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

Enhanced non-contact optical temperature sensing performance based on upconversion luminescence in Ho3+/Yb3+ codoped ferroelectric 0.94Na0.5Bi0.5TiO3-0.06BaTiO3

Open Access Open Access

Abstract

The upconversion luminescence of ferroelectric 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: xHo3+/0.01Yb3+ is investigated in detail. The prepared samples show intense characteristic emission from Ho3+ ions with a dominant green emission peak located at 547 nm under 980 nm excitation. Among these solid solutions, the 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ ceramic possesses the optimum emission. Optical temperature sensing behavior of Ho3+/Yb3+ codoped ferroelectric 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 are investigated through fluorescence intensity ratio (FIR, I656/I547) induced by non-thermally coupled energy levels 5F55I8 and 5F4/5S25I8 in the range of 303–483 K. A moderate amount of Yb3+ doping in 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: Ho3 + ceramic contributes to the improvement of their optical temperature sensing performance. At T = 303 K, the calculated highest relative sensitivity SR is 0.0042 K-1, which indicates that the optimized 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ ceramic is promising as the candidate materials for noncontact optical temperature sensing devices.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

In recent years, for Lead-based piezoelectric ceramics used as an important class of smart materials, the dominant one is still conventional Pb(ZrxTi1-x)O3 (PZT) due to its excellent piezoelectric performance, which plays a crucial role in practical applications, such as transducers, actuators, sensors and other electro-mechanical devices and systems [1,2]. However, some countries have enacted restrictions prohibiting the consumption of in electronic devices containing lead due to their hazardous and volatile nature in the manufacturing operation, posing a significant risk to human health and the environment [3,4]. As a result of the growing awareness of environmental protection and long-term development, lead-free piezoelectric materials for replacing conventional lead-containing piezoelectric materials have received increased attention [58].

Na0.5Bi0.5TiO3 is recognized as one of the most promising lead-free piezoelectric material possibilities. Raman results show that the perovskite ceramic with a highly crystalline-structure has a lower phonon energy [9]. In addition, the material Na0.5Bi0.5TiO3 has a high Curie temperature (Tc = 320 °C) [10]. This property shows that Na0.5Bi0.5TiO3 ceramic has a potential applicability in high temperature applications. Thus, Na0.5Bi0.5TiO3-based ceramics can serve as an excellent host material for photoluminescence, which drops the probability of non-radiative transitions caused by phonon vibrations and increases the activators up-conversion emission luminous intensity. However, the piezoelectric properties of the material Na0.5Bi0.5TiO3 are not ideal, limiting its practical application. To improve its piezoelectric properties, a new one or more solid solutions with perovskite structure are usually incorporated into Na0.5Bi0.5TiO3, forming new binary or multiple solid solutions. Among them, 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 has been extensively studied as a binary system with a composition near the morphological phase boundary (MPB) for the optimum piezoelectric performance [11].

It is of great interest that rare earth (RE) ion-activated 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 ferroelectrics not only exhibit additional functionality of excellent photoluminescence, and improved intrinsic ferroelectric and piezoelectric properties [12], but their photoluminescence can also be efficiently modulated by an electric field based on a physical approach [13,14], which is usually reversible. There are numerous uses for these materials in the discipline of photonics, which can expand the applications of piezoelectric nanogenerators [15]. Where photoluminescence associated to RE ions serving as a straightforward yet effective method to probe their local structure of the MPB composition [16]. Furthermore, Na0.5Bi0.5TiO3, BaTiO3 and (1-y)Na0.5Bi0.5TiO3-yBaTiO3 consisting of both activated RE ions show high up-conversion emission, and the degree of their upconversion luminescence is significantly influenced by the ambient temperature [1719].

Upconversion luminescence can generate short-wavelength higher-energy photons by absorbing two or more long-wavelength lesser-energy photons, a phenomenon known as the anti-Stokes shift [20,21]. The upconversion luminescence phenomenon is a very important and fundamental nonlinear optical process. Upconversion luminescence materials have attracted widespread interest because of their potential applications in medicinal chemistry, anti-counterfeiting, photovoltaic technologies, photocatalytic technique, displays, optical fibers, temperature sensing, and optoelectronic devices [2125]. Furthermore, temperature sensing materials can be employed in luminescent probes and thermal sensors [26,27].

For upconversion luminescence materials, trivalent Er3+, Ho3+, but also Tm3+ are the most commonly utilized as activator ions because they have special ladder-like arranged metastable energy levels [28]. To improve the fluorescence intensity in upconversion luminescence process associated to Er3+, Ho3+ and Tm3+ ions, Yb3+ is frequently adopted as a sensitizer due to their intuitive energy level configuration and large absorption cross section in the near-infrared (NIR) region, which can enhance energy absorption [29]. Subsequently, the upconversion luminescence intensity and sensing sensitivity are enhanced by effectively transferring energy from Yb3+ to these RE3+ ions. Notably, Ho3+ is among the few RE3+ with a large amount of long-lived metastable energy levels capable of photoluminescence with intense emission across a broad spectral region spanning ultraviolet (UV)-to-mid-infrared (IR) [30]. For example, Ho3+ is one of the most efficient luminescent ions for its abundant energy level transition processes producing strong visible light, e.g., green and red emissions, accompanied by relatively weak NIR emission under 980 nm excitation. Over the past few years, some researches on lead free ferroelectrics doped with Ho3+ or Ho3+/Yb3+ have been carried out since their outstanding optical temperature sensing performance [31], for instance, Na0.5Bi0.5TiO3: Ho3+ [17], (K0.47Na0.47Li0.06)(Nb0.94Bi0.06)O3: Ho3+ [32], (K, Na)NbO3–SrTiO3: Ho3+ [33] and BaTiO3: Ho3+/Yb3+ [21], and so on.

As a method of studying optical temperature sensing, the fluorescence intensity ratio (FIR) technique can avert several external disturbances in the measurement process to achieve high measurement sensing accuracy and sensitivity [34,35]. In order to further enhance the sensitivity, the FIR originated from non-thermally coupled energy levels (non-TCLs) in RE3+ activated upconversion luminescence materials with respect to temperature sensing strategy has recently attracted more and more attention [36,37]. The significant reason for this is that the emission bands induced via the non-TCLs are separated in the spectra, i.e., they do not intersect compared with the TCLs.

As a result, Ho3+/Yb3+ codoped 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 ceramics are speculated to be ideal materials with efficient upconversion luminescence for optical temperature sensing applications. In the present work, specimens of 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: Ho3+/Yb3+ were prepared. The Ho3+ doping content on crystal structure as well as upconversion luminescence characteristics were investigated in detail. Additionally, based on upconversion luminescence spectra, the temperature from 303 K to 483 K independent FIR (I656/I547) generated by non-thermally coupled levels, i.e., the red (5F55I8) emission band to the green (5F4/5S25I8), were studied. Meanwhile, the relative sensing sensitivity exists a highest value of 0.0042 K-1 at 303 K, which successfully achieves a 20% improvement over Ho3+ mono-doped 0.94Na0.5Bi0.5TiO3-0.06BaTiO3.

2. Experimental

The details of the preparation process originated to the polycrystalline lead-free ceramics with the following chemical formula 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: xHo3+/0.01Yb3+ (x = 0.0025, 0.005, 0.0075, 0.01, 0.0125, 0.015) can be found in Ref. [31]. The final sintered samples are all bulk ceramics. To assess the crystal structures of the sintered samples, a powder PANalytical X'pert3 X-ray diffractometer (Netherlands) with CuKα (λ=0.15406 nm) as the radiation resource was used to characterized X-ray diffraction (XRD) patterns. The surface morphology and element mapping were performed via a Quanta FEG450 field emission scanning electron microscope (FESEM, USA). The diffuse reflection spectra (DRS) were recorded using a Shimadzu UV-3600 UV/visible spectrophotometer (Japan). The content and surrounding temperature dependent emission spectrums of the as-prepared ceramics upon the excitation of 980 nm were collected in a Fluorlog-3 spectrofluorometer (France). The spectra measurement has a wavelength accuracy of 0.5 nm, the pulsed excitation was accomplished with a nanosecond laser diode (LD) with adjustable power. The sample was put on a heating device whose temperature could be adjusted from ambient temperature to 573 K, achieving upconversion luminescence spectra measurement at different temperatures.

3. Result and discussion

Based on room-temperature XRD diffraction patterns of 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: xHo3+/0.01Yb3+ ceramics, a single-phase perovskite-structure is discovered as presented in Fig. 1(a). All observed characteristic peaks match with the pure rhombohedral NBT (PDF No.36-0340, R3c), free of impurity peak in the XRD detection resolution range. Figure 1(b)-(d) give the amplified XRD profiles with a 2θ in range of 31.75°-47.5° to further explain phase structure of the polycrystalline. The diffraction peaks shift observed from Fig. 1(b)-(d) are slightly displaced to higher angles when the x-value (content of Ho3+ ions) increases. Ho3+ and Yb3+ ions are most likely to substitute Bi3+ ions at the A-site, both in terms of ionic radii [38] and valence considerations, followed by the lattice shrinkage phenomenon described above. Similar to our earlier research [31], in the XRD profiles of 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: xHo3+/0.01Yb3+ ceramics, as shown in Fig. 1(b) and (c), two prominent splitting peaks near 32.50° and 40.10° indicate that the rhombohedral phase is present in the as-prepared ceramics. Figure 1(d) depicts the variation of the (202) reflections that are split into doublets, suggesting the tetragonal phase is also present [39,40], which is similar to 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 matrix. These results indicate the appearance of MPB in ceramics containing both rhombohedral and tetragonal phases.

 figure: Fig. 1.

Fig. 1. a) Room-temperature XRD diffraction patterns related to 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: xHo3+/0.01Yb3+ ceramics. b)-d) are the magnified XRD patterns from 31.75° to 33.25°, 39.25° to 40.75° and 46° to 47.5° in the 2θ range, respectively.

Download Full Size | PDF

Figure 2(a) gives SEM images for the prepared ceramic with x = 0.0075. Apparently, the distribution of grain size is inhomogeneous. The present ceramics have a compact microstructure and good sinterability. Figure 2(b) shows EDS spectrum of 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ ceramic and Fig. 2(c-i) present its EDS elemental mapping images. Figure 2(b) demonstrates that the presence of Na, Ba, Yb, Ho, Bi, Ti and O. As shown in Fig. 2(c-i), all constituent elements of this compound within a specific area can be determined using the EDS elemental mappings. It is confirmed that the 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ ceramic include the all components, i.e., Na, Ba, Yb, Ho, Bi, Ti and O, and that these components are dispersed randomly all through the ceramic particles.

 figure: Fig. 2.

Fig. 2. a) Scanning electron microscope (SEM) images and b) EDS spectrum of 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ ceramic. c)-i) corresponding EDS elemental mapping images of Na, Ba, Yb, Ho, Bi, Ti and O in 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ ceramic, respectively.

Download Full Size | PDF

Figure 3 gives the UV-visible diffuse reflectance spectra (DRS, 200–800 nm) of the 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: xHo3+/0.01Yb3+ ceramics including x = 0.0075 and 0.015. The DRS displays a high reflectance (>80%) between 450 and 800 nm. Three distinct absorption bands centered at 453, 540 and 644 nm are observed in the DRS related to those two typical samples, corresponding to 4f-4f transitions of Ho3+ ions, that is 5I85G6, 5I85F4/5S2 and 5I85F5, respectively. In addition, at 485 nm, a somewhat faint absorption band is observed corresponding to the 5I85F3 transition [41]. A sharp drop near 360 nm is observed, corresponding to the absorption edge that determines its optical band gap. Below 340 nm, there is a wide absorption band induced by the electrons transition from the valence band (VB) to the conduction band (CB) [42].

 figure: Fig. 3.

Fig. 3. The UV–visible diffuse reflectance spectra (DRS) of typical 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: xHo3+/0.01Yb3+ ceramics (x = 0.0075 and 0.015). The inset exhibits the corresponding absorption spectra.

Download Full Size | PDF

To experimentally determine the band gap energy (forbidden band width) of typical 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: xHo3+/0.01Yb3+ ceramic, the absorption spectra are shown in the inset of Fig. 3. These spectra were derived from their UV–visible DRS using the Kubelka-Munk (K-M) function $F(R)$ [43], as provided in Eq. (1)

$$F(R) = \frac{{{{(1 - R)}^2}}}{{2R}} = \frac{K}{S}$$

With regard to the function, R(=Rsample/Rstandard) is the diffuse reflectance. S denote scattering coefficients, K is proportional to S. When the powder sample scatters in completely diffuse manner or when it is continuously illuminated at 60 ° incidence, then $K = 2\alpha $, $\alpha $ is linear absorption coefficient, S stands for a constant in relation to the wavelength of the incident light [31]. According to earlier publications [44,45], the Wood-Tauc relation can be used to describe the relationship of α and Eg for materials:

$$\mathrm{\alpha }hv\textrm{ = C}{({hv\textrm{ - }{E_\textrm{g}}} )^n}$$
Where =photon energy, Eg is optical band gap energy, C represents a constant termed as band tailoring parameter, n is dependent on the band gap type associated to electronic transitions, with the allowed direct being 1/2 and the allowed indirect being 2. As presented in the inset of Fig. 3, the optical band gap Eg is acquired through extrapolating the linear part of the circle with [F(R)]2 against to intersect the horizontal coordinate axis. The typical samples’ optical band gap values are 3.23 and 3.21 eV for Ho3+ ion contents of x = 0.0075 and 0.015, respectively.

Figure 4(a) displays the upconversion emission spectra of Ho3+/Yb3+ codoped 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 binary system when it is stimulated by a 980 nm laser diode fitted with a power of about 0.1W. As shown in Fig. 4(a), the sintered ceramics display strong upconversion emissions in the 450–800 nm range. In their upconversion emission spectra, there are three distinct bands of Ho3+ in the green (520-575 nm), red (625-690 nm) and NIR (730-775 nm) regions, are ascribed to the radiative transitions of Ho3+: 5F4/5S25I8, 5F55I8 and 5F4/5S25I7, respectively. When the content of Ho3+ ions increase, the peak positions of all emission bands remain almost unchanged, but the emission intensities change significantly. To further determine the dependence of the intensity related to the upconversion luminescence spectra on the amount of Ho3+ ions, the main peak intensities for different Ho3+ doping concentration located at different positions are shown in Fig. 4(b). The emission intensity increases initially, reaching the maximum when x = 0.0075, and then declines with increasing Ho3+ content. Therefore, the ideal Ho3+ doping concentration for the 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: xHo3+/0.01Yb3+ system is 0.0075. The concentration quenching effect takes place as Ho3+ content is greater than 0.0075, which is due to the fact that the distance between two adjoining Ho3+ and Ho3+ or Ho3+ and Yb3+ decreases with increasing Ho3+ content, which facilitates the non-radiative energy transfer (ET) between them and thus results in a reduction in the upconversion luminescence intensity [46].

 figure: Fig. 4.

Fig. 4. a) The upconversion emission spectrums dependent on Ho3+ ion composition in the 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: xHo3+/0.01Yb3+ ferroelectric system, b) The upconversion emission intensity of Ho3+ ions in various wavelength ranges with the different content of Ho3+.

Download Full Size | PDF

Figure 5(a) depicts the upconversion emission spectra of Ho3+ single-doped and Ho3+/Yb3+ co-doped 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 ceramics. From Fig. 5(a), we found that the introduction of appropriate Yb3+ ions greatly enhance the luminous intensity. A similar upconversion emission spectral pattern to that the one containing only Ho3+ ions is observed, which is consistent with the description in the introduction.

 figure: Fig. 5.

Fig. 5. a) When x = 0.0075, upconversion emission spectra with and without Yb3+ at pump power 0.5 W. The inset shows enlarged upconversion emission spectrum in the Yb3+ free ceramic. b) The upconversion luminescence mechanism of Ho3+/Yb3+ codoped 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 specimen.

Download Full Size | PDF

According to the upconversion luminescence spectrums of 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: xHo3+/0.01Yb3+ previously presented in Fig. 4(a) above, Fig. 5(b) gives an energy level diagram about the potential ET between Ho3+ and Yb3+ ions, resulting in upconversion luminescence processes involving possible excitation and emission paths [47], which contributes to the understanding of the upconversion luminescence mechanism. Though ground state absorption (GSA), Yb3+ ions are first excited by a 980 nm laser diode from the 2F7/2 ground state to the 2F5/2 excited state. Subsequently, an ET occurs from the excited Yb3+ to the adjacent Ho3+, which is main reason for the population of Ho3+ ions in excited levels due to the absorption cross-section of Yb3+ acted as efficient sensitizers being larger than its activators Ho3+ for 980 nm photons. And then, the upward 5I85I6, 5I75F5 and 5I65F4/5S2 transitions associated to Ho3+ ions are promoted. In which, the population on 5F4/ 5S2 level is raised through two successive ETs from Yb3+.

And then, the majority of the Ho3+ ions in the 5F4/5S2 levels radiatively transiton to the 5I8 ground state energy level, generating a green emission peak with a 547 nm center. Simultaneously, a tiny proportion of Ho3+ ions relax to the 5I7 level (5F4/5S25I7), creating a 756 nm-centered NIR emission band. In addition, there are two mechanisms that account for the population of Ho3+ in 5F5 level. The first one is the non-radiative multi-phonon relaxation (marked as ①) from 5F4/5S2 levels. The second one is the non-radiatively decay of Ho3+ ions from their 5I6 excited state (marked as ②) to the lower level 5I7. Then, through excited state absorption (ESA), Ho3+ ions in 5I7 absorb the energy from Yb3+ or a 980 nm photon and transition to their 5F5 level. The majority of Ho3+ ions in the 5F5 level radiatively transition to the ground state 5I8 level, resulting in a red emission. [47,48]

To study the mechanism and process of upconversion luminescence, it is necessary to determine the relationship between the pump power and its corresponding luminous intensity of the luminescent material. The upconversion luminescence spectra of the prepared sample with the best upconversion luminescence performance in the 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: xHo3+/0.01Yb3+ system with 980 nm excitation working at different laser pumping powers (310-570 mW) are shown in Fig. 6(a). The intensity of the emissions increases significantly with increasing pump power. In upconversion luminescence mechanism, the following relationship formula can be used to express the intensity I as a function of the pump power P [49]:

$$I \propto {P^n}$$
where n denotes the number of pump photons required to emit one photon during the upconversion emission process. As presented in Fig. 6(b), n can be obtained from a bilogarithmic plot of the integrated intensity of visible emissions with different input powers. Plots within the pump-power range exhibit a linear trend. In accordance with the linear fitting results, the slopes n for green 547 nm, red 656 nm and NIR 756 nm emissions have values 1.89, 1.91 and 1.95, respectively, indicating that two-photon absorption process is engaged in these upconversion emission mechanisms.

 figure: Fig. 6.

Fig. 6. a) Pump power dependence of the upconversion luminescence spectra of the 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ sample. b) Ln(Power) dependent Ln (Intensity) for 547, 656, and 756 nm.

Download Full Size | PDF

The upconversion luminescence performances of luminescent materials are usually sensitive to the temperature, for example color rendering and light yield, which make them to be applied to optical thermometry. To study the temperature characteristics of the upconversion luminescence in 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ ceramics, Fig. 7 displays the measured upconversion luminescence spectra at different temperatures (303-483 K), with excitation at 980 nm. The luminous intensity drops monotonically with rising temperature. Because of the interaction between crystal-field and 4f energy levels of rare earth (RE) ions, the PL performance of RE3+-doped phosphors is hypersensitive to the local symmetry of ligand environment where the luminous active ions reside [50]. The maximum temperature is 483 K, which is below the Curie temperature reported [51]. Therefore, fluctuations in luminous intensity caused by structural phase changes are excluded during the test at variable temperatures.

 figure: Fig. 7.

Fig. 7. The upconversion luminescence spectra associated to the 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.075Ho3+/0.01Yb3+ ceramic at various temperatures. The inset displays the emission intensity normalized to the emission intensity at room temperature T = 303 K.

Download Full Size | PDF

The upconversion luminescence spectra consists of three bands, namely an intense green due to the 5F4/5S25I8 transition, a red band attributed to 5F55I8 and a weaker NIR band assigned to 5F4/5S25I7 transition. Their emission intensity at different temperature normalized by the intensity at T = 303 K are shown in the inset of Fig. 7. With increasing temperature (303–483 K), the decline evolutions of normalized intensity associated with these three emission bands are very similar, which is explained to thermal quenching effects. When the ceramic is heated, the lattice vibrations increase, increasing the likelihood of non-radiative multi-phonon relaxation (MPR) processes. Therefore, the upconversion luminescence intensity derived from the radiative transition weakens, and fluorescence quenching phenomenon occurs as temperature rises.

In contrast to red light emission (656 nm), the normalized emission intensity of green light emission (547 nm) and NIR emission (756 nm) decreases rapidly with rising temperature. This result can be explained by the fact that the probability of non-radiative relaxation of Ho3+ ions between the 5F4/5S2 and 5F5 levels (①) may increase. This would result in a decrease in the number of electrons on the 5F4/5S2 levels of Ho3+ and an increase in the electronic populations on 5F5 level. Therefore, the normalized intensity of red emission with a center at 656 nm declines more gradually with increasing temperature than the normalized intensity of the green emission with a center at 547 nm [52]. I656 as well as I547 stand for the integral intensities of luminous bands with centers at 656 nm and 547 nm, respectively. As shown in Fig. 8(a), the FIR (I656/I547) increases as temperature rises. At T = 483 K, the upconversion luminescence intensity of green, red and NIR light is only 11%, 20% and 12% of their initial intensities at room temperature T = 303 K, implying that the present prepared ceramic should be a promising candidate for temperature sensitive devices, i.e., non-contact optical temperature sensing.

 figure: Fig. 8.

Fig. 8. a) Temperature dependent I656/I547 related to Ho3+/Yb3+ codoped 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 ceramic; b) The ceramic's relative sensitivity versus absolute temperature.

Download Full Size | PDF

Figure 8(a) presents the calculated fluorescence intensity ratios (FIR, i.e., I656/I547) of red to green emission peaks for 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ ceramic to investigate the potential uses of as-prepared ceramics in optical temperature sensors. Previous researches have shown that these two emissions of the activators are non-TCLs when Ho3+ is in a matrix with low phonon energy [31,53]. As the temperature rises, the FIR value gradually improves. It reaches maximum value when the temperature T = 483 K. According to the experimental data, I656/I547 = 0.00298*T-0.18733 is acquired by fitting, which is similar to the reported 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: Ho3+ [31] and KBaYSi2O7: Yb3+-Er3+/Ho3+ phosphors [53].

It is critical to investigate the sensor sensitivity of the sintered ceramics. The calculation equations for sensitivity are as follows. [54,55]

$${S_A} = \frac{{\textrm{d(FIR)}}}{{dT}}$$
$${S_R} = \frac{1}{{FIR}}\frac{{d(FIR)}}{{dT}}$$
Where SA is absolute sensitivity, SR donates elative temperature sensitivity. The SA was calculated to be 0.00298 K-1 in the 303-483 K range using Eq. (5). In the same temperature range, Fig. 8(b) illustrates the corresponding computed findings of SR versus absolute temperature in accordance with Eq. (6). The SR gradually decreases as the temperature rises from 303 K to 483 K with the highest value of SR is 0.0042 K-1 at T = 303 K, which is 20% higher than that of Ho3+ single-doped 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 ceramics related to our previous study. To assess the applicability of 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ as an optical temperature sensing material by FIR technique, the maximum SR values for the as-prepared specimens and the reported other matrix materials doped with Ho3+ doped or codoped with Ho3+/Yb3+ are listed in Table 1, indicating that the SR value of 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ is shown to be greater than many other reported Ho3+ doped or Ho3+/Yb3 + codoped upconversion luminescence materials, which further validates the suitability of this material for the applications in non-contact optical temperature sensing.

Tables Icon

Table 1. Comparative value table for the maximum SR of different samples doped with Ho3+ ions.

To reveal the precise upconversion luminescence color visually for 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ ceramic at room temperature and better understanding of the temperature dependence of luminescent crystal colors, Fig. 9 provides the CIE 1931 chromaticity diagram and the calculated coordinates of 0.0075Ho3+/0.01Yb3+ solid solution based on the upconversion luminescence spectrums at different temperature captured in Fig. 7. The relevant data of chromaticity coordinates (x, y) of CIE are summarized in Table 2, indicating that the ceramic exhibits intense green emission with the color coordinates of (0.3426, 0.6477) at room temperature T = 303 K as observed in Fig. 9. Moreover, the luminous color turns from green to yellow-green when the specimen is heated from 303 to 483 K. These findings are very comparable to those of Ho3+-doped NBT ceramics [62], which suggests the upconversion luminescence color has a very significant temperature dependence. Since the human eye is extremely sensitive to green light, the current samples can also be utilized as a safety sign in low temperature conditions in addition to non-contact optical thermometry.

 figure: Fig. 9.

Fig. 9. The temperature-dependent chromaticity coordinates of the 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ specimen in CIE-1931 chromaticity diagram.

Download Full Size | PDF

Tables Icon

Table 2. The color coordinates and color purity of the 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ ceramic.

The color purity of materials plays a dominant role in the applications of emission display. Commonly, it is calculated using the following formula [62]

$$\textrm{Color purity} = \frac{{\sqrt {{{(x - {x_i})}^2} + {{(y - {\textrm{y}_i})}^2}} }}{{\sqrt {{{({x_d} - {x_i})}^2} + {{({y_d} - {\textrm{y}_i})}^2}} }} \times 100\%$$
Where (x, y), (xi, yi) and (xd, yd) denote the chromaticity coordinates of the overall light emitted by ceramic with the best luminous performance, are the write illuminant (0.3101, 0.3162), and the dominant wavelength (547 nm) in the color space, respectively. According to upconversion luminescence spectra, Table 2 also shows the estimated values for color purity at different temperatures for the 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ ceramic. The color purity gradually decreases from 82.17% to 74.08% when the temperature T is enhanced from 303 K to 4 83 K, which means that the ceramic may have good applications in WLEDs.

4. Conclusions

In summary, Ho3+/Yb3+ codoped 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 compounds have been fabricated via the traditional solid-state reaction strategy. The XRD patterns validate that in the synthesized 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: xHo3+/0.01Yb3+ ceramics, rhombohedral and tetragonal phases with perovskite structure coexist. The impacts of Ho3+ content on the performance of upconversion luminescence are thoroughly investigated, and the optimum doping concentration of 0.0075 for this system is demonstrated. The upconversion luminescence spectra exhibit an intense green emission with three typical emission bands upon NIR excitation at 980 nm, corresponding to 5F4/5S25I8, 5F55I8 and 5F4/5S25I7 characteristic transitions of Ho3+, respectively. The thermal quenching behavior of 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ ceramic indicates that the upconversion luminescence intensity is extremely temperature sensitive in the 303-483 K range. The FIR values of I656/I547 gradually increases with elevating temperature and reaches a maximum at 483 K. At 303 K, the maximum relative sensitivity is around 0.0042 K-1. The high upconversion luminescence intensity and thermal sensitivity of the current ferroelectric ceramics suggest that they could be as a promising candidate material for operation in non-contact optical temperature thermometry.

Funding

Natural Science Foundation of Huaian (Grant No. HAB202056); Program for Jiangsu Excellent Scientific and Technological Innovation Team (Grant No. 202110323094Y).

Acknowledgments

This work was financially supported by the Undergraduate Training Program for Innovation and Entrepreneurship of Jiangsu Province (Grant No. 202110323094Y), the Natural Science Foundation of Huaian (Grant No. HAB202056), and Qing Lan Project of Universities in Jiangsu Province (for Jiang).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. C. M. Lau, X. W. Xu, and K. W. Kwok, “Photoluminescence, ferroelectric, dielectric and piezoelectric properties of Er-doped BNT–BT multifunctional ceramics,” Appl. Surf. Sci. 336, 314–320 (2015). [CrossRef]  

2. D. Li, D. Zhou, W. Liu, P. J. Wang, Y. Guo, X.-G. Yao, and H.-X. Lin, “Enhanced energy storage properties achieved in Na0.5Bi0.5TiO3-based ceramics via composition design and domain engineering,” Chem. Eng. J. 419, 129601 (2021). [CrossRef]  

3. W. Li, J. Hao, W. Li, J. Du, P. Fu, W. Sun, C. Chen, Z. Xu, and R. Chu, “Electrical properties and luminescence properties of 0.96(K0.48Na0.52)(Nb0.95Sb0.05)–0.04Bi0.5(Na0.82K0.18)0.5ZrO3-xSm lead-free ceramics,” J. Adv. Ceram. 9(1), 72–82 (2020). [CrossRef]  

4. J. Gong, P. Du, W. Li, L. Luo, and Structure, “Electrical and luminescent properties of the NBT:Er-0.3CT: Pr composite ceramics,” J. Eur. Ceram. Soc. 41(9), 4815–4825 (2021). [CrossRef]  

5. M. Li, M. J. Pietrowski, R.A. De Souza, H. Zhang, I. M. Reaney, S. N. Cook, J. A. Kilner, and D. C. Sinclair, “A family of oxide ion conductors based on the ferroelectric perovskite Na0.5Bi0.5TiO3,” Nat. Mater. 13(1), 31–35 (2014). [CrossRef]  

6. M. H. Lee, D. J. Kim, J. S. Park, S. W. Kim, T. K. Song, M.-H. Kim, W.-J. Kim, D. Do, and I.-K. Jeong, “High-performance lead-free piezoceramics with high curie temperatures,” Adv. Mater. 27(43), 6976–6982 (2015). [CrossRef]  

7. M. Pal, A. Srinivas, and S. Asthana, “Enhanced magneto-electric properties and magnetodielectric effect in lead-free (1-x)0.94Na0.5Bi0.5TiO3-0.06BaTiO3–xCoFe2O4 particulate composites,” J. Alloys Compd. 900, 163487 (2022). [CrossRef]  

8. C. M. Lau, X. Wu, and K. W. Kwok, “Effects of vacancies on luminescence of Er-doped 0.93Bi0.5Na0.5TiO3-0.07 BaTiO3 ceramics,” J. Appl. Phys. 118(3), 034107 (2015). [CrossRef]  

9. S. Saïd, P. Marchet, T. Merle-MéJean, and J.-P. Mercurio, “Raman spectroscopy study of the Na0.5Bi0.5TiO3–PbTiO3 system,” Mater. Lett. 58(9), 1405–1409 (2004). [CrossRef]  

10. M. Wu, Y. Lu, and Y. Li, “Structural and electrical properties of Er2O3-doped Na1/2Bi1/2TiO3 lead-free piezoceramics,” J. Am. Ceram. Soc. 90(11), 3642–3645 (2007). [CrossRef]  

11. B. J. Chu, D. R. Chen, G. R. Li, and Q. R. Yin, “Electrical properties of Na1/2Bi1/2TiO3–BaTiO3 ceramics,” J. Eur. Ceram. Soc. 22(13), 2115–2121 (2002). [CrossRef]  

12. X. Xia, X. Jiang, C. Chen, X. Jiang, Y. Chen, N. Tu, Y. Luo, and Y. Jiang, “Enhanced piezoelectric performance and orange-red emission of Sm3+ doped (Na1/2Bi1/2)TiO3 based lead-free ceramics,” Ceram. Int. 43(1), 376–384 (2017). [CrossRef]  

13. H. Sun, X. Wu, D. F. Peng, and K. W. Kwok, “Room-temperature large and reversible modulation of photoluminescence by in situ electric field in ergodic relaxor ferroelectrics,” ACS Appl. Mater. Interfaces 9(39), 34042–34049 (2017). [CrossRef]  

14. D. K. Khatua, A. Kalaskar, and R. Ranjan, “Tuning Photoluminescence response by electric field in electrically soft ferroelectrics,” Phys. Rev. Lett. 116(11), 117601 (2016). [CrossRef]  

15. Z. Cao, J. Wang, C. Zhang, X. Mao, and L. Luo, “Flexible piezoelectric nanogenerator based on the Er3+ doped lead-free (Na0.5Bi0.5)TiO3-BaTiO3 piezoelectric nanofibers with strong upconversion luminescence,” J. Alloys Compd. 918, 165766 (2022). [CrossRef]  

16. D. K. Khatua, A. Agarwal, N. Kumar, and R. Ranjan, “Probing local structure of the morphotropic phase boundary composition of Na0.5Bi0.5TiO3–BaTiO3 using rare-earth photoluminescence as a technique,” Acta Mater. 145, 429–436 (2018). [CrossRef]  

17. Y. Wang, Y. Wang, C. Ma, Z. Feng, C. Zuo, W. Ye, C. Zhao, Y. Li, Z. Wen, Z. Cao, Z. Cao, X. Shen, C. Wang, Y. Li, X. Yuan, and Y. Cao, “Excellent cryogenic optical thermometry based on green up-conversion of Ho3+-doped perovskite Na0.5Bi0.5TiO3 ceramics,” J. Mater. Chem. C 9(4), 1353–1361 (2021). [CrossRef]  

18. P. Singh, R. S. Yadav, P. Singh, and S. B. Rai, “Upconversion and downshifting emissions of Ho3+-Yb3+ co-doped ATiO3 perovskite phosphors with temperature sensing properties in Ho3+-Yb3+ co-doped BaTiO3 phosphor,” J. Alloys Compd. 855, 157452 (2021). [CrossRef]  

19. C. Chen, X. Xia, L. Zheng, L. Luo, N. Tu, Z. Shen, X. Jiang, and J. Liu, “Effects of phase structure on up-conversion photoluminescence and dielectric performance in Er3+ doped (Bi0. 5Na0. 5)TiO3-BaTiO3 lead-free ceramics,” J. Alloys Compd. 801, 619–625 (2019). [CrossRef]  

20. K. L. Reddy, R. Balaji, A. Kumar, and V. Krishnan, “Lanthanide doped near infrared active upconversion nanophosphors: fundamental concepts, synthesis strategies, and technological applications,” Small 14, 1801304 (2018). [CrossRef]  

21. T. Zheng, M. Runowski, I. R. Martín, S. Lis, M. Vega, and J. Llanos, “Nonlinear optical thermometry—A novel temperature sensing strategy via second harmonic generation (SHG) and upconversion luminescence in BaTiO3: Ho3+,Yb3+ Perovskite,” Adv. Opt. Mater. 9(12), 2100386 (2021). [CrossRef]  

22. N. Vijaya, P. Babu, V. Venkatramu, C. K. Jayasankar, S. F. León-Luis, U. R. Rodríguez-Mendoza, I. R. Martín, and V. Lavín, “Optical characterization of Er3+-doped zinc fluorophosphate glasses for optical temperature sensors,” Sens. Actuators, B 186, 156–164 (2013). [CrossRef]  

23. Y. Zhang, X. Wang, H. Ye, Y. Li, and X. Yao, “xLiNbO3-(1-x)(K,Na)NbO3 ceramics: A new class of phosphors with tunable upconversion luminescence by external electric field and excellent optical temperature sensing property,” J. Alloys Compd. 770, 214–221 (2019). [CrossRef]  

24. H. Suo, Q. Zhu, X. Zhang, B. Chen, J. Chen, and F. Wang, “High-security anti-counterfeiting through upconversion luminescence,” Mater. Today Phys. 21, 100520 (2021). [CrossRef]  

25. F. Mangini, M. Ferraro, M. Zitelli, A. Niang, A. Tonello, V. Couderc, and S. Wabnitz, “Multiphoton-absorption-excited up-conversion luminescence in optical fibers,” Phys. Rev. Appl. 14(5), 054063 (2020). [CrossRef]  

26. O. A. Savchuk, J. J. Carvajal, M. C. Pujol, E. W. Barrera, J. Massons, M. Aguilo, and F. Diaz, “Ho, Yb: KLu(WO4)2 Nanoparticles: A versatile material for multiple thermal sensing purposes by luminescent thermometry,” J. Phys. Chem. C 119(32), 18546–18558 (2015). [CrossRef]  

27. L. Xu, J. Liu, L. Pei, Y. Xu, and Z. Xia, “Enhanced up-conversion luminescence and optical temperature sensing in graphitic C3N4 quantum dots grafted with BaWO4:Yb3+, Er3+ phosphors,” J. Mater. Chem. C 7(20), 6112–6119 (2019). [CrossRef]  

28. X. Zhao, H. Suo, Z. Zhang, and C. Guo, “Cross-relaxation and non-steady-state processes induced up-conversion color modulation in Gd2O3: Ln3+/Er3+ (Ln = Tm, Ho),” Dyes Pigm. 151, 335–341 (2018). [CrossRef]  

29. B. Hou, M. Jia, P. Li, G. Liu, Z. Sun, and Z. Fu, “Multifunctional optical thermometry based on the rare-earth-ions-doped up-/down-Conversion Ba2TiGe2O8: Ln (Ln = Eu3+/Er3+/Ho3+/Yb3+) Phosphors,” Inorg. Chem. 58(12), 7939–7946 (2019). [CrossRef]  

30. Y. Wang, C. Xu, C. Zheng, X. Wang, L. Gong, and L. Sun, “Ho3+/Yb3+ codoped Vernier phase Y7O6F9 powders with efficient near-infrared quantum cutting, intense upconversion and downshifting from visible to mid-infrared,” Infrared Phys. Technol. 104, 103004 (2020). [CrossRef]  

31. J. Wu, C. Ma, Y. Fan, B. Sun, Y. Sheng, Z. Zhai, H. Lu, and X. Lv, “Upconversion luminescence and non-contact optical temperature sensing in Ho3+-doped 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 lead-free ceramics,” Opt. Mater. 127, 112239 (2022). [CrossRef]  

32. X. Wu, J. Lin, P. Chen, C. Liu, M. Lin, C. Lin, L. Luo, and X. Zheng, “Ho3+ doped (K, Na)NbO3 based multifunctional transparent ceramics with superior optical temperature sensing performance,” J. Am. Ceram. Soc. 102(3), 1249–1258 (2019). [CrossRef]  

33. J. Lin, Q. Lu, J. Xu, X. Wu, C. Lin, T. Lin, C. Chen, and L. Luo, “Outstanding optical temperature sensitivity and dual-mode temperature-dependent photoluminescence in Ho3+-doped (K,Na)NbO3–SrTiO3 transparent ceramics,” J. Am. Ceram. Soc. 102(8), 4710–4720 (2019). [CrossRef]  

34. Y. Jiang, Y. Tong, S. Chen, W. Zhang, F. Hu, R. Wei, and H. Guo, “A three-mode self-referenced optical thermometry based on up-conversion luminescence of Ca2MgWO6:Er3+,Yb3+ phosphors,” Chem. Eng. J. 413, 127470 (2021). [CrossRef]  

35. S. Wang, K. Zhu, T. Wang, H. Zhou, N. Zhang, J. Zhang, L. Ye, Z. Jia, and X. Tao, “Sensitive Ho3+, Yb3+ co-doped mixed sesquioxide single crystal fibers thermometry based on upconversion luminescence,” J. Alloys Compd. 891, 162062 (2022). [CrossRef]  

36. N. An, H. Zhou, K. Zhu, L. Ye, J. Qiu, and L.-G. Wang, “Improved temperature sensing performance of YAG: Ho3+/Yb3+ by doping Ce3+ ions based on up-conversion luminescence,” J. Alloys Compd. 843, 156057 (2020). [CrossRef]  

37. J. Zhang, Y. Zhang, and X. Jiang, “Investigations on upconversion luminescence of K3Y(PO4)2: Yb3+-Er3+/Ho3+/Tm3+ phosphors for optical temperature sensing,” J. Alloys Compd. 748, 438–445 (2018). [CrossRef]  

38. R. D. Shannon, “Revised effective ionic radii and systematic studies of interatomie distances in halides and chaleogenides,” Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr. 32(5), 751–767 (1976). [CrossRef]  

39. Z. Xu, X. Hao, and S. An, “Dielectric properties and energy-storage performance of (Na0.5Bi0.5)TiO3–SrTiO3 thick films derived from polyvinylpyrrolidone-modified chemical solution,” J. Alloys Compd. 639, 387–392 (2015). [CrossRef]  

40. C. Yan, X. Liu, C. Yue, H. Sun, H. Sui, and Y. Wang, “Outstanding enhanced breakdown field strength and energy storage properties in Na0.5Bi0.5TiO3-based thin film by the aging process,” J. Power Sources 508, 230331 (2021). [CrossRef]  

41. A. Roy, A. Dwivedi, H. Mishra, D. Kumar, and S. B. Rai, “Dual mode (upconversion and downshifting) behavior of Ho3+/Yb3+/Bi3+ co-doped YTaO4 phosphor and its application as a security ink,” J. Alloys Compd. 821, 153202 (2020). [CrossRef]  

42. M. Cai, B. Zhou, F. Wang, Y. Tian, J. Zhou, S. Xu, and J. Zhang, “Highly efficient mid-infrared 2 µm emission in Ho3+/Yb3+-codoped germanate glass,” Opt. Mater. Express 5(6), 1431–1439 (2015). [CrossRef]  

43. B. K. Das, S. J. Bora, M. Chakrabortty, L. Kalita, R. Chakrabarty, and R. Barman, “Structural, thermal and spectroscopic properties of supramolecular coordination solids,” J. Chem. Sci. 118(6), 487–494 (2006). [CrossRef]  

44. J. Tauc and A. Menth, “States in the gap,” J. Non-Cryst. Solids 8, 569–585 (1972). [CrossRef]  

45. R. S. Monika, A. Yadav, S. B. Rai, and Rai, “NIR light guided enhanced photoluminescence and temperature sensing in Ho3+/Yb3+/Bi3+ co-doped ZnGa2O4 phosphor,” Sci. Rep. 11(1), 4148 (2021). [CrossRef]  

46. T. Li, C.-F. Guo, Y.-M. Yang, L. Li, and N. Zhang, “Efficient green up-conversion emission in Yb3+/Ho3+ co-doped CaIn2O4,” Acta Mater. 61(19), 7481–7487 (2013). [CrossRef]  

47. W. Xu, X. Gao, L. Zheng, Z. Zhang, and W. Cao, “Short-wavelength upconversion emissions in Ho3+/Yb3+ codoped glass ceramic and the optical thermometry behavior,” Opt. Express 20(16), 18127–18137 (2012). [CrossRef]  

48. Z. Liu, H. Deng, and D. Chen, “Temperature dependent upconversion properties of Yb3+: Ho3+ co-doped Gd2O3 nanoparticles prepared by pulsed laser ablation in water,” Ceram. Int. 45(10), 13235–13241 (2019). [CrossRef]  

49. D. R. G. M. Pollnau, S. R. Lüthi, and H. U. Güdel, “Power dependence of upconversion luminescence in lanthanide and transition-metal-ion systems,” Phys. Rev. B 61(5), 3337–3346 (2000). [CrossRef]  

50. C. Ma, W. Zhou, Z. Gan, X. Wang, W. Tan, Z. Zhang, and Z. Zhai, “Upconversion photoluminescence modulation by electric field poling in Er3+ doped (Ba0.85Ca0.15) (Zr0.1Ti0.9)O3 piezoelectric ceramics,” J. Alloys Compd. 794, 325–332 (2019). [CrossRef]  

51. A. M. Balakt, C. P. Shaw, and Q. Zhang, “The decrease of depolarization temperature and the improvement of pyroelectric properties by doping Ta in lead-free 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 ceramics,” Ceram. Int. 43(4), 3726–3733 (2017). [CrossRef]  

52. J. Zhang, G. Chen, and Z. Hua, “Up-conversion luminescence of novel Yb3+/Ho3+/Er3+ doped Sr5(PO4)3Cl phosphors for optical temperature sensing,” Opt. Mater. Express 7(6), 2084–2089 (2017). [CrossRef]  

53. J. Zhang, Y. Zhang, X. Li, Z. Zhai, and Z. Hua, “KBaYSi2O7: Yb3+-Er3+/Ho3+ phosphors: Optical temperature sensing materials of high sensitivity,” J. Lumin. 227, 117562 (2020). [CrossRef]  

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

55. S. F. León-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 174, 176–186 (2012). [CrossRef]  

56. J. Liao, L. Kong, M. Wang, Y. Sun, and G. Gong, “Tunable upconversion luminescence and optical temperature sensing based on non-thermal coupled levels of Lu3NbO7:Yb3+/Ho3+ phosphors,” Opt. Mater. 98, 109452 (2019). [CrossRef]  

57. Y. Guo, D. Wang, and Y. He, “Fabrication of highly porous Y2O3: Ho,Yb ceramic and its thermometric applications,” J. Alloys Compd. 741, 1158–1162 (2018). [CrossRef]  

58. A. Pandey and V. K. Rai, “Improved luminescence and temperature sensing performance of Ho3+-Yb3+-Zn2+: Y2O3 phosphor,” Dalton Trans. 42(30), 11005–11011 (2013). [CrossRef]  

59. A. Kumar, J. C. G. E. da Silva, K. Kumar, H. C. Swart, S. K. Maurya, P. Kumar, and S. P. Tiwari, “Improvement in upconversion/downshifting luminescence of Gd2O3: Ho3+/Yb3+ phosphor through Ca2+ /Zn2+ incorporation and optical thermometry studies,” Mater. Res. Bull. 112, 28–37 (2019). [CrossRef]  

60. C. Sheng, X. Li, Y. Tian, X. Wang, S. Xu, H. Yu, Y. Cao, and B. Chen, “Temperature dependence of up-conversion luminescence and sensing properties of LaNbO4: Nd3+/Yb3+/Ho3+ phosphor under 808 nm excitation,” Spectrochim. Acta, Part A 244, 118846 (2021). [CrossRef]  

61. Z. Lu, K. Li, J. Wang, and L. Luo, “Photochromic and temperature sensing properties of Ho3+-Yb3+ codoped Bi0.495-xNa0.5TiO3 ceramics,” Opt. Mater. 111, 110718 (2021). [CrossRef]  

62. U. Caldiño, A. Lira, A. N. Meza-Rocha, I. Camarillo, and R. Lozada-Morales, “Development of sodium-zinc phosphate glasses doped with Dy3+, Eu3+ and Dy3+/Eu3+ for yellow laser medium, reddish-orange and white phosphor applications,” J. Lumin. 194, 231–239 (2018). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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 (9)

Fig. 1.
Fig. 1. a) Room-temperature XRD diffraction patterns related to 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: xHo3+/0.01Yb3+ ceramics. b)-d) are the magnified XRD patterns from 31.75° to 33.25°, 39.25° to 40.75° and 46° to 47.5° in the 2θ range, respectively.
Fig. 2.
Fig. 2. a) Scanning electron microscope (SEM) images and b) EDS spectrum of 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ ceramic. c)-i) corresponding EDS elemental mapping images of Na, Ba, Yb, Ho, Bi, Ti and O in 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ ceramic, respectively.
Fig. 3.
Fig. 3. The UV–visible diffuse reflectance spectra (DRS) of typical 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: xHo3+/0.01Yb3+ ceramics (x = 0.0075 and 0.015). The inset exhibits the corresponding absorption spectra.
Fig. 4.
Fig. 4. a) The upconversion emission spectrums dependent on Ho3+ ion composition in the 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: xHo3+/0.01Yb3+ ferroelectric system, b) The upconversion emission intensity of Ho3+ ions in various wavelength ranges with the different content of Ho3+.
Fig. 5.
Fig. 5. a) When x = 0.0075, upconversion emission spectra with and without Yb3+ at pump power 0.5 W. The inset shows enlarged upconversion emission spectrum in the Yb3+ free ceramic. b) The upconversion luminescence mechanism of Ho3+/Yb3+ codoped 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 specimen.
Fig. 6.
Fig. 6. a) Pump power dependence of the upconversion luminescence spectra of the 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ sample. b) Ln(Power) dependent Ln (Intensity) for 547, 656, and 756 nm.
Fig. 7.
Fig. 7. The upconversion luminescence spectra associated to the 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.075Ho3+/0.01Yb3+ ceramic at various temperatures. The inset displays the emission intensity normalized to the emission intensity at room temperature T = 303 K.
Fig. 8.
Fig. 8. a) Temperature dependent I656/I547 related to Ho3+/Yb3+ codoped 0.94Na0.5Bi0.5TiO3-0.06BaTiO3 ceramic; b) The ceramic's relative sensitivity versus absolute temperature.
Fig. 9.
Fig. 9. The temperature-dependent chromaticity coordinates of the 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ specimen in CIE-1931 chromaticity diagram.

Tables (2)

Tables Icon

Table 1. Comparative value table for the maximum SR of different samples doped with Ho3+ ions.

Tables Icon

Table 2. The color coordinates and color purity of the 0.94Na0.5Bi0.5TiO3-0.06BaTiO3: 0.0075Ho3+/0.01Yb3+ ceramic.

Equations (6)

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

F ( R ) = ( 1 R ) 2 2 R = K S
α h v  = C ( h v  -  E g ) n
I P n
S A = d(FIR) d T
S R = 1 F I R d ( F I R ) d T
Color purity = ( x x i ) 2 + ( y y i ) 2 ( x d x i ) 2 + ( y d y i ) 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.