powders with stoichiometric molar ratio were weighed and ground finely in an agate mortar. Subsequently, the powders were further calcined at 1300 °C for 8 hours in a muffle furnace. Finally, the resultant phosphors were obtained. The X-ray diffraction (XRD) patterns of all phosphors were examined by a powder diffractometer XD-2 (Persee General Instrument Co., Ltd., Beijing) with CuKα radiation (λ=1.5406 Å). The morphology and elemental composition were recorded by a field-emission scanning electron microscope (Hitachi, SU3500) with an attached energy dispersive X-ray (EDX) spectrometer. The room temperature excitation, emission spectra as well as luminescence decay kinetics were measured with a FLS1000 steady state spectrometer (Edinburgh Instrument Ltd., Livingston, UK). The temperature-dependent PL spectra and luminescence decay curves with the range of 298-573 K were measured using the same spectrometer equipped with high temperature attachments.

3. Results and discussion

The XRD patterns of SGLT, SGLT: 5%Eu3+, SGLT: 0.7%Mn4+, and SGLT: 5%Eu3+, 0.7%Mn4+ phosphors were measured and shown in Fig. 1(a). It can be found that the structure of these phosphors is identical with that of SrEuLiTeO6 (JCPDS card no.80-0078), which belongs to the B site ordered double perovskite of AA′BB′O6 structures [29]. The prepared SGLT phosphors possess monoclinic structure with space group P21/n. The crystal structure of samples was not changed by doping Eu3+ and Mn4+, manifesting that the solid solution was obtained. Additionally, a new minor Li3Eu3Te2O12 (JCPDS no. 25-1178) impurity phase (marked with *) can be found in the XRD patterns of the SGLT: 5%Eu3+ and SGLT: 5%Eu3+,0.7%Mn4+ samples and was very weak to be neglected. The crystal structure of SGLT was shown in Fig. 1(b). The Li and Te atoms for the B-site cations were both coordinated by six O atoms with B-site rock-salt ordering structure, which presented octahedral centers appropriate for doping Mn4+ ions in the SGLT host. The Sr2+,Gd3+/Eu3+ cations occupy the A sites at random and form A-site layered ordering structure. Furthermore, the ions radii of Te6+ and Mn4+ (coordinated number (CN) =6) are 0.56 Å and 0.53 Å, respectively, indicating that the Mn4+ ions substituted for Te6+ ions in the SGLT matrix [33]. Eu3+ ions occupy without inversion center and prefer to replace for the Gd3+ ions owing to the similar radii between Eu3+ (r=1.12 Å, CN=9) and Gd3+ (r=1.107 Å, CN=9), which is beneficial for Eu3+ luminescence.

 figure: Fig. 1.

Fig. 1. XRD patterns of SGLT, SGLT: 0.7%Mn4+, SGLT: 5%Eu3+ and SGLT: 5%Eu3+, 0.7%Mn4+ compared with ICSD no. 067852 of SrEuLiTeO6 (a).The crystal structure of SGLT (b).

Download Full Size | PDF

The emission spectra of SGLT: x%Mn4+ (x = 0.1, 0.3, 0.5, 0.6, 0.7, 0.8) phosphors under 302 nm excitation are shown in Fig. S1. As expected, the dominant red emission band centered at 695 nm is assigned to the2Eg4A2g transition of Mn4+. It was shown that the emission intensity at 695 nm increases initially as a consequence of increased amount of Mn4+ and reaches maximum when the doping concentration is 0.7%, and subsequently decreases due to the concentration quenching effect at higher Mn4+ concentrations. Therefore, the Mn4+ doping concentration was fixed at 0.7%. In order to investigate the room temperature and temperature-dependent luminescent properties, the excitation and emission spectra of singly doped Eu3+, singly doped Mn4+, and Eu3+, Mn4+ codoped SGLT samples were recorded as shown in Fig. 2, respectively. As depicted in Fig. 2(a), the excitation spectrum (the blue line) of Eu3+ single-doped SGLT monitored at 615 nm includes a wide excitation band and a few of sharp excitation peaks. The broad excitation band centered at 292 nm in the range from 243 nm to 334 nm was assigned to the overlap of Eu-O and Te-O charge transfer state. While the sharp excitation peaks located at 362 nm (7F05D4), 377 nm (7F05L7), 382 nm (7F05L8), 385 nm (7F05GJ), 395 nm (7F05L6), 416 nm (7F05D3), 465 nm (7F05D2), 527 nm (7F05D1), 535 nm (7F15D1) are due to the characteristic excitation of the intra-4f transition of Eu3+, respectively [34]. The photoluminescence spectrum (The red line) of Eu3+ single-doped SGLT excited by 302 nm appears two dominant emission peaks centered at 591 nm and 615 nm, which belong to the 5D07F1 magnetic transition and 5D07F2 electric transition of Eu3+, respectively. As depicted in Fig. 2(b), the excitation spectrum monitored at 695 nm of Mn4+ single-doped SGLT are composed of two broad excitation bands, one band centered at 323 nm is corresponding to the Mn-O charge transfer and the transition of Mn4+: 4A2g4T1g. The other one with strongest excitation at 468 nm is attributed to Mn4+: 4A2g4T2g. Under the excitation of 302nm, the emission spectrum presents an emission band ranging from 641 nm to 794 nm with the strongest emission at 695 nm, which belongs to the 2Eg4A2g transition of Mn4+ [35]. Figure 2(c) reveals the excitation and emission spectra of Eu3+, Mn4+ codoped SGLT sample. The excitation spectra exhibits the spectral overlap monitored at 615 nm of Eu3+ and 695 nm of Mn4+ ions, making probable of simultaneous excitation of Mn4+ and Eu3+ ions under 302 nm excitation. Therefore, the emission spectrum at 302 nm excitation includes two emission peaks centered at 591 nm and 615 nm originating from the 5D07F1,2 transitions of Eu3+ and the emission band of 695 nm owing to the 2Eg4A2g transition of Mn4+, respectively.

 figure: Fig. 2.

Fig. 2. Excitation and emission spectra of SGLT: 5%Eu3+ (a), SGLT: 0.7%Mn4+ (b) and SGLT: 5%Eu3+,0.7%Mn4+ (c) phosphors.

Download Full Size | PDF

The emission spectra of SGLT:5%Eu3+,0.7%Mn4+ phosphor with the temperature of 300-550 K were displayed in Fig. 3(a). It can be observed that both the Eu3+ and Mn4+ intensities declined with elevating temperature owing to thermal quenching effect. Notably, the 2Eg4A2g emission intensity of Mn4+ decreases much more rapidly as compared with that of the 5D07F2 and5D07F1 intensities of Eu3+ with the increase of temperature. Therefore, Eu3+ serves as a natural self-referencing for thermal decay of Mn4+ so that the self-referencing optical thermometry of Mn4+/Eu3+ can be constructed. The two-dimensional fluorescence topographical mapping of SGLT: 5%Eu3+,0.7%Mn4+ under 302 nm excitation was shown in Fig. 3(b). Apparently, the intensities of Eu3+ and Mn4+ can recover after one cycle process (300K→550K→300 K), indicating the repeatability and reliability of the SGLT: 5%Eu3+,0.7%Mn4+ sample for optical thermometry.

 figure: Fig. 3.

Fig. 3. Temperature-dependent emission spectra (a) and two-dimensional fluorescence topographical mapping (b) of SGLT:5%Eu3+,0.7%Mn4+ under 302 nm excitation in the temperature range from 300 K to 550 K.

Download Full Size | PDF

To gain more insight into the temperature sensing, the integrated intensities of Eu3+: 5D07F1,2 and Mn4+: 2Eg4A2g dependence on temperature were displayed in Fig. 4(a,d), respectively. The variation of the FIR1 of Eu3+: 5D07F1 (577-602 nm) and Mn4+: 2Eg4A2g (637-775 nm) and FIR2 of Eu3+:5D07F2 (604-633 nm) and Mn4+: 2Eg4A2g (637-775 nm) dependence on temperature are shown and fitted in Fig. 4(b, e), respectively. Obviously, the FIR1,2 of Eu3+: 5D07F1 (577-602 nm), 5D07F2 (604-633 nm) and Mn4+: 2Eg4A2g (637-775 nm) exhibits the linear dependence on temperature. Consequently, the temperature dependence on decay curve can be respectively expressed by the following Eqs. (1) and (2):

$$FI{R_1} ={-} 0.094T + 53.929$$
$$FI{R_2} ={-} 0.018T + 10.033$$
where FIR1,2 is the intensity ratio and T is the temperature (K). In order to evaluate the performance of temperature sensing, the absolute sensitivity (SA1,2) and relative sensitivity (SR1,2) are displayed in Fig. 4(c, f) and estimated by the following formulas [36]:
$${S_A}_1 = \left|{\frac{{d(FIR)}}{{dT}}} \right|= 9.46\%{K^{ - 1}}$$
$${S_{R1}} = \left|{\frac{1}{{FIR}} \times \frac{{d(FIR)}}{{dT}}} \right|= \frac{{0.09455}}{{ - 0.09455T + 53.92894}} \times 100\%{K^{ - 1}}$$
$${S_{A2}} = \left|{\frac{{d(FIR)}}{{dT}}} \right|= 1.76\%{K^{ - 1}}$$
$${S_{R2}} = \left|{\frac{1}{{FIR}} \times \frac{{d(FIR)}}{{dT}}} \right|= \frac{{0.01758}}{{ - 0.01758T + 10.03374}} \times 100\%{K^{ - 1}}$$

 figure: Fig. 4.

Fig. 4. The integrated intensities of Eu3+: 5D07F1(577-602 nm) and Mn4+: 2Eg4A2g(637-775 nm) (a). Temperature-dependent intensity ratio of Mn4+/Eu3+(695/591) and the fitted curve (b). The relative sensitivity SR (blue line) and the absolute sensitivity SA(red line) as a function of temperature (300–550 K) (c).The integrated intensities of Eu3+ 5D07F2(604-633 nm) and Mn4+ 2Eg4A2g(637-775 nm) (d). Temperature-dependent intensity ratio of Mn4+/Eu3+ (695/615) and the fitted curve (e). The relative sensitivity SR (blue line) and the absolute sensitivity SA (red line) as a function of temperature (300–550 K) (f).

Download Full Size | PDF

 figure: Fig. 5.

Fig. 5. Temperature-dependent decay curves of Mn4+ ions monitored at 695 nm and excited at 302 nm(a), 362 nm(d) and 465 nm (g). The lifetimes as a function of temperature under 302 nm (b), (e) 362 nm and (h) 465 nm excitation. The relative sensitivity SR as a function of temperature based on the lifetimes of Mn4+ upon 302 nm (c), 362 nm(f) and 465 nm (i) excitation.

Download Full Size | PDF

According to Eqs. (36), the absolute (SA) and relative (SR) sensitivities based on FIR of 5D07F1 (577-602 nm)/2Eg4A2g (637-775 nm) and 5D07F2 (604-633 nm)/ 2Eg4A2g (637-775 nm) for the SGLT: 5%Eu3+,0.7%Mn4+ phosphor are calculated to be 9.46%K-1, 4.9%K-1@550 K and 1.76%K-1,4.8%K-1@550 K, respectively. It was worthwhile to mention that these SA and SR values are superior to many other TM3+ / RE3+ codoped systems, such as 4.37% K-1 for Ce3+/Mn4+ doped Lu3Al5O12 [16], 2.08% K-1 for Mn4+/Eu3+ doped Ba2LaNbO6 [25], 1.51% K-1 for Mn4+/Eu3+ doped Ca2LaNbO6 [25], 4.3% K-1 for Mn4+/Tb3+ doped Lu3Al5O12 [37], 0.86% K-1 for Mn4+/Eu3+ doped NaLaMgWO6 [38], and 1.82% K-1 for Mn4+/Dy3+ doped BaLaMgNbO6 [39]. Moreover, the temperature uncertainty (ΔT) is also a crucial parameter to evaluate the temperature sensing of materials, and its value can be evaluated by the following Eq. (7) [40]:

$$\Delta T = \frac{1}{{{S_R}}}\frac{{\delta FIR}}{{FIR}}$$
where δFIR/FIR is the relative error in FIR measurement whose value is approximately 0.5%. The relative sensitivity SR1 and SR2 is 4.9%K-1@550 K and 4.8%K-1@550 K, respectively.

Therefore, the minimum temperature resolution was both evaluated to be about 0.1K at 550 K for SGLT: 5%Eu3+,0.7%Mn4+ phosphor. These results indicate that SGLT: Eu3+,Mn4+ phosphor is a promising temperature sensing material and exhibits a good temperature resolution.

Both the FIR of Mn4+ and Eu3+ and the lifetime of Mn4+ are sensitive to temperature, thus a temperature sensing mode based on lifetime was conducted. Figure 5 (a, d, g) shows the fluorescence decay curves of SGLT: 5%Eu3+, 0.7%Mn4+ sample excited at 302 nm, 362 nm and 465 nm, monitoring the emission of Mn4+ at 695 nm from 300K to 550K. As presented, it was found that under different excitations, electrons are excited from the ground state 4A2g to different excited states, such as 4T1g or 4T2g states. For various excited states, the rate of falling to2Eg is various, resulting in different lifetime of Mn4+ ions. In addition, it was supposed that the lifetime quenching can be assigned to the increment of non-radiative transition with the increase of temperature, which was attributed to the enhancement of electron-phonon coupling originating from no shielding for 3d electrons of Mn4+. Initially, the decay curves can be fitted by a double exponential function at low temperature. With elevating the temperature, the non-exponential decay curves gradually become obvious and the average lifetimes are evaluated by the following equation [41]:

$${\tau _m} = \frac{{\int_0^\infty {t \times I(t)dt} }}{{\int_0^\infty {I(t)dt} }}$$
where I(t) is the luminescence intensity at a time t. It was seen that the rise of temperature results in the continuous decrease of lifetimes, which implies that the resultant phosphor is promising in application for thermometers. As we all know, cross-over model and Struck-Fonger model can be used to describe the relation between lifetime and temperature [20,42]. In our case, the temperature dependence on decay lifetimes can be more accurately fitted using Struck-Fonger model as follows:
$$\tau (T) = \frac{{{\tau _0}}}{{1 + A\exp (\frac{{ - \Delta E}}{{kT}})}}$$
Where τ0 and τ(T) are the initial lifetime at room temperature and a certain temperature T, respectively; k is Boltzmann’s constant (0.69503 cm-1.K-1); A is a constant, and ΔE is the thermal-quenching activation energy.

The lifetime dependence on temperature under different excitation wavelengths can be well fitted by Eq. (8), as shown in Fig. 5(b, e, h). Under 302, 362 and 465 nm excitation, the ΔE were determined to be 0.25(≈ 2016 cm-1), 0.31(≈ 2500 cm-1), and 0.21 eV (≈ 1694 cm-1), respectively, while the corresponding A values were estimated to be 295, 1791.58, and 76.43, respectively. Therefore, the relative temperature sensitivity (SR) derived from decay lifetime dependence on temperature can be expressed as the following equation [43]:

$${S_R} = \left|{\frac{1}{\tau }\frac{{d\tau }}{{dT}}} \right|\times 100\%= \frac{{A\exp ( - \varDelta E/kT)}}{{1 + A\exp ( - \varDelta E/kT)}} \times \frac{{\varDelta E}}{{k{T^2}}} \times 100\%$$
The obtained SR values calculated from Eq. (9) are displayed in Fig. 5(c, f, i). Obviously, the SR values monotonously increase with elevating temperature ranging from 298-573K, the maximum SR is 0.158%, 0.229%, and 0.104% K-1@ 573K, respectively. It is worthwhile mentioning that the SR value can be modulating by changing excitation wavelengths and the SR values were comparable to those of previously based on decay lifetimes, such as ZrO2:Eu3+(0.33%K-1) [19], YAlO3:Cr3+(0.19%K-1) [44], Ca3Mo0.2W0.8O6:Eu3+ (0.142% K-1) [5] and YNbO4:Sm3+ (0.43%K-1) [45]. Hence, these results indicated that SGLT: Eu3+, Mn4+ phosphors are promising candidates for non-contact optical thermometers using FIR and lifetime-based temperature readout.

4. Conclusion

A dual-mode optical thermometry based on FIR of dual-emitting Mn4+, Eu3+ for SGLT phosphor and decay lifetime was constructed. The double perovskite Mn4+, Eu3+ codoped SGLT phosphors were obtained by high-temperature solid-state reaction method. Under 302 nm excitation, the luminescent intensity of Mn4+ dropped faster than that of Eu3+ with the increase of temperature, so the resultant phosphors exhibit an excellent temperature sensing properties due to different thermal behaviors of Mn4+ and Eu3+.The maximum absolute and relative sensitivity based on FIR of 5D07F1 (577-602 nm)/2Eg4A2g (637-775 nm) were as high as 9.46%K-1 and 4.9%K-1@550 K, respectively. Moreover, the decay lifetime of Mn4+ can be used for temperature sensing and the maximum relative sensitivity was 0.229%@ 573K.These results demonstrate that SGLT: Mn4+, Eu3+ phosphor has its promising potential in temperature measurement and provides a guidance for exploring temperature sensors with high sensitivity.

Funding

Chongqing Municipal Education Commission (KJZD-M202000601); National Natural Science Foundation of China (11674044, 11704054, 12004061, 12004062); Natural Science Foundation of Chongqing (CSTC 2019jcyj-msxm1208); Venture and Innovation Support Program for Chongqing Overseas Returnees (CX2019085).

Disclosures

The authors declare no conflicts of interest.

See Supplement 1 for supporting content.

References

1. X. Zhang, Z. Zhu, Z. Guo, Z. Sun, and Y. Chen, “A ratiometric optical thermometer with high sensitivity and superior signal discriminability based on Na3Sc2P3O12:Eu2+, Mn2+ thermochromic phosphor,” Chem. Eng. J. 356, 413–422 (2019). [CrossRef]  

2. C. Xie, P. Wang, Y. Lin, X. Wei, M. Yin, and Y. Chen, “Temperature-dependent luminescence of a phosphor mixture of Li2TiO3:Mn4+ and Y2O3:Dy3+ for dual-mode optical thermometry,” J. Alloys Compd. 821, 153467 (2020). [CrossRef]  

3. F. Chi, B. Jiang, Z. Zhao, Y. Chen, X. Wei, C. Duan, M. Yin, and W. Xu, “Multimodal temperature sensing using Zn2GeO4:Mn2+ phosphor as highly sensitive luminescent thermometer,” Sens. Actuators, B 296, 126640 (2019). [CrossRef]  

4. J. Rocha, C. D. S. Brites, and L. D. Carlos, “Lanthanide organic framework luminescent thermometers,” Chem. Eur. J. 22(42), 14782–14795 (2016). [CrossRef]  

5. J. Xue, H. M. Noh, B. C. Choi, S. H. Park, J. H. Kim, J. H. Jeong, and P. Du, “Dual-functional of non-contact thermometry and field emission displays via efficient Bi3+→Eu3+ energy transfer in emitting-color tunable GdNbO4 phosphors,” Chem. Eng. J. 382, 122861 (2020). [CrossRef]  

6. M. Xu, X. Zou, Q. Su, W. Yuan, C. Cao, Q. Wang, X. Zhu, W. Feng, and F. Li, “Ratiometric nanothermometer in vivo based on triplet sensitized upconversion,” Nat. Commun. 9(1), 2698 (2018). [CrossRef]  

7. S. S. Zhou, C. K. Duan, M. Yin, X. L. Liu, S. Han, S. B. Zhang, and X. M. Li, “Optical thermometry based on cooperation of temperature-induced shift of charge transfer band edge and thermal coupling,” Opt. Express 26(21), 27339 (2018). [CrossRef]  

8. G. Xiang, X. Liu, W. Liu, B. Wang, Z. Liu, S. Jiang, X. Zhou, L. Li, Y. Jin, and J. Zhang, “Multifunctional optical thermometry based on the Stark sublevels of Er3+ in CaO- Y2O3:Yb3+/Er3+,” J. Am. Ceram. Soc. 103(4), 2540–2547 (2020). [CrossRef]  

9. J. Zhong, D. Chen, Y. Peng, Y. Lu, X. Chen, X. Li, and Z. Ji, “A review on nanostructured glass ceramics for promising application in optical thermometry,” J. Alloys Compd. 763, 34–48 (2018). [CrossRef]  

10. L. Luo, W. Ran, P. Du, W. Li, and D. Wang, “Photocatalytic and thermometric characteristics of Er3+-activated Bi5IO7 upconverting microparticles,” Adv. Mater. Interfaces 7(11), 1902208 (2020). [CrossRef]  

11. M. Back, J. Ueda, M. G. Brik, T. Lesniewski, M. Grinberg, and S. Tanabe, “Revisiting Cr3+-doped Bi2Ga4O9 spectroscopy:Crystal field effect and optical thermometric behavior of near-infrared-emitting singly-activated phosphors,” ACS Appl. Mater. Interfaces 10(48), 41512–41524 (2018). [CrossRef]  

12. Y. Cheng, Y. Gao, H. Lin, F. Huang, and Y. S. Wang, “Strategy design for ratiometric luminescence thermometry: Circumventing the limitation of thermally coupled levels,” J. Mater. Chem. C 6(28), 7462–7478 (2018). [CrossRef]  

13. J. Xue, X. Wang, J. H. Jeong, and X. Yan, “Spectral and energy transfer in Bi3+-Ren+ (n = 2, 3, 4) co-doped phosphors: extended optical applications,” Phys. Chem. Chem. Phys. 20(17), 11516–11541 (2018). [CrossRef]  

14. A. M. Kaczmarek, J. Liu, B. Laforce, L. Vincze, K. V. Hecke, and R. V. Deun, “Cryogenic luminescent thermometers based on multinuclear Eu3+/Tb3+ mixed lanthanide polyoxometalates,” Dalton Trans. 46(18), 5781–5785 (2017). [CrossRef]  

15. T. Hu, Y. Gao, M. Molokeev, Z. Xia, and Q. Zhang, “Non-stoichiometry in Ca2Al2SiO7 enabling mixedvalent europium toward ratiometric temperature sensing,” Sci. China Mater. 62(12), 1807–1814 (2019). [CrossRef]  

16. Y. Chen, J. He, X. Zhang, M. Rong, Z. Xia, J. Wang, and Z. Q. Liu, “Dual-mode optical thermometry design in Lu3Al5O12:Ce3+/Mn4+ phosphor,” Inorg. Chem. 59(2), 1383–1392 (2020). [CrossRef]  

17. M. Sekulić, V. Đorđević, Z. Ristić, and M. Medić, “Highly sensitive dual self-referencing temperature readout from the Mn4+/Ho3+ binary luminescence thermometry probe,” Adv. Opt. Mater. 6(17), 1800552 (2018). [CrossRef]  

18. Y. Song, N. Guo, J. Li, R. Ouyang, Y. Miao, and B. Shao, “Photoluminescence and temperature sensing of lanthanide Eu3+ and transition metal Mn4+ dual-doped antimoniate phosphor through site-beneficial occupation,” Ceram. Int. 46(14), 22164–22170 (2020). [CrossRef]  

19. J. Zhou, R. Lei, H. Wang, Y. Hua, D. Li, Q. Yang, D. Deng, and S. Q. Xu, “A new generation of dual-mode optical thermometry based on ZrO2:Eu3+ nanocrystals,” Nanophotonics 8(12), 2347–2358 (2019). [CrossRef]  

20. M. D. Dramićanin, B. Milićević, V. Đorđević, Z. Ristić, J. Zhou, D. Milivojević, J. Papan, M. G. Brik, C. G. Ma, A. M. Srivastava, and M. M. Wu, “Li2TiO3:Mn4+ deep-red phosphor for the lifetime-based luminescence thermometry,” ChemistrySelect 4(24), 7067–7075 (2019). [CrossRef]  

21. Z. Fang, L. Zhao, Q. Yang, Z. Yang, Y. Cai, D. Zhou, X. Yu, J. Qiu, and X. Xu, “Optical thermometry properties of silicate glass ceramics with dual-phase for spatial isolation of Er3+ and Cr3+,” J. Lumin. 219, 116861 (2020). [CrossRef]  

22. D. Chen, S. Liu, Y. Zhou, Z. Wan, P. Huang, and Z. Ji, “Dual-activator luminescence of RE/TM:Y3Al5O12 (RE = Eu3+,Tb3+,Dy3+;TM = Mn4+,Cr3+) phosphors for self-referencing optical thermometry,” J. Mater. Chem. C 4(38), 9044–9051 (2016). [CrossRef]  

23. L. Chu, W. Ahmad, W. Liu, J. Yang, R. Zhang, Y. Sun, J. Yang, and X. Li, “Lead-free halide double perovskite materials: A new superstar toward green and stable optoelectronic applications,” Nano-Micro Lett. 11(1), 16 (2019). [CrossRef]  

24. W. Ning and F. Gao, “Structural and functional diversity in lead-free halide perovskite materials,” Adv. Mater. 31(22), 1900326 (2019). [CrossRef]  

25. P. Wang, J. Mao, L. Zhao, B. Jiang, C. Xie, Y. Lin, F. Chi, M. Yin, and Y. Chen, “Double perovskite A2LaNbO6:Mn4+, Eu3+ (A = Ba, Ca) phosphors: potential applications in optical temperature sensing,” Dalton Trans. 48(27), 10062–10069 (2019). [CrossRef]  

26. Y. Lin, L. Zhao, B. Jiang, J. Mao, F. Chi, P. Wang, C. Xie, X. Wei, Y. Chen, and M. Yin, “Temperature-dependent luminescence of BaLaMgNbO6: Mn4+, Dy3+ phosphor for dual-mode optical thermometry,” Opt. Mater. 95, 109199 (2019). [CrossRef]  

27. M. L. López, I. Alvarez, M. Gaitdn, A. Jerez, C. Pico, and M. L. Veiga, “Structural study and magnetic measurements of some perovskites MLnLiTeO6,” Solid State Ionics 63-65, 599–602 (1993). [CrossRef]  

28. V. L. Vales and G. Subodh, “Crystal structure, phonon modes and dielectric properties of B site ordered ABiLiTeO6 (A = Ba, Sr) double perovskites,” Ceram. Int. 44(11), 12036–12041 (2018). [CrossRef]  

29. W. Zhao, Z. Y. Ping, Q. H. Zheng, and W. W. Zhou, “Concentration and thermal quenching of SrGdLiTeO6:Eu3+ red-emitting phosphor for white light-emitting diode,” Acta. Phys. Sin. 67(24), 247801 (2018). [CrossRef]  

30. S. C. Lal, A. M. Aiswarya, K. S. Sibi, and G. Subodh, “Insights into the structure, photoluminescence and Judd-Ofelt analysis of red emitting SrLaLiTeO6:Eu3+ phosphors,” J. Alloys Compd. 788, 1300–1308 (2019). [CrossRef]  

31. L. Li, S. Fu, Y. Zheng, C. Li, P. Chen, G. Xiang, S. Jiang, and X. Zhou, “Near-ultraviolet and blue light excited Sm3+ doped Lu2MoO6 phosphor for potential solid state lighting and temperature sensing,” J. Alloys Compd. 738, 473–483 (2018). [CrossRef]  

32. L. Li, X. Tang, Z. Wu, Y. Zheng, S. Jiang, X. Tang, G. Xiang, and X. Zhou, “Simultaneously tuning emission color and realizing optical thermometry via efficient Tb3+(Eu3+ energy transfer in whitlockite-type phosphate multifunctional phosphors,” J. Alloys Compd. 780, 266–275 (2019). [CrossRef]  

33. K. Li, H. Lian, R. V. Deun, and M. G. Brik, “A far-red-emitting NaMgLaTeO6:Mn4+ phosphor with perovskite structure for indoor plant growth,” Dyes Pigments 162, 214–221 (2019). [CrossRef]  

34. L. Li, W. Chang, W. Chen, Z. Feng, C. Zhao, P. Jiang, Y. Wang, X. Zhou, and A. Suchocki, “Double perovskite LiLaMgWO6:Eu3+ novel red-emitting phosphors for solid state lighting: Synthesis, structure and photoluminescent properties,” Ceram. Int. 43(2), 2720–2729 (2017). [CrossRef]  

35. Y. Huang, Q. Sun, and B. Devakumar, “Novel efficient deep-red-emitting Ca2LuTaO6: Mn4+ double-perovskite phosphors for plant growth LEDs,” J. Lumin. 222, 117177 (2020). [CrossRef]  

36. M. D. Dramićanin, “Trends in luminescence thermometry,” J. Appl. Phys. 128(4), 040902 (2020). [CrossRef]  

37. C. Xie, L. Zhao, B. Jiang, J. Mao, Y. Lin, P. Wang, X. Wei, M. Yin, and Y. Chen, “Dual-activator luminescence of LuAG:Mn4+/Tb3+ phosphor for optical thermometry,” J. Am. Ceram. Soc. 102(12), 7500–7508 (2019). [CrossRef]  

38. H. Zhou, N. Guo, M. Zhu, J. Li, Y. Miao, and B. Shao, “Photoluminescence and ratiometric optical thermometry in Mn4+/Eu3+ dual-doped phosphor via site-favorable occupation,” J. Lumin. 224, 117311 (2020). [CrossRef]  

39. Y. Lin, L. Zhao, B. Jiang, J. Mao, F. Chi, P. Wang, C. Xie, X. Wei, Y. Chen, and M. Yin, “Temperature-dependent luminescence of BaLaMgNbO6:Mn4+, Dy3+ phosphor for dual-mode optical thermometry,” Opt. Mater. 95, 109199 (2019). [CrossRef]  

40. L. P. Li, F. Qin, Y. Zhou, Y. Zheng, J. Miao, and Z. G. Zhang, “Photoluminescence and time-resolved-luminescence of CaWO4:Dy3+ phosphors,” J. Lumin. 224, 117308 (2020). [CrossRef]  

41. J. Xue, M. Song, H. M. Noh, S. H. Park, B. C. Choi, J. H. Kim, J. H. Jeong, and P. Du, “Achieving non-contact optical thermometer via inherently Eu2+/Eu3+-activated SrAl2Si2O8 phosphors prepared in air,” J. Alloys Compd. 843, 155858 (2020). [CrossRef]  

42. W. H. Fonger and C. W. Struck, “Eu+35D resonance quenching to the charge transfer states in Y2O2S, La2O2S, and LaOCl,” J. Chem. Phys. 52(12), 6364–6372 (1970). [CrossRef]  

43. T. Senden, R. J. A. van Dijk-Moes, and A. Meijerink, “Quenching of the red Mn4+ luminescence in Mn4+-doped fluoride LED phosphors,” Light: Sci. Appl. 7(1), 8 (2018). [CrossRef]  

44. H. Uchiyama, H. Aizawa, T. Katsumata, S. Komuro, T. Morikawa, and E. Toba, “Fiber-optic thermometer using Cr-doped YAlO3 sensor head,” Rev. Sci. Instrum. 74(8), 3883–3885 (2003). [CrossRef]  

45. L. R. Đačanin, S. R. Lukić-Petrović, D. M. Petrović, M. G. Nikolić, and M. D. Dramićanin, “Temperature quenching of luminescence emission in Eu3+-and Sm3+-doped YNbO4 powders,” J. Lumin. 151, 82–87 (2014). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Figure S1

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 (5)

Fig. 1.
Fig. 1. XRD patterns of SGLT, SGLT: 0.7%Mn4+, SGLT: 5%Eu3+ and SGLT: 5%Eu3+, 0.7%Mn4+ compared with ICSD no. 067852 of SrEuLiTeO6 (a).The crystal structure of SGLT (b).
Fig. 2.
Fig. 2. Excitation and emission spectra of SGLT: 5%Eu3+ (a), SGLT: 0.7%Mn4+ (b) and SGLT: 5%Eu3+,0.7%Mn4+ (c) phosphors.
Fig. 3.
Fig. 3. Temperature-dependent emission spectra (a) and two-dimensional fluorescence topographical mapping (b) of SGLT:5%Eu3+,0.7%Mn4+ under 302 nm excitation in the temperature range from 300 K to 550 K.
Fig. 4.
Fig. 4. The integrated intensities of Eu3+: 5D07F1(577-602 nm) and Mn4+: 2Eg4A2g(637-775 nm) (a). Temperature-dependent intensity ratio of Mn4+/Eu3+(695/591) and the fitted curve (b). The relative sensitivity SR (blue line) and the absolute sensitivity SA(red line) as a function of temperature (300–550 K) (c).The integrated intensities of Eu3+ 5D07F2(604-633 nm) and Mn4+ 2Eg4A2g(637-775 nm) (d). Temperature-dependent intensity ratio of Mn4+/Eu3+ (695/615) and the fitted curve (e). The relative sensitivity SR (blue line) and the absolute sensitivity SA (red line) as a function of temperature (300–550 K) (f).
Fig. 5.
Fig. 5. Temperature-dependent decay curves of Mn4+ ions monitored at 695 nm and excited at 302 nm(a), 362 nm(d) and 465 nm (g). The lifetimes as a function of temperature under 302 nm (b), (e) 362 nm and (h) 465 nm excitation. The relative sensitivity SR as a function of temperature based on the lifetimes of Mn4+ upon 302 nm (c), 362 nm(f) and 465 nm (i) excitation.

Equations (10)

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

F I R 1 = 0.094 T + 53.929
F I R 2 = 0.018 T + 10.033
S A 1 = | d ( F I R ) d T | = 9.46 % K 1
S R 1 = | 1 F I R × d ( F I R ) d T | = 0.09455 0.09455 T + 53.92894 × 100 % K 1
S A 2 = | d ( F I R ) d T | = 1.76 % K 1
S R 2 = | 1 F I R × d ( F I R ) d T | = 0.01758 0.01758 T + 10.03374 × 100 % K 1
Δ T = 1 S R δ F I R F I R
τ m = 0 t × I ( t ) d t 0 I ( t ) d t
τ ( T ) = τ 0 1 + A exp ( Δ E k T )
S R = | 1 τ d τ d T | × 100 % = A exp ( Δ E / k T ) 1 + A exp ( Δ E / k T ) × Δ E k T 2 × 100 %
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.